Next Article in Journal
Arachidin-1, a Prenylated Stilbenoid from Peanut, Induces Apoptosis in Triple-Negative Breast Cancer Cells
Previous Article in Journal
Do We Have Viable Protective Strategies against Anesthesia-Induced Developmental Neurotoxicity?
Previous Article in Special Issue
Isoform-Selective NFAT Inhibitor: Potential Usefulness and Development
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Decoding the Phosphatase Code: Regulation of Cell Proliferation by Calcineurin

Department of Biochemistry, Joint Faculty of Veterinary Science, Yamaguchi University, 1677-1 Yoshida, Yamaguchi 753-8511, Japan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(3), 1122; https://doi.org/10.3390/ijms23031122
Submission received: 27 December 2021 / Revised: 18 January 2022 / Accepted: 18 January 2022 / Published: 20 January 2022
(This article belongs to the Special Issue Updates of Calcineurin/NFAT Signaling in Human Health and Diseases)

Abstract

:
Calcineurin, a calcium-dependent serine/threonine phosphatase, integrates the alterations in intracellular calcium levels into downstream signaling pathways by regulating the phosphorylation states of several targets. Intracellular Ca2+ is essential for normal cellular physiology and cell cycle progression at certain critical stages of the cell cycle. Recently, it was reported that calcineurin is activated in a variety of cancers. Given that abnormalities in calcineurin signaling can lead to malignant growth and cancer, the calcineurin signaling pathway could be a potential target for cancer treatment. For example, NFAT, a typical substrate of calcineurin, activates the genes that promote cell proliferation. Furthermore, cyclin D1 and estrogen receptors are dephosphorylated and stabilized by calcineurin, leading to cell proliferation. In this review, we focus on the cell proliferative functions and regulatory mechanisms of calcineurin and summarize the various substrates of calcineurin. We also describe recent advances regarding dysregulation of the calcineurin activity in cancer cells. We hope that this review will provide new insights into the potential role of calcineurin in cancer development.

1. Introduction

Intracellular calcium ions (Ca2+) act as pleiotropic secondary messengers in key signaling pathways involving a variety of cellular functions. While the extracellular Ca2+ concentration is 11.5 mM, the steady-state intracellular Ca2+ concentration is kept very low at several tens of nanomoles. It is well known that the concentration of Ca2+ varies greatly in different cellular compartments, for example, intracellular calcium, 100 nM; endoplasmic reticulum (ER), 0.5–1 mM; and mitochondria, 100–200 nM [1].
The binding of a hormone or growth factor to a G protein coupled receptor (GPR) or a tyrosine kinase receptor (RTK) initiates the Ca2+ signaling cascade. The activation of such receptors transmits the signals to phospholipase C (PLC), which cleaves phosphatidylinositol 4,5-biphosphate (PIP2) to produce diacylglycerol (DAG) and inositol-1,4,5-trisphosphate (InsP3). InsP3 then binds to the InsP3 receptor (InsP3R) and stimulates the release of Ca2+ from intracellular stores, such as ER [2], and allows for the entry of Ca2+ [3,4] from the extracellular space. Spatially and temporally coordinated Ca2+ release via InsP3R is regulated by complex feedback mechanisms [1].
Changes in Ca2+ concentration trigger signal transduction, which regulates a wide range of cellular events such as gene expression, cell cycle, cell motility, autophagy, and apoptosis [1,5]. Local changes in intracellular Ca2+ diffuse through the cell and have effects on the distal sites [6]. Long-term intracellular Ca2+ increases in the mitochondria and causes the release of cytochrome c and subsequently triggers apoptosis [7]. Due to the diverse roles of Ca2+, the dysregulation of calcium homeostasis can impair cell function and trigger cell death, and by doing so, may contribute to heart disease, cancer, and mental disorders. As a result, Ca2+ signals must be tightly regulated during the cell cycle. Indeed, previous studies have characterized Ca2+ transients during the cell cycle [8,9]. For example, Pande et al. (1996) measured the intracellular levels of free cytoplasmic Ca2+ in rat fibroblasts and found that complex changes occurred in calcium levels during the cell cycle [10]. They showed that the levels of Ca2+ were the lowest at the beginning of the G1 phase but were subsequently increased at the G1/S border.
Ca2+ oscillations are narrow in the sense of periodic rise and fall, while Ca2+ concentration alterations are broad in the sense of simple Ca2+ concentration changes. To facilitate different cellular processes, intracellular downstream effectors decode the frequency, amplitude, and duration of intracellular Ca2+ oscillations [11]. The most widely known family of calcium downstream effectors is calmodulin (CaM). CaM is highly conserved among eukaryotes with four EF-hand Ca2+ binding sites. CaM is involved in the regulation of cell proliferation and cell cycle through CaM-dependent phosphorylation/dephosphorylation events [12]. When Ca2+ binds to CaM, the Ca2+/CaM complex can interact with and control target proteins such as serine/threonine phosphatase calcineurin (CaN) and the family of Ca2+/CaM-dependent protein kinases (CaMK).
The molecular mechanisms through which Ca2+/CaM regulates cell cycle progression, especially through the activation state of the cyclin-dependent kinase (CDK) complex, is of great interest, and therefore has been studied intensively. It has been shown that although all eukaryotic cells require Ca2+ signaling for cell proliferation, some transformed cells and tumor cell lines are less dependent on Ca2+ for cell growth [13,14]. Ca2+-mediated signaling pathways, therefore, play important roles in tumorigenesis, tumor progression, metastasis, and invasion [6]. In this review, we discuss the latest findings of CaN in cell proliferation and its potential as a new therapeutic target in the treatment of cancer.

2. Cell Cycle and CDK

Cell division is precisely regulated by a series of CDKs, whose activity is dependent on the binding to different cell cycle-specific cyclins [15]. Cyclins confer substrate specificity and form CDK/cyclin complexes, which activate or inactivate target proteins by phosphorylation, and orchestrate coordinated cell cycle progression. In addition to CDK association with cyclins, the activity of CDKs is tightly regulated by several mechanisms, including phosphorylation, binding of CDK inhibitors (CKIs), and subcellular localization of CDK/cyclin complexes [16].
One of the critical targets of the CDK-cyclin complex in the G1 phase is the retinoblastoma protein (Rb). In the early G1 phase, CDK4/6-cyclin D initiates the phosphorylation of Rb, and, subsequently, CDK2-cyclin E phosphorylates Rb [17,18,19,20]. Phosphorylation of Rb releases the E2F transcription factor, and promotes the expression of the genes required for cell cycle progression [21,22]. This enables cells to pass through the restriction point at the G1/S boundary and to commence the S phase. CDK2-cyclin A plays an important role in S phase progression through the phosphorylation of proteins involved in DNA replication [23,24]. As cells enter the S phase, the CDK2-cyclin A complex is activated and remains activated throughout the G2 phase [25,26]. In the late G2 phase, CDK1-cyclin B is activated, allowing for the entry of cells into mitosis [27]. During G2/M transition, CDK1-cyclin A activity is necessary for prophase initiation [28].

3. Calcium, Calmodulin, and Calmodulin-Dependent Protein Kinases in Cell Proliferation

In mammalian cells, Ca2+ is required at several different points during the cell cycle. Cells are most sensitive to the depletion of extracellular Ca2+ at early G1 and near the G1/S boundary [12,29]. In early G1, Ca2+ regulates the expression of immediate-early genes, such as FOS, JUN, and MYC. Ca2+ is also required for Rb phosphorylation at the G1/S boundary. Spontaneous Ca2+ oscillations at the G1/S phase transition have been described in synchronized immortalized cell lines. These Ca2+ oscillations are accompanied by DNA replication [30].
Once hormone receptors are activated, intracellular Ca2+ levels increase. Ca2+ regulates certain targets directly and others indirectly through Ca2+-binding proteins, such as CaM. Ca2+/CaM activates various pathways involved in the regulation of cellular processes, such as secretion, cell motility, ion homeostasis, and gene transcription. CaM also modulates a large number of intracellular enzymes, including protein kinases, protein phosphatases, phosphodiesterases, adenylyl cyclases, and ion channels [31].
CaM is required for cell cycle progression through G1, specifically the G1/S transition. Consistent with this notion, CaM antagonists block cell cycle progression in early or late G1 phases [32]. Furthermore, the exit from mitosis is sensitive to changes in the concentration of CaM [33]. Upon binding with Ca2+, CaM undergoes a major conformational change, and forms a Ca2+/CaM complex to interact with a variety of target proteins such as cellular kinases [34]. The best-studied kinases involved in Ca2+/CaM signaling are the CaMKs [35,36]. CaN and CaMK play important roles in cell cycle progression by activating or inhibiting key cell cycle regulators (Figure 1). The CaN/NFAT pathway contributes to inducing the transcription of cyclin D1 [37] and CDK4, and stabilizes cyclin D1 by dephosphorylating Thr286 of cyclin D [38]. The same pathway also regulates p21, which inhibits CDK2-cyclin E and CDK4/6-cyclin D. In colorectal cancer cells, NFATc1 represses p21 expression by binding to the c-myc promoter [39]. In contrast, in breast cancer cells, the CaN/NFAT pathway promotes p21 transcription [40]. In addition, the transcription of cyclin A, which regulates the progression of S and G2 phases, is also activated by the CaN/NFAT pathway [41]. CaMK regulates cell cycle progression from G1 to S phase [42,43], and negatively regulates the expression of p27, which is an inhibitor of CDK4-cyclin D and CDK2-cyclin E [44]. In the G2/M phase, CaMK phosphorylates and activates CDC25, leading to the dephosphorylation and activation of CDK1 [45]. As a result, the inhibition of CaMK causes cell cycle arrest at G1/S [46,47,48] or G2/M transition [49], with decreased or increased expression levels of cyclin D1 and p27, respectively. G1 arrest by CaMK inhibition also accompanies CDK4 and CDK2 inhibition [43,50,51]. Not surprisingly, CaMK appears to be highly active in tumor cells, such as myeloid leukemia, glioma, and endometrial and thyroid carcinoma, where it contributes to accelerated cell proliferation [46,52,53].

4. Characteristics of Calcineurin

CaN, also known as protein phosphatase 2 B (PP2B), is a serine/threonine protein phosphatase that is conserved in all eukaryotes [54,55,56,57]. CaN is a heterodimer composed of a catalytic subunit, calcineurin A (CnA), and a Ca2+-binding regulatory subunit, calcineurin B (CnB). In mammals, three independent genes, PPP3CA, PPP3CB, and PPP3CC, encode CnAα, CnAβ, and CnAγ, respectively. CnAα and CnAβ exhibit ubiquitous expression patterns, whereas the CnAγ expression is restricted to the testis and brain [58,59,60,61]. The CaN regulatory subunits CnB1 and CnB2 are encoded by two genes (PPP3R1 and PPP3R2, respectively). The CnB1 protein is expressed ubiquitously, whereas the CnB2 protein is specifically expressed in the testes. CnA contains an amino-terminal catalytic domain followed by a CnB-binding domain, a CaM-binding domain, and an autoinhibitory domain. CaN also contains a nuclear localization signal (NLS) in the catalytic domain and a nuclear export signal (NES) in the carboxyl terminus. The autoinhibitory domain of CnA blocks the catalytic site and the NLS [56,62]. CnB contains four EF-hand Ca2+-binding motifs and an amino-terminal myristylation site. CaN is activated by the increased intracellular Ca2+ concentration in the cell and plays essential roles in multiple signaling processes [63]. The binding of Ca2+/CaM to the regulatory domain leads to the attenuation of autoinhibition, followed by dramatic enzymatic activation. Of note, although there are multiple kinases that are regulated by CaM, and CaN is the only phosphatase directly regulated by Ca2+/CaM.
As mentioned above, CaN is widely distributed in various mammalian tissues and is particularly abundant in neural tissues [64]. However, the abundance and sub-cellular localization of CnAα and CnAβ are different. CnAα is more abundantly expressed than CnAβ; furthermore, CnAα localizes in the nucleus while CnAβ localizes in the cytoplasm [65]. Interestingly, although CaN localizes predominantly in the cytoplasm of unstimulated cells [61,66,67] in response to elevated Ca2+ concentrations, and a small portion of CaN can translocate to the nucleus and interact with the target substrates [68].

5. Mechanisms Regulating Calcineurin Activity

Several proteins have been reported to inhibit CaN, including AKAP79 (A-kinase anchoring protein-79) [69], PMCA2 (plasma membrane calcium ATPase 2) [70], CHP (calcineurin homologous protein) [71,72], Cabin/Cain [73,74], calcipressin/RCAN/DSCR/CSP [75,76,77], and FK506-binding protein (FKBP) 38 [78]. Subcellular localization of CaN is regulated by its interaction with AKAP79, a scaffold protein that anchors CaN at distinct subcellular locations, leading to the inhibition of CaN [67,69,79]. In breast cancer cells, PMCA2 interacts with and sequesters CaN in the membrane, and suppresses the activation of the CaN-NFAT pathway [70].
CHP competes with CnB to bind to CnA, and inhibits CaN activity [71,72]. While Cabin1/cain inhibits CaN by interacting with CaN in a phosphorylation-dependent manner through a binding site on CaN, which is distinct from that of the drug–immunophilin complex [73,74]. Calcipressin/RCAN/DSCR/CSP has also been identified as a CnA-binding protein that inhibits CaN activity [77,80]. A conserved peptide (FLISPPxSPP) of the calcipressin family is phosphorylated and functions as a binding site for CaN. As the expression of calcipressin is induced by CaN, it functions as a feedback inhibitor of CaN signaling. Calcipressin/RCAN/DSCR/CSP binds CaN at the same site as NFAT and other substrates, with competition for binding between these molecules being a possible regulatory mechanism [81]. FKBP38 targets BCL-2 to the mitochondria and inhibits apoptosis. The same protein also binds to and inhibits CaN, even in the absence of FK506 [78]. Furthermore, histone H1 inhibits CaMKII and CaN by blocking CaM autophsophorylation [82]. CaN is also inactivated by the oxidation of key methionine residues [83,84,85,86]. Conversely, CaN is activated by the intramolecular cleavage by two different proteases, caspase 3 and Ca2+-dependent protease calpain [87,88,89]. Several studies have identified that CaN activity is regulated by its phosphorylation. The CaMKII-mediated phosphorylation of CaN on Ser197 inhibits the CaN activity [90,91,92]. This phosphorylation is blocked by Ca2+/CaM binding to CaN. In contrast, although the auto-dephosphorylation of CaN is very slow, it can be rapidly dephosphorylated by protein phosphatase IIA [91].

6. Functions of Calcineurin/NFAT

Cyclosporine A and FK506 are well-characterized immunosuppressive agents that prevent organ transplant rejection [93,94,95]. These compounds bind tightly to endogenous cytoplasmic cyclophilin A or FKBP12, respectively. Interestingly, cyclophilin A or FKBP12, in complex with cyclosporine A or FK506, bind to CaN and block the access of the CaN substrate to the active site of the CaN [96]. This indicates the possibility that the immunosuppressive effects of these drugs could be, in part, caused by the disruption of CaN functions. Consistent with this view, it has been show that CaN inhibition by cyclosporine A or FK506 delays G1/S progression in various cell types [97,98,99,100]. Mechanistically, cyclosporin A induces the expression of the cyclin inhibitor p21 and a reciprocal reduction in cyclins A and E, leading to CDK2 inactivation [101,102].
The most studied substrates of CaN are the family of nuclear factors of activated T-cell (NFATc or NFAT) transcription factors [57,103]. Four of the five members of the NFAT protein family, NFATc1 (NFAT2), NFATc2 (NFAT1), NFATc3 (NFAT4), and NFATc4 (NFAT3), are regulated by calcium signaling. In T cells, dephosphorylation of NFAT by CaN changes the structure of the NFAT protein, exposing the NLS, which is then relocated to the nucleus to regulate the transcription of immune function-associated genes [104,105,106]. In addition to T cells, NFATs are also present in a wide variety of cells and tissues, and dephosphorylation of NFAT by CaN activates transcription in the neurons and astrocytes [107,108], and also in the heart and skeletal muscle [109,110,111,112]. Of note, NFAT has been shown to regulate cell cycle-related genes and promote cell cycle progression [103,113,114,115]. Furthermore, both CaN and NFATc1 regulate cyclin D1 transcription [37]. In addition, Camp-responsive element binding protein 1 (CREB1) transcription factor, which binds to the cyclin D1 promoter and induces cyclin D1 mRNA expression, is also regulated by CaN [98]. The other key target of NFATc1 in cell cycle progression is c-Myc. CaN-mediated dephosphorylation of NFATc1 activates MYC transcription, allowing the cell to proceed to the S phase [116,117]. NFATc1 binds directly to the NFAT site in the MYC promoter [115]. In addition, NFATc1 increases c-myc transcription by activating the ERK1/2/p38/MAPK signaling pathway [118] or inducing histone acetylation, resulting in the binding of ELK1 to the MYC promoter [119]. Thus, NFATc1 upregulates MYC transcription via multiple mechanisms.
Given the diverse expression and function of each NFAT isoform, its dysregulation is known to be associated with tumorigenesis, Alzheimer’s disease, and the development of autoimmune and inflammatory diseases. A novel therapeutic strategy for treating NFAT-related diseases is to develop new ways to selectively regulate specific NFAT isoforms [120].

7. Other Substrates of Calcineurin

Approximately 600 proteins with conserved sequences that bind to CaN have been identified [121]. Many of the known CaN substrates possess PxIxIT and/or LxVP motifs [56]. Such substrates include transcription factors, proteins involved in cell cycle and apoptosis, cytoskeletal proteins, scaffold proteins, membrane channels, and receptors (Table 1) [56,58,122,123].
Key proteins associated with tumor development such as myocyte enhancer factor 2 (MEF2), kinase suppressor of ras 2 (KSR2), DAXX, c-Jun, and nuclear factor I (NFI), are known CaN substrates [128,129,134,136,138]. Furthermore, CAN dephosphorylatesd the pro-apoptotic Bcl-2 family member BAD [126] and the apoptosis promoting factor ASK1 [125]. Dephosphorylation of another apoptosis related factor CaMKIIγ is promoted by CaN, leading to the nuclear translocation of CaMKIIγ. Nuclear translocation of TFEB, a master transcriptional regulator of lysosome biogenesis and autophagy, also requires dephosphorylated by CaN [127,145]. In addition, while DARPP-32 and ElK-1 are inactivated by CaN, other targets such as nitric oxide synthase, NHE1, and TRESK are activated [142,146,147,148,149]. CaN is activated through cytoplasmic Ca2+ increase caused by mitochondrial depolarization. Activated CaN dephosphorylates DRP1, which then relocates to the mitochondria and promotes mitochondrial fission [130]. In neurons, CaN dephosphorylates dynamin1. This promotes the endocytosis of TRKA receptors and subsequent axonal growth [131]. CaN-dependent dephosphorylation of GluA1, a component of AMPA receptors (AMPARs) at the synapses, leads to the removal of AMPARs from the synapse and endocytosis [133], while dephosphorylation of Myosin phosphatase target subunit 1 (MYPT1, a component of MP) by CaN causes MP to acquire phosphatase activity [137].
Recently, we found that CaN inhibits cyclin D1 degradation by dephosphorylation of the Thr286 residue [38] (Figure 2a). Treatment with CN585, which is a specific inhibitor of CaN phosphatase activity, or by the immunosuppressant FK506, inhibits breast cancer cell proliferation accompanied by delayed G1/S progression mediated by the degradation of cyclin D1. FK506 also decreases the protein level of CDK4 via a yet unknown mechanism. Taken together, CaN activates CDK4-cyclin D1 through multiple mechanisms to promote cell cycle progression [12,38]. The overexpression of cyclin D1 has been linked to the development and progression of cancer [150]. It is also known that increased levels of cyclin D1 could frequently result from deregulated proteasomal degradation [151]. This indicates that the selective inhibition of CaN could be an effective method for the treatment of cancers that are characterized by cyclin D1 overexpression. We also found that the stability and activity of ERα, a key molecule for estrogen-dependent cancer progression, is mediated by CaN [132] (Figure 2b). CaN dephosphorylates ERα on Ser294 and activates the mechanistic target of rapamycin (mTOR). CaN is also dephosphorylate MAP2, RIIα, RCAN1, and Tau, but the significance of these dephosphorylations remains unclear [124,135,143,144].
Intriguingly, CaN may possess roles that are independent of its dephosphorylation activity. CnAβ1 functionally activates the PDK1-Akt phosphorylation cascade, increasing the phosphorylation of its target, Ser253 of FOXO3a, and the inhibition of its nuclear translocation. The inhibitory effect of FOXO3a by CnAβ1 regulates myoblast proliferation and prevents myotube atrophy under starvation conditions. Of note, this function of CnAβ1 does not require a phosphatase activity, suggesting that CaN functions independently of its dephosphorylation activity [152].

8. Activation of Calcineurin/NFAT Pathway in Cancer

The CaN/NFAT pathway is activated in multiple cancers, including breast cancer [153], lung cancer [154], prostate cancer [155], ovarian cancer [118], pancreatic cancer [115], liver cancer [156], colorectal cancer [39], glioblastoma [138,157], melanoma [158], leukemia [159,160], and lymphoma [161,162]. The dysregulated CaN/NFAT pathway observed in cancer is summarized in Table 2.
The mechanisms that drive CaN activation in cancer are not well understood. In some cancers, CaN hyperactivation has been associated with mutations. For example, the EL4 murine T-cell lymphoma cell line expresses a mutant form of CaN, in which the aspartic acid at position 477 is mutated to asparagine and the negative regulation of the phosphatase activity by the autoinhibitory domain is impaired by this mutation [167]. In contrast, the SML B-cell lymphoma cell line expresses a truncated version of CnA, which results in constitutive activation of CaN [168]. It has been reported that the expression of Ca2+ channels that regulate the intracellular Ca2+ concentration is involved in the activation of CaN. TRPv6 (TRP vanilloid family member 6), a Ca2+-selective ion channel, is highly expressed in prostate cancer and is a prognostic marker [169,170]. Increased expression of TRPv6 induces more Ca2+ to enter the cell, which in turn promotes NFAT activation [171]. It is also possible that the inflammatory response activates CaN [172,173,174]. Consistent with this notion, in intestinal cancer, changes in the bacterial community activate Toll-like receptor (TLR) signaling, which leads to subsequent calcium entry and CaN activation [175]. Moreover, inflammatory responses are activated in many cancers, and it is well understood that chronic inflammation is a risk factor for tumorigenesis [176]. Recently, we found that a high expression of CaN was correlated with a poor prognosis regarding the outcome of endocrine therapy in patients with ERα-positive breast cancer [132], which indicates that the selective inhibition of CaN could be effective to treat such cancers.
Highly activated CaN is supposed to activate its substrates, including NFAT, to promote proliferation. Similar to CaN, its target NFAT is also constitutively activated or overexpressed in numerous cancers and may contribute to cancer development and progression [177]. Conversely, the expression of NFATc1 is suppressed in several types of cancer, and a reduction of NFATc1 has been shown to be linked with aggressiveness and malignancy of cancer [164,165,166].

9. A Therapeutic Perspective for Cancer

Owing to the high frequency of CaN/NFAT activation in cancer and the contribution of these molecules in cancer progression, the CaN/NFAT pathway could be a potential therapeutic target. Indeed, the anticancer effects of CaN inhibitors have been studied extensively in the past. For instance, cyclosporine A or FK506 induced apoptosis and rapid tumor clearance, resulting in the regression of leukemia [160]. Cyclosporine A or FK506 also inhibits tumor growth in the bladder and prostate xenografts in vivo [174,178,179]. In addition, cyclosporine A itself is also directly involved in tumor growth, as it increases TGFβ production [180], activates Ras [181], suppresses PTEN expression, and increases AKT activation [182].
There is growing acceptance that targeting dysregulated Ca2+ channels/transporters/pumps can provide promising potential for the treatment of cancer patients. [183]. Buffering nuclear Ca2+ concentrations alters the expression levels of the genes involved in cell proliferation, resulting in an antitumor effect [184]. Moreover, buffering nuclear Ca2+ reduces the growth rate of tumor cells without affecting the cells in normal tissues [185]. Further research is needed to clarify the differences in the sensitivity of normal and cancer cell growth to changes in nuclear Ca2+ levels.

10. Conclusions

Nuclear Ca2+ is involved in tumor growth and alters the expression of the genes involved in cell proliferation. Furthermore, previous studies have suggested that the modulation of nuclear Ca2+ signaling may be effective in cancer therapy. Activation of CaN and its downstream dephosphorylation has been identified as a mechanism by which nuclear Ca2+ regulates cell proliferation and cell cycle progression. As discussed above, dephosphorylation of proteins by CaN plays an important role in tumor formation and progression. Therefore, in the future, it will be necessary to identify the molecular mechanism of CaN activation and the substrates that promote cancer cell growth. Targeting the interactions of activated CaN with specific substrates in cancer cells, without affecting the normal immune function of CaN, may effectively inhibit the growth of cancer cells, leading to the establishment of new tumor-specific therapies.

Author Contributions

Conceptualization, M.S.; investigation, T.M. and M.S.; data curation, T.M. and M.S.; writing—original draft preparation, M.S.; writing—review and editing, M.S.; visualization, T.M. and M.S.; supervision, M.S.; project administration, T.M. and M.S.; funding acquisition, M.S. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by a Grant-in-Aid for JSPS KAKENHI (grant numbers: 21H02403 and 20K21503) and the Yamaguchi University Project for Formation of the Core Research Center to Midori Shimada.

Acknowledgments

The authors thank the members of Shimada Laboratory.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Berridge, M.J.; Lipp, P.; Bootman, M.D. The versatility and universality of calcium signalling. Nat. Rev. Mol. Cell Biol. 2000, 1, 11–21. [Google Scholar] [CrossRef]
  2. Patel, S.; Joseph, S.; Thomas, A. Molecular properties of inositol 1,4,5-trisphosphate receptors. Cell Calcium 1999, 25, 247–264. [Google Scholar] [CrossRef] [PubMed]
  3. Kiselyov, K.; Mignery, G.A.; Zhu, M.X.; Muallem, S. The N-Terminal Domain of the IP3 Receptor Gates Store-Operated hTrp3 Channels. Mol. Cell 1999, 4, 423–429. [Google Scholar] [CrossRef]
  4. Van Rossum, D.; Patterson, R.L.; Kiselyov, K.; Boehning, D.; Barrow, R.K.; Gill, D.L.; Snyder, S.H. Agonist-induced Ca2+ entry determined by inositol 1,4,5-trisphosphate recognition. Proc. Natl. Acad. Sci. USA 2004, 101, 2323–2327. [Google Scholar] [CrossRef] [Green Version]
  5. Venkatachalam, K.; Van Rossum, D.B.; Patterson, R.L.; Ma, H.-T.; Gill, D.L. The cellular and molecular basis of store-operated calcium entry. Nat. Cell Biol. 2002, 4, E263–E272. [Google Scholar] [CrossRef] [PubMed]
  6. Parkash, J.; Asotra, K. Calcium wave signaling in cancer cells. Life Sci. 2010, 87, 587–595. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. McConkey, D.J.; Orrenius, S. The role of calcium in the regulation of apoptosis. Biochem. Biophys. Res. Commun. 1997, 239, 357–366. [Google Scholar] [CrossRef]
  8. Santella, L. The Role of Calcium in the Cell Cycle: Facts and Hypotheses. Biochem. Biophys. Res. Commun. 1998, 244, 317–324. [Google Scholar] [CrossRef]
  9. Pinto, M.C.X.; Kihara, A.; Goulart, V.A.; Tonelli, F.M.P.; Gomes, K.N.; Ulrich, H.; Resende, R.R. Calcium signaling and cell proliferation. Cell. Signal. 2015, 27, 2139–2149. [Google Scholar] [CrossRef]
  10. Pande, G.; Kumar, N.A.; Manogaran, P.S. Flow cytometric study of changes in the intracellular free calcium during the cell cycle. Cytometry 1996, 24, 55–63. [Google Scholar] [CrossRef]
  11. Smedler, E.; Uhlén, P. Frequency decoding of calcium oscillations. Biochim. Biophys. Acta (BBA) Gen. Subj. 2014, 1840, 964–969. [Google Scholar] [CrossRef] [Green Version]
  12. Kahl, C.R.; Means, A.R. Regulation of Cell Cycle Progression by Calcium/Calmodulin-Dependent Pathways. Endocr. Rev. 2003, 24, 719–736. [Google Scholar] [CrossRef] [Green Version]
  13. Whitfield, J.F. Calcium signals and cancer. Crit. Rev. Oncog. 1992, 3, 55–90. [Google Scholar]
  14. Cook, S.J.; Lockyer, P.J. Recent advances in Ca2+-dependent Ras regulation and cell proliferation. Cell Calcium 2006, 39, 101–112. [Google Scholar] [CrossRef]
  15. Morgan, D.O. Cyclin-Dependent Kinases: Engines, Clocks, and Microprocessors. Annu. Rev. Cell Dev. Biol. 1997, 13, 261–291. [Google Scholar] [CrossRef]
  16. Satyanarayana, A.; Kaldis, P. Mammalian cell-cycle regulation: Several Cdks, numerous cyclins and diverse compensatory mechanisms. Oncogene 2009, 28, 2925–2939. [Google Scholar] [CrossRef] [Green Version]
  17. Matsushime, H.; Ewen, M.E.; Strom, D.K.; Kato, J.Y.; Hanks, S.K.; Roussel, M.F.; Sherr, C.J. Identification and properties of an atypical catalytic subunit (p34PSK-J3/cdk4) for mammalian D type G1 cyclins. Cell 1992, 71, 323–334. [Google Scholar] [CrossRef]
  18. Connell-Crowley, L.; Harper, J.W.; Goodrich, D.W. Cyclin D1/Cdk4 regulates retinoblastoma protein-mediated cell cycle arrest by site-specific phosphorylation. Mol. Biol. Cell 1997, 8, 287–301. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Zarkowska, T.; Mittnacht, S. Differential phosphorylation of the retinoblastoma protein by G1/S cyclin-dependent kinases. J. Biol. Chem. 1997, 272, 12738–12746. [Google Scholar] [CrossRef] [Green Version]
  20. Bertoli, C.; Skotheim, J.M.; De Bruin, R.A.M. Control of cell cycle transcription during G1 and S phases. Nat. Rev. Mol. Cell Biol. 2013, 14, 518–528. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Weinberg, R.A. The retinoblastoma protein and cell cycle control. Cell 1995, 81, 323–330. [Google Scholar] [CrossRef] [Green Version]
  22. Dyson, N. The regulation of E2F by pRB-family proteins. Genes Dev. 1998, 12, 2245–2262. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Petersen, B.O.; Lukas, J.; Sorensen, C.; Bartek, J.; Helin, K. Phosphorylation of mammalian CDC6 by Cyclin A/CDK2 regulates its subcellular localization. EMBO J. 1999, 18, 396–410. [Google Scholar] [CrossRef]
  24. Coverley, D.; Pelizon, C.; Trewick, S.; Laskey, R. Chromatin-bound Cdc6 persists in S and G2 phases in human cells, while soluble Cdc6 is destroyed in a cyclin A-cdk2 dependent process. J. Cell Sci. 2000, 113, 1929–1938. [Google Scholar] [CrossRef] [PubMed]
  25. Yam, C.H.; Fung, T.K.; Poon, R. Cyclin A in cell cycle control and cancer. Cell. Mol. Life Sci. 2002, 59, 1317–1326. [Google Scholar] [CrossRef]
  26. Stillman, B. Cell Cycle Control of DNA Replication. Science 1996, 274, 1659–1663. [Google Scholar] [CrossRef] [PubMed]
  27. Kishimoto, T.; Okumura, E. In vivo regulation of the entry into M-phase: Initial activation and nuclear translocation of cyclin B/Cdc2. Prog. Cell Cycle Res. 1997, 3, 241–249. [Google Scholar] [CrossRef]
  28. Furuno, N.; Elzen, N.D.; Pines, J. Human Cyclin a Is Required for Mitosis until Mid Prophase. J. Cell Biol. 1999, 147, 295–306. [Google Scholar] [CrossRef] [Green Version]
  29. Boynton, A.L.; Whitfield, J.F.; Isaacs, R.J.; Tremblay, R. The control of human WI-38 cell proliferation by extracellular calcium and its elimination by SV-40 virus-induced proliferative transformation. J. Cell. Physiol. 1977, 92, 241–247. [Google Scholar] [CrossRef] [PubMed]
  30. Russa, A.D.; Maesawa, C.; Satoh, Y. Spontaneous [Ca2+]i oscillations in G1/S phase-synchronized cells. J. Electron. Microsc. 2009, 58, 321–329. [Google Scholar] [CrossRef]
  31. Chin, D.; Means, A.R. Calmodulin: A prototypical calcium sensor. Trends Cell Biol. 2000, 10, 322–328. [Google Scholar] [CrossRef]
  32. Colomer, J.; Lopezgirona, A.; Agell, N.; Bachs, O. Calmodulin Regulates the Expression of CDKS, Cyclins and Replicative Enzymes During Proliferative Activation of Human T Lymphocytes. Biochem. Biophys. Res. Commun. 1994, 200, 306–312. [Google Scholar] [CrossRef] [PubMed]
  33. Rasmussen, C.D.; Means, A.R. Calmodulin is required for cell-cycle progression during G1 and mitosis. EMBO J. 1989, 8, 73–82. [Google Scholar] [CrossRef] [Green Version]
  34. Choi, J.; Husain, M. Calmodulin-Mediated Cell Cycle Regulation: New Mechanisms for Old Observations. Cell Cycle 2006, 5, 2183–2186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Schulman, H.; Hanson, P.I. Multifunctional Ca2+/calmodulin-dependent protein kinase. Neurochem. Res. 1993, 18, 65–77. [Google Scholar] [CrossRef]
  36. Fujisawa, H. Regulation of the activities of multifunctional Ca2+/calmodulin-dependent protein kinases. J. Biochem. 2001, 129, 193–199. [Google Scholar] [CrossRef] [PubMed]
  37. Karpurapu, M.; Wang, D.; Van Quyen, D.; Kim, T.K.; Kundumani-Sridharan, V.; Pulusani, S.; Rao, G.N. Cyclin D1 is a bona fide target gene of NFATc1 and is sufficient in the mediation of injury-induced vascular wall remodeling. J. Biol. Chem. 2010, 285, 3510–3523. [Google Scholar] [CrossRef] [Green Version]
  38. Goshima, T.; Habara, M.; Maeda, K.; Hanaki, S.; Kato, Y.; Shimada, M. Calcineurin regulates cyclin D1 stability through dephosphorylation at T286. Sci. Rep. 2019, 9, 1–11. [Google Scholar] [CrossRef] [Green Version]
  39. Masuo, T.; Okamura, S.; Zhang, Y.; Mori, M. Cyclosporine A inhibits colorectal cancer proliferation probably by regulating expression levels of c-Myc, p21WAF1/CIP1 and proliferating cell nuclear antigen. Cancer Lett. 2009, 285, 66–72. [Google Scholar] [CrossRef]
  40. Perez-Neut, M.; Rao, V.R.; Gentile, S. hERG1/Kv11.1 activation stimulates transcription of p21waf/cip in breast cancer cells via a calcineurin-dependent mechanism. Oncotarget 2016, 7, 58893–58902. [Google Scholar] [CrossRef] [Green Version]
  41. Li, M.; Liu, Y.; Sun, X.; Li, Z.; Fang, P.; He, P.; Shi, H.; Xie, M.; Wang, X.; Zhang, D.; et al. Sildenafil inhibits calcineurin/NFATc2-mediated cyclin A expression in pulmonary artery smooth muscle cells. Life Sci. 2011, 89, 644–649. [Google Scholar] [CrossRef]
  42. Choi, J.; Chiang, A.; Taulier, N.; Gros, R.; Pirani, A.; Husain, M. A calmodulin-binding site on cyclin E mediates Ca2+-sensitive G1/s transitions in vascular smooth muscle cells. Circ. Res. 2006, 98, 1273–1281. [Google Scholar] [CrossRef] [Green Version]
  43. Kahl, C.R.; Means, A.R. Regulation of cyclin D1/Cdk4 complexes by calcium/calmodulin-dependent protein kinase I. J. Biol. Chem. 2004, 279, 15411–15419. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Wang, C.; Li, N.; Liu, X.; Zheng, Y.; Cao, X. A novel endogenous human CaMKII inhibitory protein suppresses tumor growth by inducing cell cycle arrest via p27 stabilization. J. Biol. Chem. 2008, 283, 11565–11574. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Patel, R.; Holt, M.; Philipova, R.; Moss, S.; Schulman, H.; Hidaka, H.; Whitaker, M. Calcium/calmodulin-dependent phosphorylation and activation of human Cdc25-C at the G2/M phase transition in HeLa cells. J. Biol. Chem. 1999, 274, 7958–7968. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Parmer, T.G.; Ward, M.D.; Hait, W.N. Effects of rottlerin, an inhibitor of calmodulin-dependent protein kinase III, on cellular proliferation, viability, and cell cycle distribution in malignant glioma cells. Cell Growth Differ. Mol. Boil. J. Am. Assoc. Cancer Res. 1997, 8, 327–334. [Google Scholar]
  47. Minami, H.; Inoue, S.; Hidaka, H. The effect of KN-62, Ca2+/calmodulin dependent protein kinase II inhibitor on cell cycle. Biochem. Biophys. Res. Commun. 1994, 199, 241–248. [Google Scholar] [CrossRef]
  48. Skelding, K.A.; Rostas, J.A.P.; Verrills, N.M. Controlling the cell cycle: The role of calcium/calmodulin-stimulated protein kinases I and II. Cell Cycle 2011, 10, 631–639. [Google Scholar] [CrossRef] [Green Version]
  49. House, S.J.; Ginnan, R.G.; Armstrong, S.E.; Singer, H.A. Calcium/calmodulin-dependent protein kinase II-delta isoform regulation of vascular smooth muscle cell proliferation. Am. J. Physiol. Cell Physiol. 2007, 292, C2276–C2287. [Google Scholar] [CrossRef] [Green Version]
  50. Tombes, R.M.; Grant, S.; Westin, E.H.; Krystal, G. G1 cell cycle arrest and apoptosis are induced in NIH 3T3 cells by KN-93, an inhibitor of CaMK-II (the multifunctional Ca2+/CaM kinase). Cell Growth Differ. Mol. Boil. J. Am. Assoc. Cancer Res. 1995, 6, 1063–1070. [Google Scholar]
  51. Morris, T.A.; DeLorenzo, R.J.; Tombes, R.M. CaMK-II inhibition reduces cyclin D1 levels and enhances the association of p27kip1 with Cdk2 to cause G1 arrest in NIH 3T3 cells. Exp. Cell Res. 1998, 240, 218–227. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Cheng, E.H.; Gorelick, F.; Czernik, A.J.; Bagaglio, D.M.; Hait, W.N. Calmodulin-dependent protein kinases in rat glioblastoma. Cell Growth Differ. Mol. Boil. J. Am. Assoc. Cancer Res. 1995, 6, 615–621. [Google Scholar]
  53. Si, J.; Collins, S.J. Activated Ca2+/calmodulin-dependent protein kinase IIgamma is a critical regulator of myeloid leukemia cell proliferation. Cancer Res. 2008, 68, 3733–3742. [Google Scholar] [CrossRef] [Green Version]
  54. Klee, C.B.; Ren, H.; Wang, X. Regulation of the Calmodulin-stimulated Protein Phosphatase, Calcineurin. J. Biol. Chem. 1998, 273, 13367–13370. [Google Scholar] [CrossRef] [Green Version]
  55. Hogan, P.G.; Li, H. Calcineurin. Curr. Biol. 2005, 15, R442–R443. [Google Scholar] [CrossRef] [Green Version]
  56. Li, H.; Rao, A.; Hogan, P.G. Interaction of calcineurin with substrates and targeting proteins. Trends Cell Biol. 2011, 21, 91–103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Rusnak, F.; Mertz, P. Calcineurin: Form and Function. Physiol. Rev. 2000, 80, 1483–1521. [Google Scholar] [CrossRef]
  58. Klee, C.B.; Draetta, G.F.; Hubbard, M.J. Calcineurin. Adv. Enzym. Relat. Areas Mol. Biol. 1988, 61, 149–200. [Google Scholar]
  59. Cohen, P.T.; Chen, M.X.; Armstrong, C.G. Novel Protein Phosphatases That May Participate in Cell Signaling. Adv. Pharmacol. 1996, 36, 67–89. [Google Scholar] [CrossRef]
  60. Goto, S.; Matsukado, Y.; Mihara, Y.; Inoue, N.; Miyamoto, E. Calcineurin in human brain and its relation to extrapyramidal system. Immunohistochemical study on postmortem human brains. Acta Neuropathol. 1986, 72, 150–156. [Google Scholar] [CrossRef] [PubMed]
  61. Kuno, T.; Mukai, H.; Ito, A.; Chang, C.D.; Kishima, K.; Saito, N.; Tanaka, C. Distinct cellular expression of calcineurin A alpha and A beta in rat brain. J. Neurochem. 1992, 58, 1643–1651. [Google Scholar] [CrossRef] [PubMed]
  62. Hallhuber, M.; Burkard, N.; Wu, R.; Buch, M.H.; Engelhardt, S.; Hein, L.; Neyses, L.; Schuh, K.; Ritter, O. Inhibition of nuclear import of calcineurin prevents myocardial hypertrophy. Circ. Res. 2006, 99, 626–635. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Park, Y.-J.; Yoo, S.-A.; Kim, M.; Kim, W.-U. The Role of Calcium–Calcineurin–NFAT Signaling Pathway in Health and Autoimmune Diseases. Front. Immunol. 2020, 11, 195. [Google Scholar] [CrossRef]
  64. Wallace, R.W.; Tallant, E.A.; Cheung, W.Y. High levels of a heat-labile calmodulin-binding protein (CaM-BP80) in bovine neostriatum. Biochemistry 1980, 19, 1831–1837. [Google Scholar] [CrossRef]
  65. Usuda, N.; Arai, H.; Sasaki, H.; Hanai, T.; Nagata, T.; Muramatsu, T.; Kincaid, R.L.; Higuchi, S. Differential subcellular localization of neural isoforms of the catalytic subunit of calmodulin-dependent protein phosphatase (calcineurin) in central nervous system neurons: Immunohistochemistry on formalin-fixed paraffin sections employing antigen retrieval by microwave irradiation. J. Histochem. Cytochem. 1996, 44, 13–18. [Google Scholar] [CrossRef] [PubMed]
  66. Anthony, F.A.; Winkler, M.A.; Edwards, H.H.; Cheung, W.Y. Quantitative subcellular localization of calmodulin-dependent phosphatase in chick forebrain. J. Neurosci. 1988, 8, 1245–1253. [Google Scholar] [CrossRef] [Green Version]
  67. Natarajan, K.; Ness, J.; Wooge, C.H.; Janovick, J.A.; Conn, P.M. Specific Identification and Subcellular Localization of Three Calmodulin-binding Proteins in the Rat Gonadotrope: Spectrin, Caldesmon, and Calcineurin. Biol. Reprod. 1991, 44, 43–52. [Google Scholar] [CrossRef] [Green Version]
  68. Shibasaki, F.; Price, E.R.; Milan, D.; McKeon, F. Role of kinases and the phosphatase calcineurin in the nuclear shuttling of transcription factor NF-AT4. Nature 1996, 382, 370–373. [Google Scholar] [CrossRef]
  69. Coghlan, V.M.; Perrino, B.A.; Howard, M.; Langeberg, L.K.; Hicks, J.B.; Gallatin, W.M.; Scott, J.D. Association of Protein Kinase A and Protein Phosphatase 2B with a Common Anchoring Protein. Science 1995, 267, 108–111. [Google Scholar] [CrossRef]
  70. Baggott, R.R.; Mohamed, T.M.; Oceandy, D.; Holton, M.; Blanc, M.C.; Roux-Soro, S.C.; Brown, S.; Brown, J.E.; Cartwright, E.; Wang, W.; et al. Disruption of the interaction between PMCA2 and calcineurin triggers apoptosis and enhances paclitaxel-induced cytotoxicity in breast cancer cells. Carcinogenesis 2012, 33, 2362–2368. [Google Scholar] [CrossRef] [Green Version]
  71. Lin, X.; Barber, D.L. A calcineurin homologous protein inhibits GTPase-stimulated Na-H exchange. Proc. Natl. Acad. Sci. USA 1996, 93, 12631–12636. [Google Scholar] [CrossRef] [Green Version]
  72. Lin, X.; Sikkink, R.A.; Rusnak, F.; Barber, D.L. Inhibition of Calcineurin Phosphatase Activity by a Calcineurin B Homologous Protein. J. Biol. Chem. 1999, 274, 36125–36131. [Google Scholar] [CrossRef] [Green Version]
  73. Sun, L.; Youn, H.-D.; Loh, C.; Stolow, M.; He, W.; Liu, J.O. Cabin 1, A Negative Regulator for Calcineurin Signaling in T Lymphocytes. Immunity 1998, 8, 703–711. [Google Scholar] [CrossRef] [Green Version]
  74. Lai, M.M.; Burnett, P.E.; Wolosker, H.; Blackshaw, S.; Snyder, S.H. Cain, A Novel Physiologic Protein Inhibitor of Calcineurin. J. Biol. Chem. 1998, 273, 18325–18331. [Google Scholar] [CrossRef] [Green Version]
  75. Rothermel, B.; Vega, R.B.; Yang, J.; Wu, H.; Bassel-Duby, R.; Williams, R.S. A Protein Encoded within the Down Syndrome Critical Region Is Enriched in Striated Muscles and Inhibits Calcineurin Signaling. J. Biol. Chem. 2000, 275, 8719–8725. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Fuentes, J.J.; Genescà, E.; Kingsbury, T.J.; Cunningham, K.W.; Perez-Riba, M.; Estivill, X.; de la Luna, S. DSCR1, overexpressed in Down syndrome, is an inhibitor of calcineurin-mediated signaling pathways. Hum. Mol. Genet. 2000, 9, 1681–1690. [Google Scholar] [CrossRef] [Green Version]
  77. Kingsbury, T.J.; Cunningham, K.W. A conserved family of calcineurin regulators. Genes Dev. 2000, 14, 1595–1604. [Google Scholar] [CrossRef] [PubMed]
  78. Shirane, M.; Nakayama, K.I. Inherent calcineurin inhibitor FKBP38 targets Bcl-2 to mitochondria and inhibits apoptosis. Nat. Cell Biol. 2003, 5, 28–37. [Google Scholar] [CrossRef]
  79. Li, H.; Pink, M.D.; Murphy, J.G.; Stein, A.; Dell’Acqua, M.L.; Hogan, P.G. Balanced interactions of calcineurin with AKAP79 regulate Ca2+–calcineurin–NFAT signaling. Nat. Struct. Mol. Biol. 2012, 19, 337–345. [Google Scholar] [CrossRef] [PubMed]
  80. Görlach, J.; Fox, D.S.; Cutler, N.S.; Cox, G.M.; Perfect, J.R.; Heitman, J. Identification and characterization of a highly conserved calcineurin binding protein, CBP1/calcipressin, inCryptococcus neoformans. EMBO J. 2000, 19, 3618–3629. [Google Scholar] [CrossRef] [Green Version]
  81. Martinez-Martinez, S.; Genescà, E.; Rodriguez, A.; Raya, A.; Salichs, E.; Were, F.; Lopez-Maderuelo, M.D.; Redondo, J.M.; de la Luna, S. The RCAN carboxyl end mediates calcineurin docking-dependent inhibition via a site that dictates binding to substrates and regulators. Proc. Natl. Acad. Sci. USA 2009, 106, 6117–6122. [Google Scholar] [CrossRef] [Green Version]
  82. Rasmussen, C.; Garen, C. Activation of calmodulin-dependent enzymes can be selectively inhibited by histone H1. J. Biol. Chem. 1993, 268, 23788–23791. [Google Scholar] [CrossRef]
  83. Wang, X.; Culotta, V.C.; Klee, C.B. Superoxide dismutase protects calcineurin from inactivation. Nature 1996, 383, 434–437. [Google Scholar] [CrossRef] [PubMed]
  84. Sommer, D.; Fakata, K.L.; Swanson, S.A.; Stemmer, P. Modulation of the phosphatase activity of calcineurin by oxidants and antioxidants in vitro. JBIC J. Biol. Inorg. Chem. 2000, 267, 2312–2322. [Google Scholar] [CrossRef] [PubMed]
  85. Namgaladze, D.; Shcherbyna, I.; Kienhöfer, J.; Hofer, H.W.; Ullrich, V. Superoxide targets calcineurin signaling in vascular endothelium. Biochem. Biophys. Res. Commun. 2005, 334, 1061–1067. [Google Scholar] [CrossRef]
  86. Zhou, X.; Mester, C.; Stemmer, P.; Reid, G.E. Oxidation-Induced Conformational Changes in Calcineurin Determined by Covalent Labeling and Tandem Mass Spectrometry. Biochemistry 2014, 53, 6754–6765. [Google Scholar] [CrossRef]
  87. Mukerjeea, N.; McGinnis, K.M.; Gnegy, M.E.; Wang, K.K. Caspase-Mediated Calcineurin Activation Contributes to IL-2 Release during T Cell Activation. Biochem. Biophys. Res. Commun. 2001, 285, 1192–1199. [Google Scholar] [CrossRef] [Green Version]
  88. Kim, M.-J.; Jo, D.-G.; Hong, G.-S.; Kim, B.J.; Lai, M.; Cho, D.-H.; Kim, K.-W.; Bandyopadhyay, A.; Hong, Y.-M.; Kim, D.H.; et al. Calpain-dependent cleavage of cain/cabin1 activates calcineurin to mediate calcium-triggered cell death. Proc. Natl. Acad. Sci. USA 2002, 99, 9870–9875. [Google Scholar] [CrossRef] [Green Version]
  89. Wu, H.-Y.; Tomizawa, K.; Oda, Y.; Wei, F.-Y.; Lu, Y.-F.; Matsushita, M.; Li, S.T.; Moriwaki, A.; Matsui, H. Critical Role of Calpain-mediated Cleavage of Calcineurin in Excitotoxic Neurodegeneration. J. Biol. Chem. 2004, 279, 4929–4940. [Google Scholar] [CrossRef] [Green Version]
  90. Hashimoto, Y.; King, M.M.; Soderling, T.R. Regulatory interactions of calmodulin-binding proteins: Phosphorylation of calcineurin by autophosphorylated Ca2+/calmodulin-dependent protein kinase II. Proc. Natl. Acad. Sci. USA 1988, 85, 7001–7005. [Google Scholar] [CrossRef] [Green Version]
  91. Hashimoto, Y.; Soderling, T.R. Regulation of calcineurin by phosphorylation. Identification of the regulatory site phosphorylated by Ca2+/calmodulin-dependent protein kinase II and protein kinase C. J. Biol. Chem. 1989, 264, 16524–16529. [Google Scholar] [CrossRef]
  92. Martensen, T.M.; Martin, B.M.; Kincaid, R.L. Identification of the site on calcineurin phosphorylated by Ca2+/CaM-dependent kinase II: Modification of the CaM-binding domain. Biochemistry 1989, 28, 9243–9247. [Google Scholar] [CrossRef] [PubMed]
  93. Liu, J.; Farmer, J.D.; Lane, W.S.; Friedman, J.; Weissman, I.; Schreiber, S.L. Calcineurin is a common target of cyclophilin-cyclosporin A and FKBP-FK506 complexes. Cell 1991, 66, 807–815. [Google Scholar] [CrossRef]
  94. Shaw, K.T.; Ho, A.M.; Raghavan, A.; Kim, J.; Jain, J.; Park, J.; Sharma, S.; Rao, A.; Hogan, P.G. Immunosuppressive drugs prevent a rapid dephosphorylation of transcription factor NFAT1 in stimulated immune cells. Proc. Natl. Acad. Sci. USA 1995, 92, 11205–11209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Hogan, P.G.; Chen, L.; Nardone, J.; Rao, A. Transcriptional regulation by calcium, calcineurin, and NFAT. Genes Dev. 2003, 17, 2205–2232. [Google Scholar] [CrossRef] [Green Version]
  96. Roy, J.; Cyert, M.S. Identifying New Substrates and Functions for an Old Enzyme: Calcineurin. Cold Spring Harb. Perspect. Biol. 2019, 12, 3. [Google Scholar] [CrossRef]
  97. Lally, C.; Healy, E.; Ryan, M.P. Cyclosporine A-induced cell cycle arrest and cell death in renal epithelial cells. Kidney Int. 1999, 56, 1254–1257. [Google Scholar] [CrossRef] [Green Version]
  98. Schneider, G.; Oswald, F.; Wahl, C.; Greten, F.R.; Adler, G.; Schmid, R.M. Cyclosporine inhibits growth through the activating transcription factor/cAMP-responsive element-binding protein binding site in the cyclin D1 promoter. J. Biol. Chem. 2002, 277, 43599–43607. [Google Scholar] [CrossRef] [Green Version]
  99. Kahl, C.R.; Means, A.R. Calcineurin Regulates Cyclin D1 Accumulation in Growth-stimulated Fibroblasts. Mol. Biol. Cell 2004, 15, 1833–1842. [Google Scholar] [CrossRef] [Green Version]
  100. Toyota, N.; Hashimoto, Y.; Matsuo, S.; Kitamura, Y.; Iizuka, H. Effects of FK506 and cyclosporin A on proliferation, histamine release and phenotype of murine mast cells. Arch. Dermatol. Res. 1996, 288, 474–480. [Google Scholar] [CrossRef]
  101. Khanna, A.K.; Hosenpud, J.D. Cyclosporine Induces The Expression of the Cyclin Inhibitor p21. Transplantation 1999, 67, 1262–1268. [Google Scholar] [CrossRef]
  102. Tomono, M.; Toyoshima, K.; Ito, M.; Amano, H.; Kiss, Z. Inhibitors of Calcineurin Block Expression of Cyclins A and E Induced by Fibroblast Growth Factor in Swiss 3T3 Fibroblasts. Arch. Biochem. Biophys. 1998, 353, 374–378. [Google Scholar] [CrossRef] [PubMed]
  103. Mognol, G.P.; Carneiro, F.; Robbs, B.; Faget, D.; Viola, J.P.B. Cell cycle and apoptosis regulation by NFAT transcription factors: New roles for an old player. Cell Death Dis. 2016, 7, e2199. [Google Scholar] [CrossRef] [Green Version]
  104. Bueno, O.F.; Brandt, E.B.; Rothenberg, M.E.; Molkentin, J.D. Defective T cell development and function in calcineurin A beta-deficient mice. Proc. Natl. Acad. Sci. USA 2002, 99, 9398–9403. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Hogan, P.G. Calcium–NFAT transcriptional signalling in T cell activation and T cell exhaustion. Cell Calcium 2017, 63, 66–69. [Google Scholar] [CrossRef] [Green Version]
  106. Vaeth, M.; Feske, S. NFAT control of immune function: New Frontiers for an Abiding Trooper. F1000Research 2018, 7, 260. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Kraner, S.D.; Norris, C.M. Astrocyte Activation and the Calcineurin/NFAT Pathway in Cerebrovascular Disease. Front. Aging Neurosci. 2018, 10, 287. [Google Scholar] [CrossRef] [Green Version]
  108. Shah, S.Z.A.; Hussain, T.; Zhao, D.; Yang, L. A central role for calcineurin in protein misfolding neurodegenerative diseases. Cell. Mol. Life Sci. 2016, 74, 1061–1074. [Google Scholar] [CrossRef] [PubMed]
  109. Sakuma, K.; Yamaguchi, A. The Functional Role of Calcineurin in Hypertrophy, Regeneration, and Disorders of Skeletal Muscle. J. Biomed. Biotechnol. 2010, 2010, 721219. [Google Scholar] [CrossRef] [Green Version]
  110. Tu, M.K.; Levin, J.B.; Hamilton, A.M.; Borodinsky, L.N. Calcium signaling in skeletal muscle development, maintenance and regeneration. Cell Calcium 2016, 59, 91–97. [Google Scholar] [CrossRef] [Green Version]
  111. Dewenter, M.; Von Der Lieth, A.; Katus, H.A.; Backs, J. Calcium Signaling and Transcriptional Regulation in Cardiomyocytes. Circ. Res. 2017, 121, 1000–1020. [Google Scholar] [CrossRef]
  112. Parra, V.; Rothermel, B.A. Calcineurin signaling in the heart: The importance of time and place. J. Mol. Cell. Cardiol. 2017, 103, 121–136. [Google Scholar] [CrossRef] [Green Version]
  113. Caetano, M.S.; Vieira-De-Abreu, A.; Teixeira, L.K.; Werneck, M.; Barcinski, M.A.; Viola, J. NFATC2 transcription factor regulates cell cycle progression during lymphocyte activation: Evidence of its involvement in the control of cyclin gene expression. FASEB J. 2002, 16, 1940–1942. [Google Scholar] [CrossRef]
  114. Baksh, S.; Widlund, H.; Frazer-Abel, A.A.; Du, J.; Fosmire, S.; Fisher, D.E.; DeCaprio, J.A.; Modiano, J.; Burakoff, S.J. NFATc2-Mediated Repression of Cyclin-Dependent Kinase 4 Expression. Mol. Cell 2002, 10, 1071–1081. [Google Scholar] [CrossRef]
  115. Buchholz, M.; Schatz, A.; Wagner, M.; Michl, P.; Linhart, T.; Adler, G.; Gress, T.M.; Ellenrieder, V. Overexpression of c-myc in pancreatic cancer caused by ectopic activation of NFATc1 and the Ca2+/calcineurin signaling pathway. EMBO J. 2006, 25, 3714–3724. [Google Scholar] [CrossRef] [Green Version]
  116. Leone, G.; DeGregori, J.; Sears, R.; Jakoi, L.; Nevins, J.R. Myc and Ras collaborate in inducing accumulation of active cyclin E/Cdk2 and E2F. Nature 1997, 387, 422–426. [Google Scholar] [CrossRef]
  117. Neal, J.W.; Clipstone, N.A. A Constitutively Active NFATc1 Mutant Induces a Transformed Phenotype in 3T3-L1 Fibroblasts. J. Biol. Chem. 2003, 278, 17246–17254. [Google Scholar] [CrossRef] [Green Version]
  118. Xu, W.; Gu, J.; Ren, Q.; Shi, Y.; Xia, Q.; Wang, J.; Wang, S.; Wang, Y.; Wang, J. NFATC1 promotes cell growth and tumorigenesis in ovarian cancer up-regulating c-Myc through ERK1/2/p38 MAPK signal pathway. Tumour. Biol. 2016, 37, 4493–4500. [Google Scholar] [CrossRef]
  119. Köenig, A.; Linhart, T.; Schlengemann, K.; Reutlinger, K.; Wegele, J.; Adler, G.; Singh, G.; Hofmann, L.; Kunsch, S.; Büch, T.; et al. NFAT-Induced Histone Acetylation Relay Switch Promotes c-Myc-Dependent Growth in Pancreatic Cancer Cells. Gastroenterology 2010, 138, 1189–1199.e2. [Google Scholar] [CrossRef] [Green Version]
  120. Kitamura, N.; Kaminuma, O. Isoform-Selective NFAT Inhibitor: Potential Usefulness and Development. Int. J. Mol. Sci. 2021, 22, 2725. [Google Scholar] [CrossRef]
  121. Sheftic, S.R.; Page, R.; Peti, W. Investigating the human Calcineurin Interaction Network using the piLxVP SLiM. Sci. Rep. 2016, 6, 38920. [Google Scholar] [CrossRef]
  122. Pallen, C.; Wang, J.H. A multifunctional calmodulin-stimulated phosphatase. Arch. Biochem. Biophys. 1985, 237, 281–291. [Google Scholar] [CrossRef]
  123. Morioka, M.; Hamada, J.-I.; Ushio, Y.; Miyamoto, E. Potential role of calcineurin for brain ischemia and traumatic injury. Prog. Neurobiol. 1999, 58, 1–30. [Google Scholar] [CrossRef]
  124. Wei, Q.; Holzer, M.; Brueckner, M.K.; Liu, Y.; Arendt, T. Dephosphorylation of tau protein by calcineurin triturated into neural living cells. Cell. Mol. Neurobiol. 2002, 22, 13–24. [Google Scholar] [CrossRef]
  125. Liu, Q.; Wilkins, B.J.; Lee, Y.J.; Ichijo, H.; Molkentin, J.D. Direct Interaction and Reciprocal Regulation between ASK1 and Calcineurin-NFAT Control Cardiomyocyte Death and Growth. Mol. Cell. Biol. 2006, 26, 3785–3797. [Google Scholar] [CrossRef] [Green Version]
  126. Wang, H.-G.; Pathan, N.; Ethell, I.M.; Krajewski, S.; Yamaguchi, Y.; Shibasaki, F.; McKeon, F.; Bobo, T.; Franke, T.F.; Reed, J.C. Ca2+-Induced Apoptosis Through Calcineurin Dephosphorylation of BAD. Science 1999, 284, 339–343. [Google Scholar] [CrossRef]
  127. Woolfrey, K.M.; Dell’Acqua, M.L. Coordination of Protein Phosphorylation and Dephosphorylation in Synaptic Plasticity. J. Biol. Chem. 2015, 290, 28604–28612. [Google Scholar] [CrossRef] [Green Version]
  128. Huang, C.-C.; Wang, J.-M.; Kikkawa, U.; Mukai, H.; Shen, M.-R.; Morita, I.; Chen, B.-K.; Chang, W.-C. Calcineurin-mediated dephosphorylation of c-Jun Ser-243 is required for c-Jun protein stability and cell transformation. Oncogene 2007, 27, 2422–2429. [Google Scholar] [CrossRef] [Green Version]
  129. Michod, D.; Bartesaghi, S.; Khelifi, A.; Bellodi, C.; Berliocchi, L.; Nicotera, P.; Salomoni, P. Calcium-Dependent Dephosphorylation of the Histone Chaperone DAXX Regulates H3.3 Loading and Transcription upon Neuronal Activation. Neuron 2012, 74, 122–135. [Google Scholar] [CrossRef] [Green Version]
  130. Cereghetti, G.M.; Stangherlin, A.; de Brito, O.M.; Chang, C.-R.; Blackstone, C.; Bernardi, P.; Scorrano, L. Dephosphorylation by calcineurin regulates translocation of Drp1 to mitochondria. Proc. Natl. Acad. Sci. USA 2008, 105, 15803–15808. [Google Scholar] [CrossRef] [Green Version]
  131. Bodmer, D.; Ascaño, M.; Kuruvilla, R. Isoform-Specific Dephosphorylation of Dynamin1 by Calcineurin Couples Neurotrophin Receptor Endocytosis to Axonal Growth. Neuron 2011, 70, 1085–1099. [Google Scholar] [CrossRef] [Green Version]
  132. Masaki, T.; Habara, M.; Sato, Y.; Goshima, T.; Maeda, K.; Hanaki, S.; Shimada, M. Calcineurin regulates the stability and activity of estrogen receptor α. Proc. Natl. Acad. Sci. USA 2021, 118, 44. [Google Scholar] [CrossRef]
  133. Sanderson, J.L.; Gorski, J.A.; Dell’Acqua, M.L. NMDA Receptor-Dependent LTD Requires Transient Synaptic Incorporation of Ca2+-Permeable AMPARs Mediated by AKAP150-Anchored PKA and Calcineurin. Neuron 2016, 89, 1000–1015. [Google Scholar] [CrossRef] [Green Version]
  134. Dougherty, M.K.; Ritt, D.A.; Zhou, M.; Specht, S.I.; Monson, D.M.; Veenstra, T.D.; Morrison, D.K. KSR2 Is a Calcineurin Substrate that Promotes ERK Cascade Activation in Response to Calcium Signals. Mol. Cell 2009, 34, 652–662. [Google Scholar] [CrossRef] [Green Version]
  135. Goto, S.; Yamamoto, H.; Fukunaga, K.; Iwasa, T.; Matsukado, Y.; Miyamoto, E. Dephosphorylation of Microtubule-Associated Protein 2? Factor, and Tubulin by Calcineurin. J. Neurochem. 1985, 45, 276–283. [Google Scholar] [CrossRef]
  136. Flavell, S.W.; Cowan, C.W.; Kim, T.-K.; Greer, P.L.; Lin, Y.; Paradis, S.; Griffith, E.C.; Hu, L.S.; Chen, C.; Greenberg, M.E. Activity-Dependent Regulation of MEF2 Transcription Factors Suppresses Excitatory Synapse Number. Science 2006, 311, 1008–1012. [Google Scholar] [CrossRef] [Green Version]
  137. Kolozsvári, B.; Bakó, É.; Bécsi, B.; Kiss, A.; Czikora, Á.; Tóth, A.; Vámosi, G.; Gergely, P.; Erdődi, F. Calcineurin regulates endothelial barrier function by interaction with and dephosphorylation of myosin phosphatase. Cardiovasc. Res. 2012, 96, 494–503. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Brun, M.; Glubrecht, D.D.; Baksh, S.; Godbout, R. Calcineurin Regulates Nuclear Factor I Dephosphorylation and Activity in Malignant Glioma Cell Lines. J. Biol. Chem. 2013, 288, 24104–24115. [Google Scholar] [CrossRef] [Green Version]
  139. Chow, C.-W.; Dong, C.; Flavell, R.A.; Davis, R.J. c-Jun NH2-Terminal Kinase Inhibits Targeting of the Protein Phosphatase Calcineurin to NFATc1. Mol. Cell. Biol. 2000, 20, 5227–5234. [Google Scholar] [CrossRef] [Green Version]
  140. Okamura, H.; Aramburu, J.; Garcia-Rodriguez, C.; Viola, J.; Raghavan, A.; Tahiliani, M.; Zhang, X.; Qin, J.; Hogan, P.G.; Rao, A. Concerted Dephosphorylation of the Transcription Factor NFAT1 Induces a Conformational Switch that Regulates Transcriptional Activity. Mol. Cell 2000, 6, 539–550. [Google Scholar] [CrossRef]
  141. Kim, H.B.; Kumar, A.; Wang, L.; Liu, G.H.; Keller, S.R.; Lawrence, J.C.; Finck, B.N.; Harris, T.E. Lipin 1 represses NFATc4 transcriptional activity in adipocytes to inhibit secretion of inflammatory factors. Mol. Cell. Biol. 2010, 30, 3126–3139. [Google Scholar] [CrossRef] [Green Version]
  142. Hendus-Altenburger, R.; Wang, X.; Sjogaard-Frich, L.M.; Pedraz-Cuesta, E.; Sheftic, S.R.; Bendsoe, A.H.; Page, R.; Kragelund, B.B.; Pedersen, S.F.; Peti, W. Molecular basis for the binding and selective dephosphorylation of Na(+)/H(+) exchanger 1 by calcineurin. Nat. Commun. 2019, 10, 3489. [Google Scholar] [CrossRef]
  143. Blumenthal, D.K.; Takio, K.; Hansen, R.S.; Krebs, E.G. Dephosphorylation of cAMP-dependent protein kinase regulatory subunit (type II) by calmodulin-dependent protein phosphatase. Determinants of substrate specificity. J. Biol. Chem. 1986, 261, 8140–8145. [Google Scholar] [CrossRef]
  144. Li, Y.; Sheftic, S.R.; Grigoriu, S.; Schwieters, C.D.; Page, R.; Peti, W. The structure of the RCAN1:CN complex explains the inhibition of and substrate recruitment by calcineurin. Sci. Adv. 2020, 6, eaba3681. [Google Scholar] [CrossRef]
  145. Medina, D.L.; Di Paola, S.; Peluso, I.; Armani, A.; De Stefani, D.; Venditti, R.; Montefusco, S.; Scotto-Rosato, A.; Prezioso, C.; Forrester, A.; et al. Lysosomal calcium signalling regulates autophagy through calcineurin and TFEB. Nat. Cell Biol. 2015, 17, 288–299. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Czirják, G.; Tóth, Z.E.; Enyedi, P. The Two-pore Domain K+ Channel, TRESK, Is Activated by the Cytoplasmic Calcium Signal through Calcineurin. J. Biol. Chem. 2004, 279, 18550–18558. [Google Scholar] [CrossRef] [Green Version]
  147. Nishi, A.; Snyder, G.L.; Nairn, A.; Greengard, P. Role of Calcineurin and Protein Phosphatase-2A in the Regulation of DARPP-32 Dephosphorylation in Neostriatal Neurons. J. Neurochem. 2008, 72, 2015–2021. [Google Scholar] [CrossRef]
  148. Sugimoto, T.; Stewart, S.; Guan, K.-L. The Calcium/Calmodulin-dependent Protein Phosphatase Calcineurin Is the Major Elk-1 Phosphatase. J. Biol. Chem. 1997, 272, 29415–29418. [Google Scholar] [CrossRef] [Green Version]
  149. Dawson, T.M.; Steiner, J.P.; Dawson, V.L.; Dinerman, J.L.; Uhl, G.R.; Snyder, S.H. Immunosuppressant FK506 enhances phosphorylation of nitric oxide synthase and protects against glutamate neurotoxicity. Proc. Natl. Acad. Sci. USA 1993, 90, 9808–9812. [Google Scholar] [CrossRef] [Green Version]
  150. Alao, J.P. The regulation of cyclin D1 degradation: Roles in cancer development and the potential for therapeutic invention. Mol. Cancer 2007, 6, 24. [Google Scholar] [CrossRef] [Green Version]
  151. Russell, A.F.; Thompson, M.A.; Hendley, J.; Trute, L.; Armes, J.E.; Germain, D.S. Cyclin D1 and D3 associate with the SCF complex and are coordinately elevated in breast cancer. Oncogene 1999, 18, 1983–1991. [Google Scholar] [CrossRef] [Green Version]
  152. Lara-Pezzi, E.; Winn, N.; Paul, A.; Mccullagh, K.; Slominsky, E.; Santini, M.P.; Mourkioti, F.; Sarathchandra, P.; Fukushima, S.; Suzuki, K.; et al. A naturally occurring calcineurin variant inhibits FoxO activity and enhances skeletal muscle regeneration. J. Cell Biol. 2007, 179, 1205–1218. [Google Scholar] [CrossRef]
  153. Quang, C.T.; Leboucher, S.; Passaro, D.; Fuhrmann, L.; Nourieh, M.; Vincent-Salomon, A.; Ghysdael, J. The calcineurin/NFAT pathway is activated in diagnostic breast cancer cases and is essential to survival and metastasis of mammary cancer cells. Cell Death Dis. 2015, 6, e1658. [Google Scholar] [CrossRef]
  154. Minami, T.; Jiang, S.; Schadler, K.; Suehiro, J.; Osawa, T.; Oike, Y.; Miura, M.; Naito, M.; Kodama, T.; Ryeom, S. The Calcineurin-NFAT-Angiopoietin-2 Signaling Axis in Lung Endothelium Is Critical for the Establishment of Lung Metastases. Cell Rep. 2013, 4, 709–723. [Google Scholar] [CrossRef] [Green Version]
  155. Manda, K.R.; Tripathi, P.; His, A.C.; Ning, J.; Ruzinova, M.B.; Liapis, H.; Bailey, M.H.; Zhang, H.; Maher, C.A.; Humphrey, P.A.; et al. NFATc1 promotes prostate tumorigenesis and overcomes PTEN loss-induced senescence. Oncogene 2016, 35, 3282–3292. [Google Scholar] [CrossRef] [Green Version]
  156. Wang, S.; Kang, X.; Cao, S.; Cheng, H.; Wang, D.; Geng, J. Calcineurin/NFATc1 Pathway Contributes to Cell Proliferation in Hepatocellular Carcinoma. Dig. Dis. Sci. 2012, 57, 3184–3188. [Google Scholar] [CrossRef]
  157. Tie, X.; Han, S.; Meng, L.; Wang, Y.; Wu, A. NFAT1 Is Highly Expressed in, and Regulates the Invasion of, Glioblastoma Multiforme Cells. PLoS ONE 2013, 8, e66008. [Google Scholar] [CrossRef] [Green Version]
  158. Shoshan, E.; Braeuer, R.R.; Kamiya, T.; Mobley, A.K.; Huang, L.; Vasquez, M.E.; Velazquez-Torres, G.; Chakravarti, N.; Ivan, C.; Prieto, V.; et al. NFAT1 Directly Regulates IL8 and MMP3 to Promote Melanoma Tumor Growth and Metastasis. Cancer Res. 2016, 76, 3145–3155. [Google Scholar] [CrossRef] [Green Version]
  159. Gachet, S.; Genescà, E.; Passaro, D.; Irigoyen, M.; Alcalde, H.; Clemenson, C.; Poglio, S.; Pflumio, F.; Janin, A.; Lasgi, C.; et al. Leukemia-initiating cell activity requires calcineurin in T-cell acute lymphoblastic leukemia. Leukemia 2013, 27, 2289–2300. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Medyouf, H.; Alcalde, H.; Berthier, C.; Guillemin, M.C.; dos Santos, N.R.; Janin, A.; Decaudin, D.; De Thé, H.; Ghysdael, J. Targeting calcineurin activation as a therapeutic strategy for T-cell acute lymphoblastic leukemia. Nat. Med. 2007, 13, 736–741. [Google Scholar] [CrossRef]
  161. Marafiot, T.; Pozzobon, M.; Hansmann, M.-L.; Ventura, R.; Pileri, S.A.; Roberton, H.; Gesk, S.; Gaulard, P.; Barth, T.F.E.; Du, M.Q.; et al. The NFATc1 transcription factor is widely expressed in white cells and translocates from the cytoplasm to the nucleus in a subset of human lymphomas. Br. J. Haematol. 2005, 128, 333–342. [Google Scholar] [CrossRef]
  162. Pham, L.V.; Tamayo, A.T.; Yoshimura, L.C.; Lin-Lee, Y.C.; Ford, R.J. Constitutive NF-kappaB and NFAT activation in aggressive B-cell lymphomas synergistically activates the CD154 gene and maintains lymphoma cell survival. Blood 2005, 106, 3940–3947. [Google Scholar] [CrossRef]
  163. Xin, B.; Ji, K.Q.; Liu, Y.S.; Zhao, X.D. NFAT Overexpression Correlates with CA72-4 and Poor Prognosis of Ovarian Clear-Cell Carcinoma Subtype. Reprod. Sci. 2021, 28, 745–756. [Google Scholar] [CrossRef]
  164. Wang, J.; Zhang, Y.; Liu, L.; Cui, Z.; Shi, R.; Hou, J.; Liu, Z.; Yang, L.; Wang, L.; Li, Y. NFAT2 overexpression suppresses the malignancy of hepatocellular carcinoma through inducing Egr2 expression. BMC Cancer 2020, 20, 966. [Google Scholar] [CrossRef] [PubMed]
  165. Xu, S.; Shu, P.; Zou, S.; Shen, X.; Qu, Y.; Zhang, Y.; Sun, K.; Zhang, J. NFATc1 is a tumor suppressor in hepatocellular carcinoma and induces tumor cell apoptosis by activating the FasL-mediated extrinsic signaling pathway. Cancer Med. 2018, 7, 4701–4717. [Google Scholar] [CrossRef] [Green Version]
  166. Akimzhanov, A.; Krenacs, L.; Schlegel, T.; Klein-Hessling, S.; Bagdi, E.; Stelkovics, E.; Kondo, E.; Chuvpilo, S.; Wilke, P.; Avots, A.; et al. Epigenetic Changes and Suppression of the Nuclear Factor of Activated T Cell 1 (NFATC1) Promoter in Human Lymphomas with Defects in Immunoreceptor Signaling. Am. J. Pathol. 2008, 172, 215–224. [Google Scholar] [CrossRef] [Green Version]
  167. Fruman, D.A.; Pai, S.Y.; Burakoff, S.J.; Bierer, B.E. Characterization of a mutant calcineurin A alpha gene expressed by EL4 lymphoma cells. Mol. Cell. Biol. 1995, 15, 3857–3863. [Google Scholar] [CrossRef] [Green Version]
  168. Gross, K.L.; Cioffi, E.A.; Scammell, J.G. Increased Activity of the Calcineurin–Nuclear Factor of Activated T Cells Pathway In Squirrel Monkey B-Lymphoblasts Identified By Powerblot™. In Vitro Cell. Dev. Biol. Anim. 2004, 40, 57–63. [Google Scholar] [CrossRef]
  169. Lehen’Kyi, V.; Flourakis, M.; Skryma, R.; Prevarskaya, N. TRPV6 channel controls prostate cancer cell proliferation via Ca2+/NFAT-dependent pathways. Oncogene 2007, 26, 7380–7385. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Fixemer, T.; Wissenbach, U.; Flockerzi, V.; Bonkhoff, H. Expression of the Ca2+-selective cation channel TRPV6 in human prostate cancer: A novel prognostic marker for tumor progression. Oncogene 2003, 22, 7858–7861. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Van Abel, M.; Hoenderop, J.G.J.; Bindels, R.J.M. The epithelial calcium channels TRPV5 and TRPV6: Regulation and implications for disease. Naunyn-Schmiedebergs Arch. Exp. Pathol. Pharmakol. 2005, 371, 295–306. [Google Scholar] [CrossRef] [Green Version]
  172. Furman, J.L.; Norris, C.M. Calcineurin and glial signaling: Neuroinflammation and beyond. J. Neuroinflamm. 2014, 11, 158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Musson, R.E.; Smit, N.P. Regulatory mechanisms of calcineurin phosphatase activity. Curr. Med. Chem. 2011, 18, 301–315. [Google Scholar] [CrossRef]
  174. Kawahara, T.; Kashiwagi, E.; Ide, H.; Li, Y.; Zheng, Y.; Ishiguro, H.; Miyamoto, H. The role of NFATc1 in prostate cancer progression: Cyclosporine A and tacrolimus inhibit cell proliferation, migration, and invasion. Prostate 2015, 75, 573–584. [Google Scholar] [CrossRef]
  175. Peuker, K.; Muff, S.; Wang, J.; Kunzel, S.; Bosse, E.; Zeissig, Y.; Luzzi, G.; Basic, M.; Strigli, A.; Ulbricht, A.; et al. Epithelial calcineurin controls microbiota-dependent intestinal tumor development. Nat. Med. 2016, 22, 506–515. [Google Scholar] [CrossRef] [Green Version]
  176. Grivennikov, S.I.; Greten, F.R.; Karin, M. Immunity, inflammation, and cancer. Cell 2010, 140, 883–899. [Google Scholar] [CrossRef] [Green Version]
  177. Qin, J.; Nag, S.; Wang, W.; Zhou, J.; Zhang, W.-D.; Wang, H.; Zhang, R. NFAT as cancer target: Mission possible? Biochim. Biophys. Acta (BBA) Bioenerg. 2014, 1846, 297–311. [Google Scholar] [CrossRef] [Green Version]
  178. Kawahara, T.; Kashiwagi, E.; Ide, H.; Li, Y.; Zheng, Y.; Miyamoto, Y.; Netto, G.J.; Ishiguro, H.; Miyamoto, H. Cyclosporine A and tacrolimus inhibit bladder cancer growth through down-regulation of NFATc1. Oncotarget 2015, 6, 1582–1593. [Google Scholar] [CrossRef] [Green Version]
  179. Siamakpour-Reihani, S.; Caster, J.; Bandhu Nepal, D.; Courtwright, A.; Hilliard, E.; Usary, J.; Ketelsen, D.; Darr, D.; Shen, X.J.; Patterson, C.; et al. The Role of Calcineurin/NFAT in SFRP2 Induced Angiogenesis—A Rationale for Breast Cancer Treatment with the Calcineurin Inhibitor Tacrolimus. PLoS ONE 2011, 6, e20412. [Google Scholar] [CrossRef] [Green Version]
  180. Hojo, M.; Morimoto, T.; Maluccio, M.; Asano, T.; Morimoto, K.; Lagman, M.; Shimbo, T.; Suthanthiran, M. Cyclosporine induces cancer progression by a cell-autonomous mechanism. Nature 1999, 397, 530–534. [Google Scholar] [CrossRef]
  181. Datta, D.; Contreras, A.; Basu, A.; Dormond, O.; Flynn, E.; Briscoe, D.; Pal, S. Calcineurin Inhibitors Activate the Proto-Oncogene Ras and Promote Protumorigenic Signals in Renal Cancer Cells. Cancer Res. 2009, 69, 8902–8909. [Google Scholar] [CrossRef] [Green Version]
  182. Han, W.; Ming, M.; He, T.-C.; He, Y.-Y. Immunosuppressive Cyclosporin A Activates AKT in Keratinocytes through PTEN Suppression: Implications In Skin Carcinogenesis. J. Biol. Chem. 2010, 285, 11369–11377. [Google Scholar] [CrossRef] [Green Version]
  183. Cui, C.; Merritt, R.; Fu, L.; Pan, Z. Targeting calcium signaling in cancer therapy. Acta Pharm. Sin. B 2017, 7, 3–17. [Google Scholar] [CrossRef]
  184. Pusl, T.; Wu, J.J.; Zimmerman, T.L.; Zhang, L.; Ehrlich, B.E.; Berchtold, M.W.; Hoek, J.B.; Karpen, S.J.; Nathanson, M.H.; Bennett, A.M. Epidermal Growth Factor-mediated Activation of the ETS Domain Transcription Factor Elk-1 Requires Nuclear Calcium. J. Biol. Chem. 2002, 277, 27517–27527. [Google Scholar] [CrossRef] [Green Version]
  185. Andrade, L.M.; Geraldo, J.M. Nucleoplasmic Calcium Buffering Sensitizes Human Squamous Cell Carcinoma to Anticancer Therapy. J. Cancer Sci. Ther. 2012, 4, 5. [Google Scholar] [CrossRef]
Figure 1. Regulation of cell cycle progression by calcineurin and CaMK. CaN dephosphorylates the NFAT transcription factor, which in turn activates p21, cyclin D1, CDK4, c-myc, and cyclin A. CaN also stabilizes cyclin D1 by dephosphorylation. Furthermore, the CaN/NFAT pathway and its downstream target c-myc regulate p21. p21 is a well-known inhibitor of CDK2-cyclin E and CDK4/6-cyclin D. CaMK negatively regulates the expression of p27, which is an inhibitor of CDK4-cyclin D and CDK2-cyclin E. CDK2-cyclin E and CDK4/6-cyclin D complexes phosphorylate Rb, leading to the activation of E2F1 and the subsequent G1/S progression. In G2/M, CaMK phosphorylates and activates cdc25, leading to downstream dephosphorylation and the activation of CDK1. Solid red lines indicate phosphorylation, red dotted lines indicate dephosphorylation, and green dotted lines indicate transcriptional activation.
Figure 1. Regulation of cell cycle progression by calcineurin and CaMK. CaN dephosphorylates the NFAT transcription factor, which in turn activates p21, cyclin D1, CDK4, c-myc, and cyclin A. CaN also stabilizes cyclin D1 by dephosphorylation. Furthermore, the CaN/NFAT pathway and its downstream target c-myc regulate p21. p21 is a well-known inhibitor of CDK2-cyclin E and CDK4/6-cyclin D. CaMK negatively regulates the expression of p27, which is an inhibitor of CDK4-cyclin D and CDK2-cyclin E. CDK2-cyclin E and CDK4/6-cyclin D complexes phosphorylate Rb, leading to the activation of E2F1 and the subsequent G1/S progression. In G2/M, CaMK phosphorylates and activates cdc25, leading to downstream dephosphorylation and the activation of CDK1. Solid red lines indicate phosphorylation, red dotted lines indicate dephosphorylation, and green dotted lines indicate transcriptional activation.
Ijms 23 01122 g001
Figure 2. A new role of calcineurin in cell proliferation. (a) CaN regulates the stability and transcription of Cyclin D1. Phosphorylation of Thr286 of cyclin D1 promotes polyubiquitination and the subsequent degradation of cyclin D1. CaN stabilizes cyclin D1 by dephosphorylating the Thr286 of cyclin D1. Furthermore, NFAT activated by CaN induces the transcription of cyclin D1. Thus, CaN contributes to cancer cell proliferation by regulating the stability and transcription of cyclin D1. (b) Calcineurin regulates the stability and activity of Erα. Phosphorylation of Ser294 of ERα promotes polyubiquitination by the E3 ligase E6AP and the subsequent degradation of ERα. By dephosphorylating the Ser294 of ERα, CaN releases E6AP from ERα and stabilizes ERα (left). Furthermore, CaN activates the Akt-mTOR pathway via the phosphorylation of Ser118 of ERα (right). Thus, CaN contributes to the proliferation of cancer cells by regulating the stability and activity of ERα.
Figure 2. A new role of calcineurin in cell proliferation. (a) CaN regulates the stability and transcription of Cyclin D1. Phosphorylation of Thr286 of cyclin D1 promotes polyubiquitination and the subsequent degradation of cyclin D1. CaN stabilizes cyclin D1 by dephosphorylating the Thr286 of cyclin D1. Furthermore, NFAT activated by CaN induces the transcription of cyclin D1. Thus, CaN contributes to cancer cell proliferation by regulating the stability and transcription of cyclin D1. (b) Calcineurin regulates the stability and activity of Erα. Phosphorylation of Ser294 of ERα promotes polyubiquitination by the E3 ligase E6AP and the subsequent degradation of ERα. By dephosphorylating the Ser294 of ERα, CaN releases E6AP from ERα and stabilizes ERα (left). Furthermore, CaN activates the Akt-mTOR pathway via the phosphorylation of Ser118 of ERα (right). Thus, CaN contributes to the proliferation of cancer cells by regulating the stability and activity of ERα.
Ijms 23 01122 g002
Table 1. Calcineurin substrates.
Table 1. Calcineurin substrates.
Calcineurin SubstratesDephosphorylation SitesReaction by DephosphorylationReference
12E8Ser262, Ser356ND[124]
ASK1Ser967Promotes the dissociation of ASK1 from the 14-3-3 protein, resulting in the activation of ASK1[125]
AT270Thr181ND[124]
BADSer155Promotes heterodimerization of BAD and Bcl-xL, which induces apoptosis[126]
CaMKIIγSer334Translocates CaMKIIγ to the nucleus[127]
c-JunSer243Stabilizes c-Jun, promotes the interaction between c-Jun and Sp1[128]
Cyclin D1Thr286Stabilizes cyclin D1, inducing G1/S progression[38]
DAXXSer669Promotes H3.3 uptake by DAXX[129]
DRP1Ser637Splites the organelle by Drp1[130]
Dynamin 1Ser774, Ser778Promotes endocytosis of TrkA receptors and axonal growth[131]
ERαSer294Stabilizes ERα and promotes the activity of ERα[132]
GluA1Ser845Promotes removal of AMPARs from synapses and endocytosis[133]
KSR2Ser198, Thr287, Ser310Activates ERK and induces membrane localization of KSR2[134]
MAP2 ND[135]
MEF2ASer221, Ser255, Ser408Activates MEF2A and promotes the change from sumoylation to acetylation of Leu403[136]
MYPT1Thr696Affects actin polymerization by activating MP and improves endothelial barrier function[137]
NF1NDActivates transcription[138]
NFATC1Ser172Promotes nuclear transfer of NFATC1[139]
NFATC2 *Five residues among following sites, Ser170, Ser173, Ser174, Ser176, Ser177, Ser179, Ser182, in SRR-1 domainPromotes nuclear transfer of NFATC2[140]
Ser215, Ser219, Ser223 in SP-2 domain
Ser270, Ser276, Ser278, Ser282 in SP-3 domain
Ser328 in KTS motif
NFATC4Ser170Promotes nuclear transfer of NFATC4[141]
NHE1Thr779Inhibits NHE1 activity[142]
PHF1Ser396, Ser404ND[124]
RIIαSer95ND[143]
RCAN1Ser108, Ser112, Thr124, Thr192ND[144]
Tau1Ser199, Ser202ND[124]
TFEBSer142, Ser211Promotes nuclear transfer of TFEB[145]
TRESKSer276Increases K+ current, decreases channel responsiveness to calcium signals[146]
The amino acid residues of the protein indicated by asterisks (*) are dephosphorylated when the cells are stimulated with ionomycin. It is not known whether calcineurin directly dephosphorylates them. ND means not determined.
Table 2. Dysregulation of the calcineurin/NFAT pathway in cancer.
Table 2. Dysregulation of the calcineurin/NFAT pathway in cancer.
FactorAlterations in CancerTypes of CancerReference
Calcineurin (CnA)Overexpressionglioma (malignant gliomas, including grades III and IV astrocytomas)[138]
breast cancer (ER-α–positive)[132]
Activationlymphomas (lymphoid malignancies)[160]
Overexpression, activationcolon cancer[39]
NFATc1Overexpressionovarian cancer (clear-cell carcinoma)[118,163]
liver cancer (hepatocellular carcinoma)[156]
prostate cancer[155]
lymphomas (large B-cell lymphoma)[162]
Nuclear localizationlymphomas (diffuse large B-cell lymphomas)[161]
breast cancer (triple-negative)[153]
Suppressionliver cancer (hepatocellular carcinoma)[164,165]
lymphomas (anaplastic large cell lymphomas and classical Hodgkin’s lymphomas)[166]
Overexpression, nuclear localizationpancreatic cancer (pancreatic adenocarcinoma)[115]
NFATc2Overexpressionmelanoma[158]
glioma (glioblastoma)[157]
Nuclear localizationlung cancer[154]
NFATc1, NFATc3Dephosphorylationleukemia (T-cell acute lymphoblastic leukemia)[159]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Masaki, T.; Shimada, M. Decoding the Phosphatase Code: Regulation of Cell Proliferation by Calcineurin. Int. J. Mol. Sci. 2022, 23, 1122. https://doi.org/10.3390/ijms23031122

AMA Style

Masaki T, Shimada M. Decoding the Phosphatase Code: Regulation of Cell Proliferation by Calcineurin. International Journal of Molecular Sciences. 2022; 23(3):1122. https://doi.org/10.3390/ijms23031122

Chicago/Turabian Style

Masaki, Takahiro, and Midori Shimada. 2022. "Decoding the Phosphatase Code: Regulation of Cell Proliferation by Calcineurin" International Journal of Molecular Sciences 23, no. 3: 1122. https://doi.org/10.3390/ijms23031122

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop