Next Article in Journal
Drosophila as a Model for Infectious Diseases
Next Article in Special Issue
Decoding the Phosphatase Code: Regulation of Cell Proliferation by Calcineurin
Previous Article in Journal
Towards Quantitative and Standardized Serological and Neutralization Assays for COVID-19
Previous Article in Special Issue
Finely-Tuned Calcium Oscillations in Osteoclast Differentiation and Bone Resorption
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Isoform-Selective NFAT Inhibitor: Potential Usefulness and Development

1
Laboratory of Allergy and Immunology, Tokyo Metropolitan Institute of Medical Science, Tokyo 156-8506, Japan
2
Department of Disease Model, Research Institute of Radiation Biology and Medicine, Hiroshima University, Hiroshima 734-8553, Japan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(5), 2725; https://doi.org/10.3390/ijms22052725
Submission received: 10 February 2021 / Revised: 26 February 2021 / Accepted: 1 March 2021 / Published: 8 March 2021
(This article belongs to the Special Issue Updates of Calcineurin/NFAT Signaling in Human Health and Diseases)

Abstract

:
Nuclear factor of activated T cells (NFAT), which is the pharmacological target of immunosuppressants cyclosporine and tacrolimus, has been shown to play an important role not only in T cells (immune system), from which their name is derived, but also in many biological events. Therefore, functional and/or structural abnormalities of NFAT are linked to the pathogenesis of diseases in various organs. The NFAT protein family consists of five isoforms, and each isoform performs diverse functions and has unique expression patterns in the target tissues. This diversity has made it difficult to obtain ideal pharmacological output for immunosuppressants that inhibit the activity of almost all NFAT family members, causing serious and wide-ranging side effects. Moreover, it remains unclear whether isoform-selective NFAT regulation can be achieved by targeting the structural differences among NFAT isoforms and whether this strategy can lead to the development of better drugs than the existing ones. This review summarizes the role of the NFAT family members in biological events, including the development of various diseases, as well as the usefulness of and problems associated with NFAT-targeting therapies, including those dependent on current immunosuppressants. Finally, we propose a novel therapeutic strategy based on the molecular mechanisms that enable selective regulation of specific NFAT isoforms.

1. Introduction

Nuclear factor of activated T cells (NFAT) was identified as a transcriptional regulator of interleukin-2 (IL-2) in activated T cells by Crabtree’s group in 1988 [1]. In 1993, Rao’s group reported that the pharmacological actions of both cyclosporine and FK506 (tacrolimus), which are used to attenuate transplant rejection, were achieved via inhibition of NFAT activity in T cells [2]. Thus, much attention has been paid to the importance of NFAT in T cell function and differentiation [3,4]. However, NFAT family members are involved in the development, differentiation, and function of various tissues, such as the nervous, cardiovascular, and bone metabolism systems, as well as the immune system [5,6,7].
The NFAT gene family consists of five isoforms (NFATc1-c4, NFAT5), commonly involving a Rel-homology domain (RHD) that is responsible for DNA binding. NFAT5 is always distributed in both the cytoplasm and nucleus, and its activity is mainly regulated by osmotic pressure. Under high osmotic stress, NFAT5 participates in maintaining cellular functions and homeostasis [8,9]. NFAT5 also plays a role, even under isotonic conditions, in adjusting the intracellular environment suitable for cell proliferation [9]. On the other hand, NFATc1-c4 (NFATcs) are normally present in the cytoplasm, and their activities are regulated by the Ca2+/calmodulin-dependent serine/threonine phosphatase, calcineurin (CN) [10,11]. Upon activation with an increase in intracellular Ca2+ concentration, CN directly binds to the calcium regulatory domain (CRD) of NFATcs located at the N-terminal region of RHD and dephosphorylates its multiple phosphoserine residues (Figure 1). The subsequent conformational changes and exposure of the nuclear localization signal (NLS) enable NFATcs to translocate into the nucleus and promote the transcription of target genes by binding to the corresponding sequence of their transcription modulatory region [12,13,14]. Since the NFAT family is involved in many biological events, its functional and structural abnormalities often cause various diseases. Therefore, many researchers have proposed new means to regulate NFAT, but it has not yet reached clinical application. To overcome the present situation, this review describes the significance and feasibility of isoform-selective regulation of NFATcs, with the introduction of previous studies including our recent findings.

2. Functional Diversity of NFAT Family Members

NFAT is expressed in various cells and tissues and participates in the transcription of many genes, not just the initially discovered IL-2. Moreover, individual NFAT isoforms do not show the same expression patterns or functional characteristics in each cell type and/or organ (Supplementary Figures S1–S5). Based on the history of their discovery, T cells have been regarded as the principal sites of NFAT function, although NFATc4 is rarely expressed in T cells [15]. NFATc3 is more dominantly expressed than NFATc1 or NFATc2 in the thymus, while NFATc1 is highly expressed in muscle cells [12,16]. Studies using genetically modified mice have also revealed multiple functions of NFAT in the developmental processes of various organs (Table 1). NFATc1 is involved in the development of the endocardium, valves, and septum, while NFATc3 and NFATc4 control the expression of genes that regulate vascular patterning and vascularization. Consistently, mice deficient for NFATc1 or NFATc3/NFATc4 exhibit embryonic lethality due to organ hypoplasia [5,17,18]. NFAT5 knockout mice are mostly embryonic lethal caused by renal atrophy or decreased cardiac function [19,20]. NFATc2-deficient mice are born and grow normally but exhibit enhanced immune responsiveness, including splenomegaly [21,22]. NFATc4-deficiency has little effect on growth but does affect abnormal spatial memory formation in mice [23]. Conversely, NFATc4 transgenic mice develop cardiac hypertrophy and heart failure [24].
Furthermore, NFAT plays important roles in modulating cell survival, differentiation, and proliferation by regulating the expression of various genes involved in cell cycle and apoptosis, such as CDK4, cyclin, c-Myc, and FasL, as well as many cytokines and growth factors [12,25,26,27]. Through the induction of neural activity-regulating molecules, NFAT participates in the development and maintenance of the neural system [6,23]. NFATc1 plays a key role in bone metabolism, especially as a major regulator of osteoclasts [28,29].
Since NFAT regulates many important genes as described above, abnormal NFAT function often leads to the development and progression of various diseases. Constitutive activation of NFATc1 facilitates cell cycle progression and induces transformation, resulting in tumor development [30,31]. Moreover, NFATc1 participates in cancer infiltration and metastasis by inducing angiogenesis and lymphangiogenesis through the up-regulation of cyclooxygenase 2, vascular endothelial growth factor (VEGF), and CXC chemokine receptor 7 expression [32,33,34]. Clinically, NFATc1 is overexpressed in various types of tumors, such as pancreatic cancer, colorectal cancer, various leukemias, and lymphomas [31,35,36,37,38,39]. NFATc1 overexpression has been reported to be associated with poor prognosis in some cancers [40]. On the other hand, NFATc2 is implicated in an inhibitory effect on tumorigenesis, including induction of cell cycle arrest and apoptosis; thus, spontaneous lymphoma development was observed in NFATc2-deficient mice [41,42]. Nevertheless, enhanced expression of NFATc2 has been reported in patients with metastatic cancers, invasive cancers, and highly malignant cancers [34,43,44,45], suggesting that functional changes in NFATc2 may induce angiogenic dysregulation in cancer cells. The participation of NFATc3 and NFATc4 in tumorigenesis has been suggested [46,47], whereas their inhibitory effects on metastasis and infiltration in estrogen receptor-positive breast cancer [48], as well as that on lymphoma formation caused by retroviral infection [49], were also reported. Each NFAT isoform appears to play different roles depending on the type and progression of cancer.
Among the neurodegenerative disorders, a strong causal relationship has been suggested between Alzheimer’s disease (AD) and abnormal activity of NFATcs. Analysis of the human postmortem brain demonstrated that NFATc2 and NFATc4 accumulate in the hippocampus in subjects with mild cognitive impairment, severe dementia, and AD, respectively [50]. The levels of amyloid β, which is known to increase in the brain according to the severity of dementia, was correlated with that of nuclear NFATc4 [50]. Introduction of constitutively active NFATc4 into nerve cells reproduced amyloid β-induced morphological neurodegenerative changes in vitro [51]. In AD, dysregulation of intracellular calcium, which affects the enzymatic activity of CN, was observed in nerve cells [50]. The involvement of the CN/NFAT-dependent pathway in α-synuclein-induced degeneration of midbrain dopaminergic neurons in Parkinson’s disease has been suggested [52].
NFATcs are also key molecules that induce immune tolerance because they regulate the expression of Foxp3 and CTLA4, which are essential for the development and function of regulatory T cells [53,54,55,56,57]. Furthermore, NFATcs control activation-induced cell death through FasL upregulation [53,58,59]. Accordingly, anomalous NFAT activity results in the development of autoimmune and inflammatory diseases. NFAT is involved in many cellular processes typically observed in the synovium of patients with rheumatoid arthritis (RA), including activation of inflammatory cells, production of various cytokines, VEGF-mediated pathologic angiogenesis, and osteoclast formation [60]. Mice lacking leucine-rich repeat kinase 2, which is a negative regulator of NFAT and a major susceptibility gene in Crohn’s disease, displayed exacerbated experimental colitis accompanied by the activation of NFATc2 [61]. The severity of experimental allergic encephalomyelitis was diminished in mice deficient in both NFATc1 and NFATc2 [62], suggesting that NFAT also contributes to the development of multiple sclerosis. Through the regulation of macrophage activity [63] and high osmolarity-dependent pathogenic Th17 cell induction [64], NFAT5 may participate in the pathogenesis of RA and other autoimmune diseases.

3. Usefulness and Problems of Direct NFAT Regulation

According to the progress in elucidating the role of NFAT in various diseases, the potential application of NFAT-targeting therapies has been expanded. Cyclosporine and/or tacrolimus, initially developed as immunosuppressive agents for organ transplantation, are currently used for many allergic and autoimmune diseases, such as atopic dermatitis, bronchial asthma, RA, ulcerative colitis, lupus nephritis, myasthenia gravis, Behcet’s disease, regenerative anemia, psoriasis, nephrotic syndrome, and Kawasaki disease. It has been reported that the incidence of dementia in patients who received cyclosporine or tacrolimus after transplantation was significantly lower than that in the general population [65]. However, serious and wide-ranging side effects, including increased risk of infectious diseases due to immunosuppression; organ disorders, such as those of the kidney, liver, and pancreas; cardiovascular disorders, including hypertension; nervous system disorders; and malignancies are major problems in using these drugs [66,67,68]. Their prolonged application further exacerbates side effects, thereby deteriorating the quality of life (QOL) of patients.
Numerous side effects of cyclosporine and tacrolimus are caused, at least in part, by the fact that they do not act directly on NFAT, but on CN, which catalyzes the dephosphorylation of NFATcs. Cyclosporine and tacrolimus form a complex with distinct intracellular proteins, cyclophilin and FK-binding proteins, respectively. These drug–protein complexes bind to similar sites in CN and inhibit phosphatase activity [69]. Although the possible treatment of breast cancer via CN-mediated cyclin D1 dephosphorylation has also been suggested [70], more than 50 substrate molecules other than NFATcs are regulated by CN [71,72,73]. Due to the widespread function of CN, the strategy for NFAT regulation by targeting CN activity may not lead to an improvement in patient QOL.
Therefore, to address the concerns regarding the side effects of CN inhibition, new agents that directly regulate NFATcs activity have been developed. NFATcs have been shown to interact with CN via two CN binding regions (CNBRs) located in the CRD. Based on the common sequence of the N-terminal side CNBR (CNBR1) of NFATcs, a modified 16 amino acid peptide (MAGPHPVIVITGPHEE) that exhibited enhanced affinity for CN was developed. The peptide, named VIVIT, based on its core sequence [74], suppressed the activation of NFATcs by competing with the conjugation of NFATcs to CN. The existing immunosuppressant-like pharmacological effects without expected side effects were obtained by the VIVIT peptide. Thus, the prevention of allograft rejection in islet-transplanted mice by tacrolimus was associated with a dose-dependent decrease in insulin secretion. On the other hand, essentially the same inhibitory effect on allograft rejection was achieved in mice administered the polyarginine-conjugated VIVIT peptide (11R-VIVIT) without impacting insulin secretion [75]. Introduction of the VIVIT peptide using an adenovirus-associated vector into AD model mice ameliorated the morphological neurodegenerative changes around amyloid β plaques [51]. Furthermore, the suppressive effect of the VIVIT peptide has been indicated in mouse models of heart disease, colitis, bronchial asthma, and type 2 diabetes [76,77,78,79]. Potential applications of the other types of direct NFAT inhibitors, such as A-285222 and INCA-6, to type 2 diabetes, diabetic retinopathy, and age-related macular degeneration, have also been demonstrated [80,81] (Table 1).

4. Significance of Selective Control of Specific NFAT Isoforms

Molecules that directly regulate NFATcs, such as VIVIT, may be useful for treating multiple diseases, although their clinical applications have not yet been achieved. Although the fewer side effects of the direct-acting NFAT inhibitors compared to cyclosporine or tacrolimus are promising, whether a more potent pharmacological action is expected in the direct inhibitors is unclear. Since the phenotypes of genetically modified mice for each NFAT are divergent, as mentioned above, outputs in the transcription of various target genes by regulating each NFAT isoform are expected to be different. The expression of IL-4, a typical T cell cytokine, is strongly suppressed in NFATc1-deficient T cells [82,83] but enhanced in some NFATc2-deficient mice [22,84,85]. For controlling IL-4 production by T cells, selective inhibition of NFATc1 alone may be more effective than targeting the entire NFAT family. However, by utilizing genetically modified mice systems alone, it is difficult to elucidate the detailed contribution of each NFAT isoform to the transcription of individual genes.
The potency of NFATc1 and NFATc2 to activate the transcription of cytokine genes in T cells and the corresponding mechanisms were comparatively analyzed by their ectopic introduction into T cells with careful adjustment of the expression levels. Both NFATc1 and NFATc2 augmented stimulation-induced IL-2 and GM-CSF expression. However, for TNF-α and IL-13 transcription, NFATc2 strongly enhanced transcription, whereas NFATc1 showed little contribution [86]. Conversely, IL-4 expression was more strongly induced by NFATc1. The difference in the potency of these isoforms to activate the transcription of cytokines might be linked to the different phenotypes observed in the corresponding knockout mice, as described above, typically in the case of IL-4. Moreover, by employing T cells introduced with various chimeric NFAT molecules in which each functional domain was exchanged between NFATc1 and NFATc2, it was clarified that the functional difference between these isoforms is caused, at least in part, by the deficiency of the C-terminal transcriptional activation domain in a dominant NFATc1 variant [86].
NFAT collaborates with other transcription factors to exhibit its transcription activation property in many genes [12,13,14]. In particular, on the IL-2 and TNF-α promoters, NFAT has been shown to associate with different co-effector heterodimer molecules, such as Jun/Fos (activator protein 1) and Jun/activating transcription factor 2, respectively [12,14,87]. NFATc1 and NFATc2 may exert different regulatory functions on the same cytokine gene by distinctively interacting with their individual co-effectors based on the different domain structures [86].
The differences in the tissue expression patterns of NFATcs have been investigated. In T cells, where NFAT plays an important role in their function, the NFATc4 level is much lower than that of other NFATcs [15]. Remarkably, cytokine production induced in human T cells upon activation was strongly suppressed by ectopic expression of NFATc4 [88]. The difference in the role of each NFAT isoform in the regulation of IL-2 expression was further examined by knockdown of endogenous NFATcs in Jurkat cells, a human T cell line that expresses NFATc4 at a relatively high level, by introducing the corresponding small interfering RNA. Stimulation-induced IL-2 expression was diminished by NFATc1, NFATc2, or NFATc3 knockdown but enhanced by NFATc4 knockdown, suggesting that NFATc4, which is slightly expressed in T cells, plays a suppressive role in cytokine expression. T cells might have evolved to acquire their characteristic features, the cytokine-producing ability, by reducing the expression of the suppressive NFAT isoform, NFATc4. On the other hand, NFATc4 is highly expressed in aortic smooth muscle cells (ASMCs) and regulates the expression of genes involved in their development and differentiation. The difference in the expression and resulting distinct function of NFATc4 in T cells and ASMC were regulated by the expression level of T-box transcription factor TBX5 (Figure 2) [88].
In addition to Ca2+-independent regulation of NFAT5, differences in the regulatory mechanisms of NFATcs have been reported. Kar et al. demonstrated that NFATc2 was activated by Ca2+ microdomains near open store-operated Ca2+ release-activated Ca2+ channels, whereas NFATc3 activation further required nuclear Ca2+ mobilization [89]. It seems that NFAT isoforms play individual complex roles through diversity in expressed tissues, functions, and regulatory mechanisms.

5. New Intramolecular Regions Involved in CN/NFAT Interactions

The identification of the opposite function among NFAT isoforms has improved the significance of selective means to control specific NFAT isoforms. Thus, NFAT isoform inhibitors may be superior to cyclosporine and tacrolimus not only in reducing side effects but also in augmenting the pharmacological efficacy. Therefore, the molecular mechanisms that enable isoform-selective control of NFAT have been investigated. In our recent study, detailed CN-binding properties were compared among CNBR1, CNBR2, and their intermediate region proteins, as well as the entire CRD of NFATcs, by employing recombinant proteins expressed in Escherichia coli (E. coli) (Figure 1). The catalytic subunit of CN, CNA, was also expressed and purified, and the binding affinity of each NFAT region with CNA was quantitatively assessed. However, despite employing several techniques for evaluating protein–protein interactions, such as the biophysical interaction analysis method using surface plasmon resonance (such as Biacore) and amplified emission proximity homogeneous assay technology, we could not obtain reasonable findings. Recombinant proteins normally contain impurities, and their purity depends on not only the size, sequence, and properties of their amino acids but also the expression and purification methods. Since these sophisticated techniques we initially applied are designated for analyzing interactions between highly purified molecules, they might not distinguish the target proteins and their impurities. To circumvent this distress, a new quantitative immunoprecipitation method was developed by employing highly purified proteins obtained by utilizing the DYKDDDDK tag inserted at the C-terminal end, in which only the full-length proteins expressed in E. coli were theoretically included. Owing to these technical improvements, the binding affinity between each NFAT region and CNA was successfully compared (Table 2).
Similar CNA-binding activity with CRD was observed for all NFATcs. The CNBR1 region contributed to the interaction with CNA in all NFATcs, albeit with some differences in affinity. The CNBR2 region of NFATc1, NFATc3, and NFATc4 participated in CNA binding almost equally, while the binding affinity between NFATc2-CNBR2 and CNA was low, about 1/10th of that of other NFATcs. These results were in close agreement with previous findings obtained from qualitative experiments [90,91]. Remarkably, we found a novel CNA binding region in CRD that might exhibit the isoform-selective function. The intermediate region between CNBR1 and CNBR2 of NFATc1 and NFATc4, but not the other NFATcs, showed significantly strong binding activity to CNA (Table 2). Ectopic expression of this region, named CNBR3, in BHK cells that constitutively expressed fluorescence-labeled CRD proteins suppressed stimulation-induced nuclear translocation of NFATc1-CRD but not NFATc2-CRD. The interaction between NFATc1-CNBR3 and CNA was not affected by the VIVIT peptide or another CNA binding sequence peptide (DSSGDQFLSVPSPFTW) derived from the CNBR2 sequence, suggesting that the CNBR3 binding region in CNA is different from the regions responsible for interactions with CNBR1 and CNBR2. The binding regions of CNBR3 and CNA were narrowed down by competition assay using NFATc1-derived partial peptides and mass spectrometry with photoaffinity technology. Eighteen amino acids in NFATc1 (Arg258 to Pro275) and 13 amino acids in CNA (Asn77 to Gly89) were identified as the region involved in this binding (Figure 3) [92]. Amino acid substitution experiments, based on the binding model derived from the molecular operating environment (MOE) integrated computational system, further revealed that the interaction between Cys263 in NFATc1 and Asp82 in CNA was particularly essential for their binding (Figure 3).

6. Prospects for Disease Treatments Targeting the New Binding Region

The potency and quality of novel means to control NFAT activity by targeting CNBR3, which probably affects only NFATc1 and NFATc4, probably differ from those of cyclosporine and tacrolimus. The stronger pharmacological effects of these immunosuppressants are expected in the new approach, especially for treating diseases where functional incompatibility occurs among NFAT isoforms. As described above, the relatively potent involvement of NFATc1 has been suggested in several diseases such as cancer, osteoporosis, and allergic disorders, while the selective role of NFATc4 has been suggested in AD. The effectiveness of immunosuppressants and/or direct NFAT inhibitors on those diseases was experimentally proven, although their clinical usages have not been approved. Therefore, the application of isoform-selective NFAT control methods promisingly achieved by targeting CNBR3 seems to be more appropriate and feasible at least to treat those diseases (Figure 4, Table 3). Since it was recently reported that cyclosporine and tacrolimus reduce sperm motility [93], the development of isoform-selective NFAT inhibitors that can be distinguished from classical immunosuppressants may be helpful as a countermeasure against the declining birth rate, which is becoming a serious problem in some countries.

7. Conclusions

The NFAT family of transcription factors with diverse functions is important not only in the immune system but also in various biological events. The findings of many studies to date indicate the overall picture of the physiological role of the NFAT family members and their concrete functions associated with several diseases. NFAT isoforms exhibit vast diversity because each isoform functions in an overlapping or opposite manner with respect to expression patterns within cells and tissues as well as to the functional characteristics, including participation in various diseases. Therefore, selective control of the individually contributing NFAT isoforms may be a key strategy for treating diseases, particularly those involving the corresponding isoforms. Our latest results show that pyrogallol derived from Awa tea associates with CNBR3 and consequently inhibits NFATc1 activity and IL-9 gene expression (manuscript in preparation). Even half a century after the discovery of cyclosporine, there are still many issues restricting the clinical application of therapeutic methods based on the direct control of NFAT. It is further complicated by recent investigations indicating the new regulatory mechanism of NFAT activity through SUMOylation [94] and the regulatory T cell induction dependent on the threshold value of total NFAT members [95]. However, the identification of new CN-binding regions that selectively contribute to specific NFAT-isoforms may lead to the development of innovative therapeutic strategies for improving the QOL of many patients suffering from refractory diseases or diseases for which there is no cure.

Supplementary Materials

The following are available online at https://www.mdpi.com/1422-0067/22/5/2725/s1.

Author Contributions

Conceptualization, O.K.; methodology, N.K. and O.K.; software, O.K.; validation, N.K.; formal analysis, N.K. and O.K.; investigation, N.K. and O.K.; resources, N.K. and O.K.; data curation, N.K. and O.K.; writing—original draft preparation, N.K.; writing—review and editing, O.K.; visualization, N.K. and O.K.; supervision, O.K.; project administration, N.K. and O.K.; funding acquisition, N.K. and O.K. 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 to Noriko Kitamura (No. 24590107) and Osamu Kaminuma (No. 21590086) and by funding from the Terumo Life Science Foundation to Osamu Kaminuma.

Acknowledgments

The authors thank Takachika Hiroi and staff at the Laboratory of Allergy and Immunology, Tokyo Metropolitan Institute of Medical Science, and Kento Miura and staff at the Department of Disease Model, Research Institute of Radiation Biology and Medicine, Hiroshima University, for excellent assistance and advice.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shaw, J.-P.; Utz, P.J.; Durand, D.B.; Toole, J.J.; Emmel, E.A.; Crabtree, G.R. Identification of putative regulator of ealry T cell activation genes. Science 1988, 241, 202–205. [Google Scholar] [CrossRef]
  2. McCaffrey, P.G.; Perrino, B.A.; Soderling, T.R.; Rao, A. NF-ATp, a T lymphocyte DNA-binding protein that is a target for calcineurin and immunosuppressive drugs. J. Biol. Chem. 1993, 268, 3747–3752. [Google Scholar] [CrossRef]
  3. Chow, C.W.; Rincón, M.; Davis, R.J. Requirement for transcription factor NFAT in interleukin-2 expression. Mol. Cell. Biol. 1999, 19, 2300–2307. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Serfling, E.; Berberich-Siebelt, F.; Chuvpilo, S.; Jankevics, E.; Klein-Hessling, S.; Twardzik, T.; Avots, A. The role of NF-AT transcription factors in T cell activation and differentiation. Biochim. Biophys. Acta 2000, 1498, 1–18. [Google Scholar] [CrossRef] [Green Version]
  5. Ranger, A.M.; Grusby, M.J.; Hodge, M.R.; Gravallese, E.M.; de la Brousse, F.C.; Hoey, T.; Mickanin, C.; Baldwin, H.S.; Glimcher, L.H. The transcription factor NF-ATc is essential for cardiac valve formation. Nature 1998, 392, 186–190. [Google Scholar] [CrossRef] [PubMed]
  6. Benedito, A.B.; Lehtinen, M.; Massol, R.; Lopes, U.G.; Kirchhausen, T.; Rao, A.; Bonni, A. The transcription factor NFAT3 mediates neuronal survival. J. Biol. Chem. 2005, 280, 2818–2825. [Google Scholar] [CrossRef] [Green Version]
  7. Zayzafoon, M. Calcium/calmodulin signaling controls osteoblast growth and differentiation. J. Cell. Biochem. 2006, 97, 56–70. [Google Scholar] [CrossRef]
  8. Maeoka, Y.; Wu, Y.; Okamoto, T.; Kanemoto, S.; Guo, X.P.; Saito, A.; Asada, R.; Matsuhisa, K.; Masaki, T.; Imaizumi, K.; et al. NFAT5 up-regulates expression of the kidney-specific ubiquitin ligase gene Rnf183 under hypertonic conditions in inner-medullary collecting duct cells. J. Biol. Chem. 2019, 294, 101–115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Ho, S.N. The role of NFAT5/TonEBP in establishing an optimal intracellular environment. Arch. Biochem. Biophys. 2003, 413, 151–157. [Google Scholar] [CrossRef]
  10. 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]
  11. Im, S.H.; Rao, A. Activation and deactivation of gene expression by Ca2+/calcineurin-NFAT-mediated signaling. Mol. Cells 2004, 18, 1–9. [Google Scholar]
  12. Rao, A.; Luo, C.; Hogan, P.G. Transcription factors of the NFAT family: Regulation and function. Annu. Rev. Immunol. 1997, 15, 707–747. [Google Scholar] [CrossRef] [PubMed]
  13. Okamura, H.; Aramburu, J.; Garcia-Rodriguez, C.; Viola, J.P.; 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]
  14. Macian, F.; Lopez-Rodriguez, C.; Rao, A. Partners in transcription: NFAT and AP-1. Oncogene 2001, 20, 2476–2489. [Google Scholar] [CrossRef] [Green Version]
  15. Lyakh, L.; Ghosh, P.; Rice, N.R. Expression of NFAT-family proteins in normal human T cells. Mol. Cell. Biol. 1997, 17, 2475–2484. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Masuda, E.S.; Naito, Y.; Tokumitsu, H.; Campbell, D.; Saito, F.; Hannum, C.; Arai, K.; Arai, N. NFATx, a novel member of the nuclear factor of activated T cells family that is expressed predominantly in the thymus. Mol. Cell. Biol. 1995, 15, 2697–2706. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. De la Pompa, J.L.; Timmerman, L.A.; Takimoto, H.; Yoshida, H.; Elia, A.J.; Samper, E.; Potter, J.; Wakeham, A.; Marengere, L.; Langille, B.L.; et al. Role of the NF-ATc transcription factor in morphogenesis of cardiac valves and septum. Nature 1998, 392, 182–186. [Google Scholar] [CrossRef]
  18. Graef, I.A.; Chen, F.; Chen, L.; Kuo, A.; Crabtree, G.R. Signals transduced by Ca2+/calcineurin and NFATc3/c4 pattern the developing vasculature. Cell 2001, 105, 863–875. [Google Scholar] [CrossRef] [Green Version]
  19. Mak, M.C.; Lam, K.M.; Chan, P.K.; Lau, Y.B.; Tang, W.H.; Yeung, P.K.; Ko, B.C.; Chung, S.M.; Chung, S.K. Embryonic lethality in mice lacking the nuclear factor of activated T cells 5 protein due to impaired cardiac development and function. PLoS ONE 2011, 6, e19186. [Google Scholar]
  20. López-Rodríguez, C.; Antos, C.L.; Shelton, J.M.; Richardson, J.A.; Lin, F.; Novobrantseva, T.I.; Bronson, R.T.; Igarashi, P.; Rao, A.; Olson, E.N. Loss of NFAT5 results in renal atrophy and lack of tonicity-responsive gene expression. Proc. Natl. Acad. Sci. USA 2004, 101, 2392–2397. [Google Scholar] [CrossRef] [Green Version]
  21. Xanthoudakis, S.; Viola, J.P.; Shaw, K.T.; Luo, C.; Wallace, J.D.; Bozza, P.T.; Luk, D.C.; Curran, T.; Rao, A. An enhanced immune response in mice lacking the transcription factor NFAT1. Science 1996, 272, 892–895. [Google Scholar] [CrossRef]
  22. Hodge, M.R.; Ranger, A.M.; Charles de la Brousse, F.; Hoey, T.; Grusby, M.J.; Glimcher, L.H. Hyperproliferation and dysregulation of IL-4 expression in NF-ATp-deficient mice. Immunity 1996, 4, 397–405. [Google Scholar] [CrossRef] [Green Version]
  23. Quadrato, G.; Benevento, M.; Alber, S.; Jacob, C.; Floriddia, E.M.; Nguyen, T.; Elnaggar, M.Y.; Pedroarena, C.M.; Molkentin, J.D.; Di Giovanni, S. Nuclear factor of activated T cells (NFATc4) is required for BDNF-dependent survival of adult-born neurons and spatial memory formation in the hippocampus. Proc. Natl. Acad. Sci. USA 2012, 109, E1499–E1508. [Google Scholar] [CrossRef] [Green Version]
  24. Molkentin, J.D.; Lu, J.R.; Antos, C.L.; Markham, B.; Richardson, J.; Robbins, J.; Grant, S.R.; Olson, E.N. A calcineurin-dependent transcriptional pathway for cardiac hypertrophy. Cell 1998, 93, 215–228. [Google Scholar] [CrossRef] [Green Version]
  25. Baksh, S.; Widlund, H.R.; Frazer-Abel, A.A.; Du, J.; Fosmire, S.; Fisher, D.E.; DeCaprio, J.A.; Modiano, J.F.; Burakoff, S.J. NFATc2-mediated repression of cyclin-dependent kinase 4 expression. Mol. Cell 2002, 10, 1071–1081. [Google Scholar] [CrossRef]
  26. Caetano, M.S.; Vieira-de-Abreu, A.; Teixeira, L.K.; Werneck, M.B.; Barcinski, M.A.; Viola, J.P. 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] [PubMed]
  27. Mognol, G.P.; de Araujo-Souza, P.S.; Robbs, B.K.; Teixeira, L.K.; Viola, J.P. Transcriptional regulation of the c-Myc promoter by NFAT1 involves negative and positive NFAT-responsive elements. Cell Cycle 2012, 11, 1014–1028. [Google Scholar] [CrossRef] [Green Version]
  28. Takayanagi, H.; Kim, S.; Koga, T.; Nishina, H.; Isshiki, M.; Yoshida, H.; Saiura, A.; Isobe, M.; Yokochi, T.; Inoue, J.; et al. Induction and activation of the transcription factor NFATc1 (NFAT2) integrate RANKL signaling in terminal differentiation of osteoclasts. Dev. Cell 2002, 3, 889–901. [Google Scholar] [CrossRef] [Green Version]
  29. Izawa, N.; Kurotaki, D.; Nomura, S.; Fujita, T.; Omata, Y.; Yasui, T.; Hirose, J.; Matsumoto, T.; Saito, T.; Kadono, Y.; et al. Cooperation of PU.1 With IRF8 and NFATc1 Defines Chromatin Landscapes During RANKL-Induced Osteoclastogenesis. J. Bone Miner. Res. 2019, 34, 1143–1154. [Google Scholar] [CrossRef]
  30. 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] [PubMed] [Green Version]
  31. 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]
  32. Suehiro, J.; Kanki, Y.; Makihara, C.; Schadler, K.; Miura, M.; Manabe, Y.; Aburatani, H.; Kodama, T.; Minami, T. Genome-wide approaches reveal functional vascular endothelial growth factor (VEGF)-inducible nuclear factor of activated T cells (NFAT) c1 binding to angiogenesis-related genes in the endothelium. J. Biol. Chem. 2014, 289, 29044–29059. [Google Scholar] [CrossRef] [Green Version]
  33. Mena, M.P.; Papiewska-Pajak, I.; Przygodzka, P.; Kozaczuk, A.; Boncela, J.; Cierniewski, C.S. NFAT2 regulates COX-2 expression and modulates the integrin repertoire in endothelial cells at the crossroads of angiogenesis and inflammation. Exp. Cell Res. 2014, 324, 124–136. [Google Scholar] [CrossRef] [PubMed]
  34. Shou, J.; Jing, J.; Xie, J.; You, L.; Jing, Z.; Yao, J.; Han, W.; Pan, H. Nuclear factor of activated T cells in cancer development and treatment. Cancer Lett. 2015, 361, 174–184. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Le Roy, C.; Deglesne, P.A.; Chevallier, N.; Beitar, T.; Eclache, V.; Quettier, M.; Boubaya, M.; Letestu, R.; Lévy, V.; Ajchenbaum-Cymbalista, F.; et al. The degree of BCR and NFAT activation predicts clinical outcomes in chronic lymphocytic leukemia. Blood 2012, 120, 356–365. [Google Scholar] [CrossRef] [Green Version]
  36. Tripathi, M.K.; Deane, N.G.; Zhu, J.; An, H.; Mima, S.; Wang, X.; Padmanabhan, S.; Shi, Z.; Prodduturi, N.; Ciombor, K.K.; et al. Nuclear factor of activated T-cell activity is associated with metastatic capacity in colon cancer. Cancer Res. 2014, 74, 6947–6957. [Google Scholar] [CrossRef] [Green Version]
  37. Metzelder, S.K.; Michel, C.; von Bonin, M.; Rehberger, M.; Hessmann, E.; Inselmann, S.; Solovey, M.; Wang, Y.; Sohlbach, K.; Brendel, C.; et al. NFATc1 as a therapeutic target in FLT3-ITD-positive AML. Leukemia 2015, 29, 1470–1477. [Google Scholar] [CrossRef] [PubMed]
  38. Lucena, P.I.; Faget, D.V.; Pachulec, E.; Robaina, M.C.; Klumb, C.E.; Robbs, B.K.; Viola, J.P. NFAT2 Isoforms Differentially Regulate Gene Expression, Cell Death, and Transformation through Alternative N-Terminal Domains. Mol. Cell. Biol. 2016, 36, 119–131. [Google Scholar] [CrossRef] [Green Version]
  39. Marafioti, T.; Pozzobon, M.; Hansmann, M.L.; Ventura, R.; Pileri, S.A.; Roberton, H.; Gesk, S.; Gaulard, P.; Barth, T.F.; 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] [PubMed]
  40. Shakhova, I.; Li, Y.; Yu, F.; Kaneko, Y.; Nakamura, Y.; Ohira, M.; Izumi, H.; Mae, T.; Varfolomeeva, S.R.; Rumyantsev, A.G.; et al. PPP3CB contributes to poor prognosis through activating nuclear factor of activated T-cells signaling in neuroblastoma. Mol. Carcinog. 2019, 58, 426–435. [Google Scholar] [CrossRef]
  41. Robbs, B.K.; Cruz, A.L.; Werneck, M.B.; Mognol, G.P.; Viola, J.P. Dual roles for NFAT transcription factor genes as oncogenes and tumor suppressors. Mol. Cell. Biol. 2008, 28, 7168–7181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. May, S.L.; Zhou, Q.; Lewellen, M.; Carter, C.M.; Coffey, D.; Highfill, S.L.; Bucher, C.M.; Matise, I.; Morse, H.C., 3rd; O’Sullivan, M.G.; et al. Nfatc2 and Tob1 have non-overlapping function in T cell negative regulation and tumorigenesis. PLoS ONE 2014, 9, e100629. [Google Scholar] [CrossRef] [PubMed]
  43. Chen, Z.L.; Zhao, S.H.; Wang, Z.; Qiu, B.; Li, B.Z.; Zhou, F.; Tan, X.G.; He, J. Expression and unique functions of four nuclear factor of activated T cells isoforms in non-small cell lung cancer. Chin. J. Cancer 2011, 30, 62–68. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Baumgart, S.; Glesel, E.; Singh, G.; Chen, N.M.; Reutlinger, K.; Zhang, J.; Billadeau, D.D.; Fernandez-Zapico, M.E.; Gress, T.M.; Singh, S.K.; et al. Restricted heterochromatin formation links NFATc2 repressor activity with growth promotion in pancreatic cancer. Gastroenterology 2012, 142, e1–e7. [Google Scholar] [CrossRef] [Green Version]
  45. Zaichuk, T.A.; Shroff, E.H.; Emmanuel, R.; Filleur, S.; Nelius, T.; Volpert, O.V. Nuclear factor of activated T cells balances angiogenesis activation and inhibition. J. Exp. Med. 2004, 199, 1513–1522. [Google Scholar] [CrossRef]
  46. Urso, K.; Fernández, A.; Velasco, P.; Cotrina, J.; de Andrés, B.; Sánchez-Gómez, P.; Hernández-Laín, A.; Hortelano, S.; Redondo, J.M.; Cano, E. NFATc3 controls tumour growth by regulating proliferation and migration of human astroglioma cells. Sci. Rep. 2019, 9, 9361. [Google Scholar] [CrossRef]
  47. Xiao, T.; Zhu, J.J.; Huang, S.; Peng, C.; He, S.; Du, J.; Hong, R.; Chen, X.; Bode, A.M.; Jiang, W.; et al. Phosphorylation of NFAT3 by CDK3 induces cell transformation and promotes tumor growth in skin cancer. Oncogene 2017, 36, 2835–2845. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Fougere, M.; Gaudineau, B.; Barbier, J.; Guaddachi, F.; Feugeas, J.P.; Auboeuf, D.; Jauliac, S. NFAT3 transcription factor inhibits breast cancer cell motility by targeting the Lipocalin 2 gene. Oncogene 2010, 29, 2292–2301. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Glud, S.Z.; Sørensen, A.B.; Andrulis, M.; Wang, B.; Kondo, E.; Jessen, R.; Krenacs, L.; Stelkovics, E.; Wabl, M.; Serfling, E.; et al. A tumor-suppressor function for NFATc3 in T-cell lymphomagenesis by murine leukemia virus. Blood 2005, 106, 3546–3552. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Abdul, H.M.; Sama, M.A.; Furman, J.L.; Mathis, D.M.; Beckett, T.L.; Weidner, A.M.; Patel, E.S.; Baig, I.; Murphy, M.P.; LeVine, H., 3rd; et al. Cognitive decline in Alzheimer’s disease is associated with selective changes in calcineurin/NFAT signaling. J. Neurosci. 2009, 29, 12957–12969. [Google Scholar] [CrossRef]
  51. Hudry, E.; Wu, H.Y.; Arbel-Ornath, M.; Hashimoto, T.; Matsouaka, R.; Fan, Z.; Spires-Jones, T.L.; Betensky, R.A.; Bacskai, B.J.; Hyman, B.T. Inhibition of the NFAT pathway alleviates amyloid β neurotoxicity in a mouse model of Alzheimer’s disease. J. Neurosci. 2012, 32, 3176–3192. [Google Scholar] [CrossRef] [Green Version]
  52. Luo, J.; Sun, L.; Lin, X.; Liu, G.; Yu, J.; Parisiadou, L.; Xie, C.; Ding, J.; Cai, H. A calcineurin- and NFAT-dependent pathway is involved in α-synuclein-induced degeneration of midbrain dopaminergic neurons. Hum. Mol. Genet. 2014, 23, 6567–6574. [Google Scholar] [CrossRef] [Green Version]
  53. Ranger, A.M.; Oukka, M.; Rengarajan, J.; Glimcher, L.H. Inhibitory function of two NFAT family members in lymphoid homeostasis and Th2 development. Immunity 1998, 9, 627–635. [Google Scholar] [CrossRef] [Green Version]
  54. Wu, Y.; Borde, M.; Heissmeyer, V.; Feuerer, M.; Lapan, A.D.; Stroud, J.C.; Bates, D.L.; Guo, L.; Han, A.; Ziegler, S.F.; et al. FOXP3 controls regulatory T cell function through cooperation with NFAT. Cell 2006, 126, 375–387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Gibson, H.M.; Hedgcock, C.J.; Aufiero, B.M.; Wilson, A.J.; Hafner, M.S.; Tsokos, G.C.; Wong, H.K. Induction of the CTLA-4 gene in human lymphocytes is dependent on NFAT binding the proximal promoter. J. Immunol. 2007, 179, 3831–3840. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Shin, D.S.; Jordan, A.; Basu, S.; Thomas, R.M.; Bandyopadhyay, S.; de Zoeten, E.F.; Wells, A.D.; Macian, F. Regulatory T cells suppress CD4+ T cells through NFAT-dependent transcriptional mechanisms. EMBO Rep. 2014, 15, 991–999. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Bopp, T.; Palmetshofer, A.; Serfling, E.; Heib, V.; Schmitt, S.; Richter, C.; Klein, M.; Schild, H.; Schmitt, E.; Stassen, M. NFATc2 and NFATc3 transcription factors play a crucial role in suppression of CD4+ T lymphocytes by CD4+ CD25+ regulatory T cells. J. Exp. Med. 2005, 201, 181–187. [Google Scholar] [CrossRef] [Green Version]
  58. Serfling, E.; Klein-Hessling, S.; Palmetshofer, A.; Bopp, T.; Stassen, M.; Schmitt, E. NFAT transcription factors in control of peripheral T cell tolerance. Eur. J. Immunol. 2006, 36, 2837–2843. [Google Scholar] [CrossRef]
  59. Kwon, H.K.; Kim, G.C.; Hwang, J.S.; Kim, Y.; Chae, C.S.; Nam, J.H.; Jun, C.D.; Rudra, D.; Surh, C.D.; Im, S.H. Transcription factor NFAT1 controls allergic contact hypersensitivity through regulation of activation induced cell death program. Sci. Rep. 2016, 6, 19453. [Google Scholar] [CrossRef] [Green Version]
  60. 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]
  61. Liu, Z.; Lee, J.; Krummey, S.; Lu, W.; Cai, H.; Lenardo, M.J. The kinase LRRK2 is a regulator of the transcription factor NFAT that modulates the severity of inflammatory bowel disease. Nat. Immunol. 2011, 12, 1063–1070. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Dietz, L.; Frommer, F.; Vogel, A.L.; Vaeth, M.; Serfling, E.; Waisman, A.; Buttmann, M.; Berberich-Siebelt, F. NFAT1 deficit and NFAT2 deficit attenuate EAE via different mechanisms. Eur. J. Immunol. 2015, 45, 1377–1389. [Google Scholar] [CrossRef]
  63. Choi, S.; You, S.; Kim, D.; Choi, S.Y.; Kwon, H.M.; Kim, H.S.; Hwang, D.; Park, Y.J.; Cho, C.S.; Kim, W.U. Transcription factor NFAT5 promotes macrophage survival in rheumatoid arthritis. J. Clin. Investig. 2017, 127, 954–969. [Google Scholar] [CrossRef]
  64. Kleinewietfeld, M.; Manzel, A.; Titze, J.; Kvakan, H.; Yosef, N.; Linker, R.A.; Muller, D.N.; Hafler, D.A. Sodium chloride drives autoimmune disease by the induction of pathogenic TH17 cells. Nature 2013, 496, 518–522. [Google Scholar] [CrossRef] [PubMed]
  65. Taglialatela, G.; Rastellini, C.; Cicalese, L. Reduced Incidence of Dementia in Solid Organ Transplant Patients Treated with Calcineurin Inhibitors. J. Alzheimers. Dis. 2015, 47, 329–333. [Google Scholar] [CrossRef] [Green Version]
  66. Kiani, A.; Rao, A.; Aramburu, J. Manipulating immune responses with immunosuppressive agents that target NFAT. Immunity 2000, 12, 359–372. [Google Scholar] [CrossRef] [Green Version]
  67. 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] [PubMed]
  68. Fung, J.J.; Alessiani, M.; Abu-Elmagd, K.; Todo, S.; Shapiro, R.; Tzakis, A.; Van Thiel, D.; Armitage, J.; Jain, A.; McCauley, J.; et al. Adverse effects associated with the use of FK 506. Transplant. Proc. 1991, 23, 3105–3108. [Google Scholar]
  69. Jin, L.; Harrison, S.C. Crystal structure of human calcineurin complexed with cyclosporin A and human cyclophilin. Proc. Natl. Acad. Sci. USA 2002, 99, 13522–13526. [Google Scholar] [CrossRef] [Green Version]
  70. 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]
  71. Cyert, M.S. Calcineurin signaling in Saccharomyces cerevisiae: How yeast go crazy in response to stress. Biochem. Biophys. Res. Commun. 2003, 311, 1143–1150. [Google Scholar] [CrossRef]
  72. Werlen, G.; Jacinto, E.; Xia, Y.; Karin, M. Calcineurin preferentially synergizes with PKC-q to activate JNK and IL-2 promoter in T lymphocytes. EMBO J. 1998, 17, 3101–3111. [Google Scholar] [CrossRef] [PubMed]
  73. Trushin, S.A.; Pennington, K.N.; Algeciras-Schimnich, A.; Paya, C.V. Protein kinase C and calcineurin synergize to activate IkappaB kinase and NF-kappaB in T lymphocytes. J. Biol. Chem. 1999, 274, 22923–22931. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Aramburu, J.; Yaffe, M.B.; Lopez-Rodriguez, C.; Cantley, L.C.; Hogan, P.G.; Rao, A. Affinity-driven peptide selection of an NFAT inhibitor more selective than cyclosporin A. Science 1999, 285, 2129–2133. [Google Scholar] [CrossRef] [PubMed]
  75. Noguchi, H.; Matsushita, M.; Okitsu, T.; Moriwaki, A.; Tomizawa, K.; Kang, S.; Li, S.T.; Kobayashi, N.; Matsumoto, S.; Tanaka, K.; et al. A new cell-permeable peptide allows successful allogeneic islet transplantation in mice. Nat. Med. 2004, 10, 305–309. [Google Scholar] [CrossRef] [PubMed]
  76. Kuriyama, M.; Matsushita, M.; Tateishi, A.; Moriwaki, A.; Tomizawa, K.; Ishino, K.; Sano, S.; Matsui, H. A cell-permeable NFAT inhibitor peptide prevents pressure-overload cardiac hypertrophy. Chem. Biol. Drug Des. 2006, 67, 238–243. [Google Scholar] [CrossRef] [Green Version]
  77. Elloumi, H.Z.; Maharshak, N.; Rao, K.N.; Kobayashi, T.; Ryu, H.S.; Mühlbauer, M.; Li, F.; Jobin, C.; Plevy, S.E. A cell permeable peptide inhibitor of NFAT inhibits macrophage cytokine expression and ameliorates experimental colitis. PLoS ONE 2012, 7, e34172. [Google Scholar] [CrossRef] [Green Version]
  78. Choi, J.M.; Sohn, J.H.; Park, T.Y.; Park, J.W.; Lee, S.K. Cell permeable NFAT inhibitory peptide Sim-2-VIVIT inhibits T-cell activation and alleviates allergic airway inflammation and hyper-responsiveness. Immunol. Lett. 2012, 143, 170–176. [Google Scholar] [CrossRef]
  79. Zhang, L.; Li, R.; Shi, W.; Liang, X.; Liu, S.; Ye, Z.; Yu, C.; Chen, Y.; Zhang, B.; Wang, W.; et al. NFAT2 inhibitor ameliorates diabetic nephropathy and podocyte injury in db/db mice. Br. J. Pharmacol. 2013, 170, 426–439. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Garcia-Vaz, E.; McNeilly, A.D.; Berglund, L.M.; Ahmad, A.; Gallagher, J.R.; Dutius Andersson, A.M.; McCrimmon, R.J.; Zetterqvist, A.V.; Gomez, M.F.; Khan, F. Inhibition of NFAT Signaling Restores Microvascular Endothelial Function in Diabetic Mice. Diabetes 2020, 69, 424–435. [Google Scholar] [CrossRef] [Green Version]
  81. Bretz, C.A.; Savage, S.; Capozzi, M.; Penn, J.S. The role of the NFAT signaling pathway in retinal neovascularization. Investig. Ophthalmol. Vis. Sci. 2013, 54, 7020–7027. [Google Scholar] [CrossRef] [PubMed]
  82. Ranger, A.M.; Hodge, M.R.; Gravallese, E.M.; Oukka, M.; Davidson, L.; Alt, F.W.; de la Brousse, F.C.; Hoey, T.; Grusby, M.; Glimcher, L.H. Delayed lymphoid repopulation with defects in IL-4-driven responses produced by inactivation of NF-ATc. Immunity 1998, 8, 125–134. [Google Scholar] [CrossRef] [Green Version]
  83. Yoshida, H.; Nishina, H.; Takimoto, H.; Marengere, L.E.; Wakeham, A.C.; Bouchard, D.; Kong, Y.Y.; Ohteki, T.; Shahinian, A.; Bachmann, M.; et al. The transcription factor NF-ATc1 regulates lymphocyte proliferation and Th2 cytokine production. Immunity 1998, 8, 115–124. [Google Scholar] [CrossRef] [Green Version]
  84. Viola, J.P.; Kiani, A.; Bozza, P.T.; Rao, A. Regulation of allergic inflammation and eosinophil recruitment in mice lacking the transcription factor NFAT1: Role of interleukin-4 (IL-4) and IL-5. Blood 1998, 91, 2223–2230. [Google Scholar] [CrossRef] [PubMed]
  85. Erb, K.J.; Twardzik, T.; Palmetshofer, A.; Wohlleben, G.; Tatsch, U.; Serfling, E. Mice deficient in nuclear factor of activated T-cell transcription factor c2 mount increased Th2 responses after infection with Nippostrongylus brasiliensis and decreased Th1 responses after mycobacterial infection. Infect. Immun. 2003, 71, 6641–6647. [Google Scholar] [CrossRef] [Green Version]
  86. Kaminuma, O.; Kitamura, F.; Kitamura, N.; Hiroi, T.; Miyoshi, H.; Miyawaki, A.; Miyatake, S. Differential contribution of NFATc2 and NFATc1 to TNF-alpha gene expression in T cells. J. Immunol. 2008, 180, 319–326. [Google Scholar] [CrossRef] [PubMed]
  87. Tsai, E.Y.; Jain, J.; Pesavento, P.A.; Rao, A.; Goldfeld, A.E. Tumor necrosis factor alpha gene regulation in activated T cells involves ATF-2/Jun and NFATp. Mol. Cell. Biol. 1996, 16, 459–467. [Google Scholar] [CrossRef] [Green Version]
  88. Kaminuma, O.; Kitamura, N.; Nishito, Y.; Nemoto, S.; Tatsumi, H.; Mori, A.; Hiroi, T. Downregulation of NFAT3 Due to Lack of T-Box Transcription Factor TBX5 Is Crucial for Cytokine Expression in T Cells. J. Immunol. 2018, 200, 92–100. [Google Scholar] [CrossRef]
  89. Kar, P.; Mirams, G.R.; Christian, H.C.; Parekh, A.B. Control of NFAT Isoform Activation and NFAT-Dependent Gene Expression through Two Coincident and Spatially Segregated Intracellular Ca2+ Signals. Mol. Cell 2016, 64, 746–759. [Google Scholar] [CrossRef] [Green Version]
  90. Garcia-Cozar, F.J.; Okamura, H.; Aramburu, J.F.; Shaw, K.T.; Pelletier, L.; Showalter, R.; Villafranca, E.; Rao, A. Two-site interaction of nuclear factor of activated T cells with activated calcineurin. J. Biol. Chem. 1998, 273, 23877–23883. [Google Scholar] [CrossRef] [Green Version]
  91. Park, S.; Uesugi, M.; Verdine, G.L. A second calcineurin binding site on the NFAT regulatory domain. Proc. Natl. Acad. Sci. USA 2000, 97, 7130–7135. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Kitamura, N.; Shindo, M.; Ohtsuka, J.; Nakamura, A.; Tanokura, M.; Hiroi, T.; Kaminuma, O. Identification of novel interacting regions involving calcineurin and nuclear factor of activated T cells. FASEB J. 2020, 34, 3197–3208. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Miyata, H.; Satouh, Y.; Mashiko, D.; Muto, M.; Nozawa, K.; Shiba, K.; Fujihara, Y.; Isotani, A.; Inaba, K.; Ikawa, M. Sperm calcineurin inhibition prevents mouse fertility with implications for male contraceptive. Science 2015, 350, 442–445. [Google Scholar] [CrossRef] [PubMed]
  94. Xiao, Y.; Qureischi, M.; Dietz, L.; Vaeth, M.; Vallabhapurapu, S.D.; Klein-Hessling, S.; Klein, M.; Liang, C.; König, A.; Serfling, E.; et al. Lack of NFATc1 SUMOylation prevents autoimmunity and alloreactivity. J. Exp. Med. 2021, 218, e20181853. [Google Scholar] [CrossRef]
  95. Vaeth, M.; Schliesser, U.; Müller, G.; Reissig, S.; Satoh, K.; Tuettenberg, A.; Jonuleit, H.; Waisman, A.; Müller, M.R.; Serfling, E.; et al. Dependence on nuclear factor of activated T-cells (NFAT) levels discriminates conventional T cells from Foxp3+ regulatory T cells. Proc. Natl. Acad. Sci. USA 2012, 109, 16258–16263. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Differences in the structure, functional domains, and amino acid sequences of nuclear factor of activated T cell (NFAT) family members.
Figure 1. Differences in the structure, functional domains, and amino acid sequences of nuclear factor of activated T cell (NFAT) family members.
Ijms 22 02725 g001
Figure 2. The different functional role of NFATc4 in aortic smooth muscle cells (ASMCs) and T cells through T-box factor 5 (TBX5)-mediated transcriptional regulation. CASP1; caspase 1, CD163L1; CD163 molecule like 1, EREG; epiregulin, GMPR; guanosine monophosphate reductase, TBE; T-box factor binding element, TNFAIP2; tumor necrosis factor-alpha-induced protein 2.
Figure 2. The different functional role of NFATc4 in aortic smooth muscle cells (ASMCs) and T cells through T-box factor 5 (TBX5)-mediated transcriptional regulation. CASP1; caspase 1, CD163L1; CD163 molecule like 1, EREG; epiregulin, GMPR; guanosine monophosphate reductase, TBE; T-box factor binding element, TNFAIP2; tumor necrosis factor-alpha-induced protein 2.
Ijms 22 02725 g002
Figure 3. Putative three-dimensional structure of calcineurin A (CNA; green) and NFATc1 (pink) complex. The amino acids essential for the complex formation between NFATc1 (Cys263) and CNA (Asp82) are indicated by the red frame.
Figure 3. Putative three-dimensional structure of calcineurin A (CNA; green) and NFATc1 (pink) complex. The amino acids essential for the complex formation between NFATc1 (Cys263) and CNA (Asp82) are indicated by the red frame.
Ijms 22 02725 g003
Figure 4. Schematic points of action of selective and non-selective NFAT inhibitors. CN; calcineurin, CNBR3; CN-binding region 3.
Figure 4. Schematic points of action of selective and non-selective NFAT inhibitors. CN; calcineurin, CNBR3; CN-binding region 3.
Ijms 22 02725 g004
Table 1. Functions and diseases associated with NFAT family members.
Table 1. Functions and diseases associated with NFAT family members.
Isoform
  • Regulation
  • Intracellular Localization
Involvement in Organ/Tissue Formation or Cell FunctionAbnormal NFAT Activity
HumanKnockout Mouse
NFATc1
  • Calcineurin
  • Cytoplasm (in resting cells) Nucleus (in activated cells)
Heart valve/septum formation [5,17], angiogenesis [32,33], osteoclast formation [28], T cell proliferation/Th2 differentiation [82,83]Pancreatic cancer [31], colon cancer (increased risk of metastasis) [36], leukemia [35,37], lymphoma [39]Embryonic lethality (circulatory failure) [5]
NFATc2Cell cycle arrest/cell growth inhibition (normal cells) [26,41], activation-induced cell death [59], regulation of angiogenesis [45]Pancreatic cancer [44], lung adenocarcinoma (> stage II) [43], mild cognitive impairment [50]Increased immune reactivity [21,22], allergic contact hypersensitivity [59], lymphoma [42], lymphocyte hyperplasia (NFATc2/c3 DKO) [53],
NFATc3-Lung squamous cell carcinoma (well-differentiated cancer) [43], glioma [46]lymphocyte hyperplasia (NFATc2/c3 DKO) [53], embryonic lethality (vascular hypoplasia, NFATc3/c4 DKO) [18]
NFATc4Nervous system assembly/spatial memory formation
[6,23]
Skin cancer [47], severe cognitive impairment [50], Alzheimer’s disease [50]Spatial memory dysplasia [23], embryonic lethality (vascular hypoplasia, NFATc3/c4 DKO) [18]
NFAT5
  • mainly osmotic pressure
  • cytoplasm and nucleus
Cell protection under hypertonic stress [8], homeostasis [9]Rheumatoid arthritis [63]Mostly embryonic lethality (decreased cardiac function, renal atrophy) [19,20]
DKO: double knock-out.
Table 2. Contribution of calcineurin (CN) binding regions (CNBRs) to CN/NFAT interaction.
Table 2. Contribution of calcineurin (CN) binding regions (CNBRs) to CN/NFAT interaction.
Domain/RegionNFATc1NFATc2NFATc3NFATc4
CRD++++++++++++
CNBR1+++++++++
CNBR2+++++++
CNBR3++++ 1
1 Affinity to CNA (Kd, μM): <0.1; +++, 0.1 < 1; ++, 1 < 10; +, 10<; −. CRD: calcium regulatory domain.
Table 3. Current and potential application of NFAT inhibitors.
Table 3. Current and potential application of NFAT inhibitors.
ApplicationCyclosporin TacrolimusDirect NFAT InhibitorCNBR3-Targeted Inhibitor
Transplantation
Atopic dermatitis
Bronchial asthma
Rheumatoid arthritis
Ulcerative colitis
Lupus nephritis
Myasthenia gravis
Behcet’s disease
Regenerative anemia
Psoriasis
Nephrotic syndrome
Kawasaki disease
Alzheimer’s disease
Cardiac hypertrophy
Type2 diabetes
Osteoporosis
Cancer
Retinopathy
Existing and potential side effectsSeverModerateMild
○: clinically approved, △: experimentally proven, ■: suggested.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kitamura, N.; Kaminuma, O. Isoform-Selective NFAT Inhibitor: Potential Usefulness and Development. Int. J. Mol. Sci. 2021, 22, 2725. https://doi.org/10.3390/ijms22052725

AMA Style

Kitamura N, Kaminuma O. Isoform-Selective NFAT Inhibitor: Potential Usefulness and Development. International Journal of Molecular Sciences. 2021; 22(5):2725. https://doi.org/10.3390/ijms22052725

Chicago/Turabian Style

Kitamura, Noriko, and Osamu Kaminuma. 2021. "Isoform-Selective NFAT Inhibitor: Potential Usefulness and Development" International Journal of Molecular Sciences 22, no. 5: 2725. https://doi.org/10.3390/ijms22052725

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