Skip to main content

REVIEW article

Front. Immunol., 18 March 2015
Sec. T Cell Biology
This article is part of the Research Topic T cell regulation by the environment View all 11 articles

Vitamin D Actions on CD4+ T Cells in Autoimmune Disease

  • 1Department of Biochemistry, College of Agricultural and Life Sciences, University of Wisconsin-Madison, Madison, WI, USA
  • 2Department of Statistics, College of Letters and Sciences, University of Wisconsin-Madison, Madison, WI, USA

This review summarizes and integrates research on vitamin D and CD4+ T-lymphocyte biology to develop new mechanistic insights into the molecular etiology of autoimmune disease. A deep understanding of molecular mechanisms relevant to gene–environment interactions is needed to deliver etiology-based autoimmune disease prevention and treatment strategies. Evidence linking sunlight, vitamin D, and the risk of multiple sclerosis and type 1 diabetes is summarized to develop the thesis that vitamin D is the environmental factor that most strongly influences autoimmune disease development. Evidence for CD4+ T-cell involvement in autoimmune disease pathogenesis and for paracrine calcitriol signaling to CD4+ T lymphocytes is summarized to support the thesis that calcitriol is sunlight’s main protective signal transducer in autoimmune disease risk. Animal modeling and human mechanistic data are summarized to support the view that vitamin D probably influences thymic negative selection, effector Th1 and Th17 pathogenesis and responsiveness to extrinsic cell death signals, FoxP3+CD4+ T-regulatory cell and CD4+ T-regulatory cell type 1 (Tr1) cell functions, and a Th1–Tr1 switch. The proposed Th1–Tr1 switch appears to bridge two stable, self-reinforcing immune states, pro- and anti-inflammatory, each with a characteristic gene regulatory network. The bi-stable switch would enable T cells to integrate signals from pathogens, hormones, cell–cell interactions, and soluble mediators and respond in a biologically appropriate manner. Finally, unanswered questions and potentially informative future research directions are highlighted to speed delivery of etiology-based strategies to reduce autoimmune disease.

Introduction

Autoimmune diseases afflict ~50 million Americans and contribute >$100 billion to US health care costs (1). The global burden has risen with the near tripling in the last half-century of multiple sclerosis (MS) (2, 3), type 1 diabetes (T1D) (4), and other autoimmune diseases. A deep understanding of disease mechanisms will be needed to deliver etiology-based strategies to reverse this vexing trend. Indeed, “functional and mechanistic work on the molecular etiology of disease remains one of the major challenges in modern biology” (5).

This review highlights recent advances in vitamin D and T-lymphocyte biology in an effort to harness vitamin D’s potential to reduce the impact of autoimmune diseases. Gene–environment interactions, sunlight and vitamin D, and T lymphocytes as autoimmune disease initiators and vitamin D targets are discussed to explain the rationale for reviewing vitamin D mechanisms in T cells. Research on vitamin D regulation of thymocyte selection, Th1 and Th17 cells, T-cell programed cell death, and T-regulatory (Treg) cells is summarized and integrated into model mechanisms. Finally, unanswered questions relating to vitamin D mechanisms in CD4+ T cells are highlighted to promote further research that may lead to a deeper understanding of autoimmune disease molecular etiology.

Genes, Sunlight, Vitamin D, and T Lymphocytes

Autoimmune Diseases

Autoimmune diseases represent a failure of self-identification leading to an immune-mediated assault on host tissues. More than 100 autoimmune syndromes exist (1). We drew mainly on MS and T1D research because intensive investigation has generated detailed insights into vitamin D mechanisms in these diseases and provided valuable guidance for research on other autoimmune diseases. Other autoimmune diseases are included where robust mechanistic data exist. A recent chapter (6) and a review (7) have summarized vitamin D mechanisms in autoimmune diseases more generally.

Multiple sclerosis and T1D have distinct target organs, genetic risk factors, onset ages, and female to male ratios, but target organ-specific T cells as initiators unite these diseases. MS is the leading cause of non-traumatic neurological disability in young adults. It results from an autoimmune attack on the axon–myelin unit (8). Neurological dysfunction in MS is attributed to focal demyelinated lesions in the central nervous system (CNS). The neurodegenerative process is believed to occur sub-clinically before the disease is typically diagnosed in the third decade of life. Most MS patients have a relapsing-remitting form of the disease, and among them, women outnumber men 3:1.

Type 1 diabetes is a common chronic disease of childhood with an onset typically between ages 6 and 15. Boys with T1D slightly outnumber girls. T1D results from an autoimmune attack on insulin-producing pancreatic β-cells (9). Metabolic dysfunction in T1D is attributed to β-cell destruction and loss of insulin production and blood glucose control. The β-cell degenerative process is also believed to occur sub-clinically before a T1D diagnosis, when as much as 70% of β-cell mass has been destroyed and insulin supply no longer meets demand (10). We searched for analogies between MS and T1D that might reveal over-arching environmental influences on T lymphocytes in autoimmunity.

Gene–Environment Interactions in Autoimmune Disease

Interacting genetic, environmental, and hormonal influences drive autoimmune disease. There is a weak genetic component in autoimmune disease susceptibility. MS risk increases a modest 2–6% among first-degree biological relatives of an MS index case (11), but candidate risk loci are of such modest aggregate influence that genetic burden scores cannot accurately predict disease (12, 13). Autoimmune disease discordance between monozygotic twins is generally between 60 and 75% and cannot be explained by somatic mosaicism in MS (14), T1D (15), Crohn’s disease (CD) (16, 17), or systemic lupus erythematosus (SLE) (18). The low penetrance of candidate risk loci, high monozygotic twin discordance rates, and rapidly rising disease incidence rates underscore the hypothesis that an autoimmune disease phenotype emerges when modifiable environmental stressors act on a disease-susceptible genotype, and exposure to at least one environmental stressor is increasing. The key questions are (i) what are the dominant, modifiable environmental stressors, (ii) by what mechanisms do they interact with susceptibility genes to propel the disease process, and (iii) can mechanistic knowledge of gene–environment interactions be exploited to craft effective, etiology-based disease prevention and treatment strategies.

Sunlight Exposure and Risk of MS and T1D Disease

There is a large, latitude-linked, non-transmissible environmental component acting in a female-biased manner at the population level that determines whether the MS phenotype emerges from an MS-susceptible genotype (19, 20). Supporting this interpretation are the gradient in MS prevalence as a function of latitude (21, 22), alteration of MS risk by migration (23), and the equivalent MS risk between non-biological relatives of an MS case and the general population (24). The latitudinal gradient has dissipated as incidence has risen in low latitude regions. Peak MS prevalence was ~130/105 population in 1960 (21) and is now ~400/105 (22). Young women have borne the brunt of rising MS incidence (25, 26). The rise in female MS incidence implies a significant proportion of new female MS cases may be preventable (27).

Low sunlight exposure is postulated to be the major latitude-linked component in MS risk (21). Ultraviolet (UV) irradiance varies 400-fold with latitude (28), correlating inversely with the ~400-fold variation in MS prevalence (22). UV irradiance also varies seasonally at high latitudes, where increases in MS disease activity have been noted lagging seasonal declines in ambient UV light (2931). Furthermore, childhood and occupational sunlight exposure correlated inversely with MS disease risk (3235) and mortality (36).

Low sunlight exposure is also postulated to be a major component in T1D risk. Globally, T1D incidence varied ~350-fold (37) correlating inversely with the 400-fold latitudinal variation in UV irradiance (28). Global T1D incidence is increasing, while the proportion of T1D patients with the major HLA risk genotype is decreasing (13), implicating a modifiable environmental factor. T1D onset peaked between October and January and reached a nadir between June and August in the northern hemisphere, with a reverse pattern in the southern hemisphere (38). This correlation disappeared after adjustment for latitude. The inverse correlation between ambient winter UV radiation and T1D (r = −0.80) (39) was nearly as robust as that for MS (r = −0.9) (21). Thus, ambient UV irradiance is the leading candidate for the large, latitude-linked environmental risk factor in both MS and T1D (40, 41).

The correlative data are consistent with but do not prove that low sunlight exposure is a major component in autoimmune disease risk. Other environmental risk factors have been proposed (42), but they have markedly smaller effect sizes. For example, symptomatic Epstein-Barr virus (EBV) infection in adolescence (mononucleosis), the strongest of the environmental risk factors that appear to act independently of latitude, season, and UV light (43), correlates with a two to threefold increase in MS risk (44). In sharp contrast, residence in a low ambient UV light region correlates with a >100-fold increase in MS risk (22). Whether EBV infection and UV light are completely independent risk factors is currently debated (45). As discussed below, genetic data suggest the VDR gene influences HLA-DRB1 presentation of peptides to CD4+ T lymphocytes, and structural data show pathogenic T cells did not distinguish a DRB1*1501-restricted myelin peptide from a DRB5*0101-restricted EBV peptide suggesting a molecular mimicry mechanism underlying pathogenesis. In summary, the UV-linked component exerts its influence on multiple autoimmune diseases and on populations with disparate ancestries, distinct dietary and smoking habits, and dissimilar exposures to infectious and commensal organisms. These qualities of strength and universality provide the rationale for close investigation of this component’s identity and mechanisms.

Vitamin D and Transduction of Sunlight’s Protective Signals in MS and T1D

Vitamin D3 and its hormonal form, 1α,25-dihydroxyvitamin D3 (calcitriol) have been proposed as the major biological transducers of sunlight’s protective signals in autoimmune disease due to calcitriol’s ability to selectively regulate T-cell-mediated autoimmune responses (46). Evidence supporting this hypothesis has been reviewed (6, 7, 42, 47, 48). Persuasive evidence for vitamin D as sunlight’s signal transducer derives from genetic linkage studies. Rare loss-of-function mutations in the CYP27B1 gene correlated with a significantly increased autoimmune disease risk. This association was first reported for T1D (4954), Addison’s disease (55), Hashimoto’s thyroiditis, and Graves’ disease (56). It was subsequently reported for MS (5760). In rare multi-incident MS families, 35 of 35 cases inherited one defective CYP27B1 allele, an inheritance pattern with small odds (one in a billion) of occurring by chance (58). Because CYP27B1 mutations are highly penetrant but exceedingly rare, they do not contribute genetic risk in the vast majority of disease cases. In fact, genome-wide association studies (GWAS) and some case–control studies did not detect an association between CYP27B1 variants and MS or T1D (6165). However, the replicated positive genetic findings indelibly mark calcitriol synthesis as a key determinant of MS and T1D risk.

Correlations between VDR alleles and MS susceptibility have also been reported (6668). An early study found a VDR and MS association in patients who carried the high-risk HLA-DRB1*1501 allele (69). Later research identified a putative vitamin D-responsive element (VDRE; see below) in the HLA-DRB1*1501 promoter (70). The VDR association data have been inconsistent between populations, and some VDR polymorphisms studied do not have known functional impacts. The VDR FokI polymorphism is an exception; VDRF (without the FokI site) encodes a 424 amino acid protein with higher transcriptional activity than the 427 amino acid protein encoded by VDRf (a FokI site in the first ATG codon) (71, 72). The less active VDRf protein was associated with higher serum 25-hydroxyvitamin D (25-OHD) levels, lower MS risk, and lower MS disability (7376). Another exception is the Cdx-2G variant, which has 70% reduced promoter activity (77); this variant correlated with an increased risk of MS in children who had ≤2 h/day of winter sun exposure (78). GWAS did not detect a VDR and MS association (61).

Some family studies have also detected linkage between VDR polymorphisms and T1D, but concerns about inconsistencies between populations and unknown functional impacts also apply here (79). Reasoning that a VDR and T1D association might only be evident if 25-OHD is sufficient to support calcitriol synthesis in cells relevant to T1D, investigators searched for this association as a function of latitude (79). They found a VDRF and T1D association that varied in strength according to ambient winter UV light. These data emphasize the importance of analyzing genetic data in the context of environmental variables. GWAS did not detect a VDR and T1D association (62, 80).

Intriguing data suggest an epistatic interaction between VDR alleles and HLA-DR susceptibility loci in T1D as in MS. The VDRF–T1D association was only evident in patients who carried the high-risk HLA-DRB1*0301 allele (81). The HLA-DRB1*0301 allele like the HLA-DRB1*1501 allele harbors a putative VDRE in its promoter (70). Alleles without the putative VDRE were associated with disease resistance. These parallels suggest an influence of UV light and vitamin D on HLA-DRB1 gene expression and presentation to CD4+ T lymphocytes of peptides relevant to T1D and MS etiology. The nature of the peptides and the timing and outcome of the presentation event are unknown, but could relate to thymic tolerance or peripheral T-cell responses to peptides from infectious agents. In any case, the positive findings regarding VDR polymorphisms provide genetic support for calcitriol and vitamin D receptor (VDR)-regulated transcriptional events as determinants of MS and T1D risk.

Additional evidence for vitamin D and calcitriol as sunlight’s signal transducers derives from vitamin D studies. An early study closely correlated childhood dental disease, serving as an accessible biomarker of exposure to low vitamin D status (82), with worldwide MS mortality (r = 0.78, p < 0.002) (83). The first vitamin D3 interventional study to inhibit autoimmune disease was performed in the murine experimental autoimmune encephalomyelitis (EAE) model of MS (84). Human case–control studies have correlated low vitamin D intake and low circulating 25-OHD with high MS risk (8587). Importantly, MS risk correlated inversely with circulating 25-OHD independently of personal UV light exposure (88). Abundant data have now correlated circulating 25-OHD inversely with MS disease activity (30, 8894). These data and the CYP27B1 and VDR evidence contradict the view that UV light’s protective effects in demyelinating disease do not involve vitamin D (95).

In MS patients who had low vitamin D3 levels and were not taking disease-modifying drugs, supplementary vitamin D3 as a stand-alone intervention significantly reduced disease progression (96), and decreased new lesion formation and progression from optic neuritis to clinically definite MS (97). A vitamin D2 supplementation study did not report similar findings (98), but significant methodological flaws were noted in that study (99). Moreover, it is well known that vitamin D2 and vitamin D3 are not biologically comparable as regards their metabolism and ability to transcriptionally activate the VDR (100102). Ongoing clinical studies are testing whether vitamin D3 as an add-on to disease-modifying drug therapy will improve drug efficacy (103). While this is a valid pharmaceutical question, it is separate from the question of whether vitamin D3 could be an etiology-based intervention to reduce the impact of autoimmune disease. Studies of vitamin D3 as a stand-alone intervention in individuals who have low vitamin D3 levels and who are not using disease-modifying drugs will be needed to address the etiological question. Such studies are needed and fully justified.

The T1D data also support vitamin D as sunlight’s protective signal transducer (6, 104, 105). The EURODIAB study correlated vitamin D3 supplementation in infancy with decreased T1D risk (odds ratio 0.67) (106), a finding that has been replicated (107). Finland, at 60–70°N, has the highest T1D incidence in the world (108). Finnish T1D incidence quadrupled as recommended vitamin D3 intakes for children decreased from 4000 to 400 IU/day between 1965 and 2005 (109). Another retrospective analysis found childhood vitamin D3 supplementation correlated with an 88% lower risk of T1D (110). These data imply that recommending higher vitamin D3 intakes for pregnant women, infants, and children might decrease T1D incidence by >75%. A very recent study demonstrated that life-long, high-dose vitamin D3 supplementation significantly reduced spontaneous diabetes in non-obese diabetic (NOD) mice (111). Finally, vitamin D3 interventions have improved glycemic control in T1D patients (112114).

The association of UV irradiance and vitamin D with autoimmune disease risk is strong and consistent, shows a dose–response relationship, is temporally plausible and appears to be universal with respect to genotypes, dietary and smoking habits, and exposure to infectious and commensal organisms. Thus, it is reasonable to suggest that vitamin D is probably the environmental factor with the greatest influence on the emergence of an autoimmune disease phenotype given a disease-susceptible genotype. What remains to be done to satisfy the Bradford Hill criteria (115) is to rigorously test this relationship experimentally in humans, and to uncover plausible biological mechanisms that cohere with known facts of autoimmune disease. There is a growing consensus that a vitamin D3 interventional study for autoimmune disease prevention is needed and fully justified.

T Lymphocytes in Autoimmune Disease

The effort to decipher protective vitamin D mechanisms in autoimmunity focuses on T lymphocytes, because in failing to correctly discriminate between self and non-self, auto-reactive T cells drive target organ destruction (116). In animal models, auto-reactive T-cell transfer drives target organ destruction (117). The role of pathogenic auto-reactive T cells as autoimmune disease initiators unites these diseases under a common mechanistic umbrella.

Unequivocal evidence of T-lymphocyte involvement in human T1D came from transplantation studies. Transplanting pancreatic tissue from a healthy subject into his/her T1D-affected identical twin failed as a T1D therapy due to a T-cell-mediated attack on the transplanted pancreatic tissue unless T-cell immunosuppressive therapy was administered (118). Evidence for pathogenic CD4+ T lymphocytes in T1D is now very strong (10, 119).

Original evidence of T-lymphocyte involvement in MS pathogenesis came from studies showing T cells rapidly migrated from the periphery into the CNS of MS patients, where clonally restricted, activated T cells accumulated in nascent MS lesions (120). The T cells from individual MS patients showed dominant usage of specific T-cell receptor (TCR) alpha and beta chains and VDJ sequences, and specificity for myelin basic protein (MBP) peptides presented by HLA class II molecules (121). These observations argue strongly for involvement of HLA class II-restricted, auto-reactive CD4+ T cells in MS pathogenesis (122).

Newer experimental approaches have confirmed the pioneering studies. Analysis of the mRNA transcriptome in blood cells from MS patients pointed to antigen presentation, the immune synapse, and T-cell deregulation in MS pathogenesis (123, 124). GWAS identified genes related to CD4+ T-cell function in regions harboring putative susceptibility loci (61). Analysis of expression quantitative trait loci in autoimmune disease states (125) and genetic analysis of inter-individual variability in T helper (Th) activation as a function of ancestry and autoimmune disease susceptibility (126) also implicated CD4+ Th cell activation in autoimmune disease pathogenesis. In summary, consideration of gene–environment interactions, sunlight and vitamin D, and T lymphocytes as autoimmune disease initiators provides the rationale for investigating vitamin D mechanisms in CD4+ T cells in an effort to understand the molecular etiology of autoimmune disease.

Vitamin D Metabolism from a T-Lymphocyte Perspective

Calcitriol’s Non-Calcemic Actions

The vitamin D system is an evolutionarily ancient and versatile system that coordinates a plethora of biological processes like cellular metabolism and growth, differentiation and death, organismal growth, reproduction, and immunity according to sunlight’s cues. The system’s signal transducing molecules are calcitriol, a small lipophilic hormone, and the VDR, a hormone-responsive transcriptional regulator (127). Calcitriol and mammalian VDR orthologs have functioned as sunlight sensors throughout >750 million years of evolution (128, 129). Many of the ancient organisms lacked calcified structures, so the vitamin D system must have originally supported non-calcemic functions (129). We focus on calcitriol’s actions in CD4+ T lymphocytes, but actions in other tissues, for example promotion of remyelination in the brain (130), insulin release from the pancreas (131), and intestinal barrier function in the colon (132), undoubtedly contribute to calcitriol’s potency as a chronic autoimmune and neurodegenerative disease inhibitor.

Enzymes of Vitamin D Activation

Vitamin D metabolism has been well described elsewhere (127, 133136). Cutaneous vitamin D3 generated by high energy UVB photons (290–315 nm) provides 90% of the human vitamin D requirement (108). Vitamin D3 synthesis varies seasonally at high latitudes, reaching a nadir 2 months after the winter solstice, and a zenith 2 months after the summer solstice (137). The biological half-life of 25-OHD3 is ~2 months, so this metabolite effectively integrates sunlight’s energy signal over time.

At High Latitudes, There is a Period of “Light Starvation” When Vitamin D Synthesis is Negligible

The higher the latitude, the greater is the period of light starvation. At 42°N, cutaneous vitamin D synthesis is negligible from November through February (138, 139). Vitamin D synthesis decreases with increasing skin pigmentation, sunscreen use, and advancing age. Vitamin D deficiency has rapidly become a worldwide health problem, due to life-style changes (indoor living and working; sunscreen use) that have reduced sunlight exposure (140, 141). Now, more than 75% of Caucasians and >90% of Blacks, Hispanics, and Asians in the USA have 25-OHD3 <75 nmol/L (30 ng/mL), double the number one decade ago (142). The rapid rise in vitamin D insufficiency correlates temporally with the rapid rise in MS and T1D incidence.

Two hydroxylation reactions produce the calcitriol from vitamin D3. The CYP2R1-encoded 25-hydroxylase converts vitamin D3 into 25-OHD3. The CYP27B1-encoded 1α-hydroxylase converts 25-OHD3 into calcitriol. Calcitriol induces CYP24A1 gene encoding the 24-hydroxylase to convert calcitriol into inactive calcitroic acid. These hydroxylases are cytochrome P450 enzymes (133, 134). They are encoded in nuclear DNA, but the enzymes themselves localize to the mitochondrial membrane.

Many chronic diseases provisionally associated with vitamin D disproportionately impact individuals of African ancestry (AA). Lower serum 25-OHD in people of AA may be contributing to these health disparities (143). Serum 25-OHD levels are subject to genetic regulation (74), and polymorphisms in the GC, DHCR7, and CYP2R1 genes have been correlated with circulating 25-OHD3 levels in people of European ancestry (EA) (144, 145). Motivated by the need to understand racial health disparities, recent research has investigated potential genetic contributions to circulating 25-OHD3 in AA and EA subjects. The AA subjects had different GC alleles encoding the vitamin D-binding protein (DBP) and lower circulating 25-OHD (146). In addition to binding vitamin D metabolites, the DBP functions in fatty acid transport, macrophage activation, and chemotaxis (147). A study of older male subjects from urban areas confirmed the association of CYP2R1 variants with serum 25-OHD after correction for vitamin D intake, season of sampling, BMI, and other variables (148). This association was more robust in EA than AA subjects. The GC variants were associated with serum 25-OHD only in the EA subjects. Age, ancestry, vitamin D intake, and season of sampling explained 19 and 24% of the variance in AA and EA subjects, respectively. Adding genetic variants to the model explained an additional 1% (AA) and 4% (EA) of the variance. However, 72% (EA) and 80% (AA) of the variance was unexplained. Thus, genetic variants are very minor contributors to racial disparities in serum 25-OHD. The data also suggest some relevant parameters were either not considered or were subject to a large error in measurement. Unlike other reports, skin pigmentation was unrelated to serum 25-OHD in this study, and UVR exposure was marginally related to serum 25-OHD only in EA subjects. These data indicate serum 25-OHD derived primarily from vitamin D ingestion, which was assessed by a questionnaire. The questionnaire may have introduced error. Whether it assessed vitamin D supplement use was not stated. The authors previously reported that their EA subjects were more likely to use vitamin D supplements than the AA subjects (149). To generate insight into disparities in chronic diseases provisionally associated with vitamin D, it will be essential to discover dominant variables by integrating data on age, gender, life-style choices, skin color, UVR exposure, diet, vitamin D supplement use, and genetic variation with data on 25-OHD and disease phenotype.

Paracrine Calcitriol Signaling to T Cells

An important question from a CD4+ T-lymphocyte perspective is whether the calcitriol signal derives from the kidney (endocrine signaling) or from tissue-resident cells at sites of inflammation (paracrine signaling). Calcitriol has been classically viewed as an endocrine hormone, with renal synthesis and systemic distribution regulating mineral ion homeostasis and skeletal maintenance in target tissues. Blood calcitriol levels are maintained within very narrow limits and show minimal seasonal fluctuation (150), whereas many autoimmune diseases show fluctuating periods of relapse and remission. For example, MS disease activity fluctuates seasonally, correlating with and lagging changes in ambient UVB sunlight (31) and 25-OHD3 (29). These data are not consistent with endocrine signaling and instead support paracrine signaling to T lymphocytes.

Evidence that tissue-resident antigen-presenting cells (APC) produce calcitriol (151) and T cells express the VDR (152, 153) first suggested paracrine signaling between immune cells. Adding 25-OHD3 to cultures altered human T-cell responses only when APC were present to produce calcitriol (154). The APC required activation through pathogen-associated pattern recognition receptors, for example the toll-like receptors (TLR), and/or stimulation by cytokines, most notably interferon-gamma (IFN-γ), interleukin-2 (IL-2), and IL-15, to become calcitriol producers (136, 154157).

Calcitriol synthesis has now been shown in many non-calcified mammalian tissues frequented by roving T lymphocytes, most notably the skin (158, 159), lung (160), colon (161, 162), brain (163), placenta, and other reproductive tissues (154, 157, 164), so paracrine signaling to T cells is a well-established concept. The CYP27B1 transcripts are more abundant in tissues with barrier (skin, lung, colon) or reproductive function (maternal decidua, fetal trophoblast, testis) than in the kidney (158), supporting the thesis that biological protection at host–environment interfaces and environmental impacts on reproduction may have driven vitamin D system evolution. Available data suggest that paracrine calcitriol signaling to T cells within tissues is likely the major pathway by which sunlight exerts its influence on the emergence of an autoimmune disease phenotype.

It is valuable to consider how calcitriol signaling to CD4+ T lymphocytes could be compromised causing immune-mediated tissue damage. Research in the EAE model of MS demonstrated that insufficient vitamin D3 disrupted calcitriol synthesis in the CNS (84). Research in the NOD model of T1D demonstrated that reduced 1α-hydroxylase activity in APC disrupted signaling and contributed to diabetes (165). As detailed above, CYP27B1 gene lesions compromise paracrine signaling. It is not widely appreciated that corticosteroids like prednisone (166), prednisolone (167), or dexamethasone (151) compromise paracrine signaling to CD4+ T cells because they inhibit calcitriol synthesis by activated innate immune cells in vivo. Corticosteroid inhibition of calcitriol synthesis in the airway may have contributed to negative results in the VIDA trial of vitamin D3 supplementation in asthma patients who were receiving concurrent corticosteroid therapy (168).

Deregulation of the CYP24A1 gene promoter might also disrupt calcitriol signaling and promote a pro-inflammatory state. Epigenetic silencing of the CYP24A1 promoter in placental tissue promoted calcitriol accumulation and an immune tolerant state at the maternal–fetal interface (169, 170). Expression of the CYP24A1 gene differed between males and females. In the CNS of rodents with EAE, Cyp24a1 gene expression was higher and calcitriol responsiveness lower in males than females (84, 171). Similarly in human T cells from MS patients and healthy controls, CYP24A1 gene expression was higher and calcitriol responsiveness was lower in males than females (172). The rodent and human data suggest males may produce calcitriol at a rate equal to females but inactivate it faster. Gender differences in CYP24A1 expression and vitamin D metabolism may underlie the stronger inverse correlation between ambient UV light and MS risk in women compared to men (173). Estradiol addition to male T cells decreased the CYP24A1 transcripts and increased calcitriol responsiveness (172), suggesting estradiol may silence the CYP24A1 gene to promote calcitriol accumulation (169, 170).

Vitamin D Receptor Expression by CD4+ T Lymphocytes

Vitamin D Receptor and Vitamin D-Responsive Elements

The VDR enables cells to respond to calcitriol. Early researchers demonstrated a protein with high affinity (Kd 0.1 nM) for calcitriol in activated human CD4+ T lymphocytes that was later identified as the VDR (152, 153). The VDR is a nuclear protein that dimerizes with the retinoid X receptor to regulate gene expression through VDRE in calcitriol-responsive genes (127). A VDRE is composed of two hexameric half-sites, arranged as direct repeats separated by three random base pairs, for example GGTTCACGAGGTTCA (174, 175). Depending on the type of cell, the ligand-activated VDR–RXR complex recruits either coactivator complexes and cooperating transcriptional machinery or corepressor complexes to determine the nature of the transcriptional response from VDRE-containing target genes.

Antibody specificity problems have confounded investigations of VDR protein expression in T cells. All but one of the commercially available antibodies to the VDR bound non-specifically to cells and tissues from DeMay VDR-null mice (176, 177). These were confirmed by flow cytometry; all available antibodies either stained CD4+ T cells from DeMay VDR-null mice or did not yield a signal (171). The specificity problems have slowed progress in assessing VDR protein expression in T-cell subsets.

VDR and CD4+ T-Cell Activation

Resting and activated T cells differ significantly in VDR expression; resting cells expressed fewer than 1000 VDR/cell and activation increased this number 10-fold (178). VDR expression peaked at 48 h post stimulation (179). In rodents, activated CD8+ T cells had higher VDR expression than CD4+ T cells (180).

Vitamin D receptor expression was needed for optimal human T-cell activation in vitro (181). Naïve T cells were VDR negative and responded weakly to TCR stimulation. Weak TCR signaling via the mitogen-activated protein kinase p38 pathway induced VDR expression. The T cells subsequently up-regulated phospholipase C-gamma1 (PLC-γ1) expression 75-fold, enabling them to flux calcium and become fully activated. PLC-γ1 induction appeared to be VDR-dependent. The dependence of T-cell activation on VDR expression has not been demonstrated in vivo.

VDR and CD4+ T-Cell Subsets

Among rodent CD4+ T-cell subsets, IL-4-producing Th2 cells, IFN-γ-producing Th1 cells, and IL-17-producing Th17 cells all had abundant Vdr transcripts (182185). Importantly, a source of estradiol in female mice (186), and a functional Ifng gene in male and female mice (171) were essential for Th1 and Th17 cell Vdr gene expression. Recent data on human Th1 and Th17 cells produced in vitro also show high Vdr gene expression (179).

Whether CD4+Foxp3+ Treg cells express the VDR is unclear. Very low Vdr transcript levels were observed in rodent CD4+Foxp3+ Treg cells produced in vitro (187) and rodent CD4+EGFP+ Treg cells generated in vivo and flow-sorted from the spleens of Foxp3EGFP reporter mice (184). Those in vivo studies revealed an inverse relationship between Foxp3 and Vdr gene expression, with many Foxp3 transcripts and few Vdr transcripts in the Foxp3+EGFP+ T cells and the reverse pattern in the EGFP T cells. Consistent with low Vdr gene expression, calcitriol had no impact on flow-sorted CD4+Foxp3+EGFP+ T cells during activation in vitro (184). Calcitriol also had no impact on CD4+Foxp3+ T-cell proportions in the spleens of WT B6. CreVDRfl/fl mice and B6. Cre+VDRfl/fl mice with CD4+ T-cell-specific Vdr targeting, whether or not they were treated with calcitriol (184). These data suggest fully differentiated, mature rodent CD4+Foxp3+ T cells may not be calcitriol responsive. Contrary to this view, VDR-dependent down-regulation (187) and up-regulation (185) of Foxp3 transcription in rodent CD4+Foxp3+ T cells have both been reported. It will be interesting to learn the molecular details of Vdr and Foxp3 gene expression control as research in this area progresses.

CD4+ T Cells are Calcitriol Targets in Autoimmune Disease

Selective Vdr gene inactivation experiments in rodents provided unequivocal evidence that calcitriol targets CD4+ T lymphocytes for the purpose of immune system regulation (184). Reciprocal mixed bone marrow chimera studies established that hematopoietic cell Vdr gene expression was necessary for calcitriol to inhibit EAE induction. In fact, chimeric mice lacking the VDR in hematopoietic cells had a particularly aggressive EAE disease course. Subsequently, mice with CD4+ T-cell-specific Vdr targeting were constructed. Evaluation of T-cell subsets in the periphery of naïve T-cell Vdr-targeted mice assured that this genetic manipulation did not influence T-cell proportions during thymic selection. Neither vitamin D3 nor calcitriol inhibited EAE induction in mice with CD4+ T-cell-specific Vdr targeting (184). Thus, EAE data establish that calcitriol exerts protective biological effects against autoimmunity in vivo through the nuclear VDR in CD4+ T lymphocytes. These data do not rule out vitamin D effects on myeloid cells for a pathogen protective immune response (188). Studies in mice with CD4+ T-cell-specific Vdr targeting (184), and in humans with loss-of-function mutations in the VDR gene (189) support the view that the outcome of calcitriol signaling within CD4+T cells is likely a major determinant of sunlight’s influence on the emergence of an autoimmune disease phenotype.

Vitamin D and Thymocyte Selection

Thymic Negative Selection in Autoimmune Disease

The TCR repertoire is shaped in the mammalian postnatal thymus, and autoimmune disease risk is believed to reflect in part defects in this process (190). When immature CD4+CD8+ thymocytes engage cortical thymic epithelial cells expressing self peptides embedded in major histocompatibility complex (MHC) molecules, those that bind self MHC molecules with adequate affinity undergo positive selection and develop into CD4+ or CD8+ T cells. Those that bind with inadequate affinity for survival signaling undergo apoptosis. Surviving thymocytes migrate to the thymic medulla where they engage medullary thymic epithelial cells (mTEC) presenting self peptides, including peptides derived from tissue-restricted proteins. Cells that bind self peptides with high affinity receive apoptotic signals. The survivors are released to form the pre-immune TCR repertoire.

An analysis of human autoimmune TCR–peptide–MHC complexes revealed structural anomalies compared to anti-microbial TCR–peptide–MHC complexes suggesting these TCR may have escaped thymic negative selection (191). For example, a TCR from an MS patient recognized both a DRB1*1501-restricted MBP peptide and a DRB5*0101-restricted EBV peptide, suggesting this TCR may have escaped negative selection become activated by a pathogen (192). A negative selection failure during early life is believed to release potentially pathogenic T cells that initiate autoimmune disease when a cross-reactive pathogen peptide activates them (116).

Vitamin D and Thymic Negative Selection

Sunlight and vitamin D3 exert their strongest influence on MS risk during childhood, which coincides with the peak period for T-cell selection in the postnatal thymus (19). Intriguingly, recent data quantifying signal joint TCR excision circles (sjTRECs) in human cord blood T cells as a function of season demonstrated seasonal variability in thymic output (193). May-born infants had significantly lower circulating 25-OHD3 and higher sjTRECs/105 T cells than November-born infants. The sjTRECs are extra-chromosomal DNA circles formed during TCR gene rearrangement. They are not replicated as mature peripheral T cells divide, so they serve as markers of recent thymic emigrants in the peripheral T-cell repertoire. The data suggest a positive influence of 25-OHD3 on thymic negative selection. Rigorous investigation of this possibility is warranted because of the significance such an influence would have on autoimmune disease risk acquisition, and the insight it would provide into the timing of risk acquisition and therefore the timing of etiology-based intervention strategies to prevent disease.

Vitamin D and CD4+ Th1 and Th17 T Cells

Effector T-Cell Anti-Microbial Response, Cross-Reactivity, and Autoimmune Disease

Exposure to common infectious or commensal organisms triggers activation of those CD4+ T cells capable of recognizing foreign antigens in an MHCII context. The biological imperative to provide adequate immune cover for the host inevitably results in some effector T-cell cross-reactivity between foreign and self peptides due to molecular mimicry (116). Peptide antigen diversity is orders of magnitude larger than TCR diversity (194), and TCR–peptide–MHC binding shows conformational plasticity, relatively low affinity, and rapid off-kinetics (191, 195). The immune system has evolved mechanisms to prevent overly aggressive T-cell-mediated responses that cross-react with host peptides from damaging host tissue (196). These mechanisms include inhibition of pro-inflammatory cytokine synthesis, and Treg cell-mediated restriction of T-cell expansion through inhibitory receptors like cytotoxic T-lymphocyte antigen 4 (CTLA4), induction of T-cell anergy, and cell death signaling (197). Intriguing new research has documented a central role for the vitamin D system in the development of Treg cells and the termination of effector T-cell responses.

Modeling Human Autoimmune Disease

Two animal models have been particularly useful in developing mechanistic knowledge of direct calcitriol actions in CD4+ T cells, murine EAE (198, 199), and the NOD mouse (200). These animal models show strong parallels to their respective human diseases, especially in the immunological aspects of disease. The parallels are imperfect, for example regarding the role of human microbial exposures, because our understanding of human autoimmunity is incomplete. Nevertheless, hypothesis testing is more facile in animal models where direct manipulation of contributing factors is possible. Calcitriol inhibited autoimmune disease in the EAE and NOD models (201203). Working back and forth between animal models and human disease has allowed rapid forward progress to be made in understanding how the vitamin D system modulates immunity.

Vitamin D and Th1 Cells

Disease prevention studies in animal models have suggested that calcitriol directly inhibits encephalitogenic Th1 cells. Calcitriol inhibited EAE induction (182, 183, 185, 202205), and targeting the Vdr gene specifically in CD4+ T cells abrogated this inhibition (184). However, calcitriol had no effect on CD4+ T-cell priming in the periphery, myelin-specific T-cell trafficking into the CNS, or IFN-γ synthesis by freshly explanted, purified CD4+ T cells from EAE mice (182). Instead, the CNS-infiltrating myelin-specific CD4+ T cells displayed an anergic phenotype in the calcitriol-treated mice. These findings have been confirmed (171, 184). In another strain, calcitriol inhibited EAE induction, but the mechanism reported was a decreased Th1 cell frequency in peripheral lymphoid tissues (206).

The mechanisms involved in amelioration of established autoimmune disease may differ from those involved in autoimmune disease prevention. Administering calcitriol to rodents with established EAE resulted in very rapid disease remission that correlated with fewer CNS CD4+ T cells, loss of IFN-γ-production, and a significant decrease in CNS pathology (207210). Induction of Treg cells in the EAE studies is discussed below.

A short course of oral calcitriol treatment in healthy male volunteers had no measurable effect on circulating IFN-γ (211). Likewise, in two studies carried out in MS patients, vitamin D3 supplementation had no measurable effect on circulating IFN-γ (96, 212). In a third study performed in healthy volunteers, vitamin D3 supplementation decreased the mean percentage of Th1 cells in the circulation from 20 to 17% (213).

Vitamin D and Th17 Cells

Disease prevention studies in animal models have also suggested that calcitriol directly inhibits encephalitogenic Th17 cells. Experimental autoimmune uveitis (EAU) induced by immunization of mice with a retinal antigen serves as model for human autoimmune uveitis. Oral calcitriol prevented as well as partly reversed EAU disease (214). In this model, calcitriol treatment in vivo of mice impaired T-cell commitment to the Th17 lineage as well as Th17 production of IL-17.

Similar results were reported in the EAE model. The spleens of calcitriol-treated mice had fewer splenic Th17 cells and lower IL-17 production than the placebo controls (183, 187). Mice with global inactivation of the Vdr gene had activated Th17 cells that overproduced the cytokine (215). Calcitriol treatment reduced Th17 cells in the CNS in an EAE prevention study (171) and in two EAE treatment studies (185, 216). Different mechanisms were suggested to explain these findings (Figure 1A). In one study, calcitriol did not suppress Il17 gene transcription, but inhibited Th17 cell IL-17F production in a VDR-dependent manner by a post-transcriptional mechanism (183). In another study, calcitriol suppressed Il17 gene transcription by blocking nuclear factor for activated T cells (NFAT), recruiting histone deacetylase, and sequestering Runt-related transcription factor 1 (Runx1) (185). In the third study, elimination of Th17 cells by a programed cell death mechanism was suggested (216). The reconciliation of these divergent mechanisms awaits further investigation.

FIGURE 1
www.frontiersin.org

Figure 1. Proposed calcitriol mechanisms in effector Th1, Th17, and FoxP3+ Treg cells. (A) Calcitriol and IL-17 synthesis in Th17 cells. Calcitriol inhibited IL-17 synthesis in a VDR-dependent manner by a post-transcriptional mechanism (183, 187). It also blocked NFAT, recruited histone deacetylase, and sequestered Runx1 to suppress murine Il17 gene transcription (185). (B) Calcitriol and effector CD4+ T-cell apoptosis. Calcitriol increased CD4+ T-cell sensitivity to extrinsic cell death signals that may have been transduced through the FAS ligand–FAS–caspase-8 and/or the galectin-9–TIM3–calpain pathways (208210, 216). (C) Calcitriol and FOXP3 gene transcription. Calcitriol increased the transcription of the murine FoxP3 (185) and human FOXP3 genes (217). Three VDREs that functioned as calcitriol-dependent enhancers were identified in a conserved non-coding sequence (abbreviated CNS in this diagram) between exons 1 and 2. Calcitriol also increased CTLA4 protein expression (218220).

Studies of calcitriol and Th17 cell activity in humans are at an early stage. Cell culture studies have shown that adding calcitriol to activated CD4+ T cells reduced the frequency of Th17 cells (218, 221), an effect that was more pronounced in cultures of T cells from women than from men (172). New data have extended these observations in a vitamin D3 dose escalation study performed in healthy controls during the UV-restricted winter months (213). As supplementary vitamin D3 increased from 2000 to 8000 IU/day, the mean serum 25-OHD rose from 30 ± 12 to 159 ± 29 nmol/L, and the proportions of Th1 and Th17 cells in the peripheral blood decreased. There appeared to be a threshold effect with at least 70 nmol/L of 25-OHD triggering a mean 40% decrease in circulating Th17 cells in the healthy controls. Other studies found no correlation between Th17 cells and 25-OHD in MS patients (222), and no effect of supplementary vitamin D3 on Th17 cells in MS patients (223). Why the human data are conflicting is not clear, but vitamin D3 status at enrollment, vitamin D3 dose, dose frequency (135), use of disease-modifying drugs, and timing of sampling relative to estrogen cycling in women are potential confounding factors. Nevertheless, given the strong animal modeling data and some human in vivo data that are consistent with animal modeling, it is reasonable to suggest that vitamin D may significantly influence the emergence of an autoimmune disease phenotype by dampening pathogenic Th17 cells and IL-17 synthesis.

Vitamin D and Th2 Cells

Analyses of Th2 cells in vivo have also yielded inconsistent results. Administering calcitriol to mice before EAE induction increased the IL-4 transcripts in the lymph nodes and in the CNS compared to the placebo controls (204). Also, targeted disruption of the IL-4 gene moderately decreased the protective function of calcitriol in EAE (224). A subsequent report found no significant differences between calcitriol-pretreated and placebo-pretreated B10.PL mice with regard to IL-4 mRNA in the lymph nodes or the CNS after immunization with MBP (182). There was also no effect of calcitriol on the IL-4 protein synthesis per Th2 cell. Another report found administering calcitriol to Biozzi AB/H mice before EAE induction had no effect on the IL-4-producing Th2 cell frequency (206). Thus, there are some inconsistencies regarding IL-4 that remain to be resolved. Since calcitriol inhibition of EAE decreased slightly in IL-4-null mice, there may be some IL-4 contribution to the mechanism (224), but it is possible that Th2 cells were not the source of the protective IL-4.

The IL-5-producing Th2 cells have a pathogenic role in human asthma. When CD4+ T cells from asthma patients were stimulated with dust mite allergen in the presence of calcitriol, the hormone decreased IL-5, IL-9, and IL-13 production (225). Calcitriol’s actions to promote IL-10-producing Treg cells in asthma patients are discussed below.

Vitamin D and CD4+ T-Cell Programed Cell Death

Effector T-Cell Apoptosis and Autoimmune-Mediated Tissue Damage

When effector CD4+ T cells are no longer needed for anti-microbial defense, programed cell death mechanisms remove them to limit immune-mediated tissue damage. Several mechanisms used by CD4+ Treg cells (197, 226), astrocytes and neurons (227, 228), and a few other types of cells to control effector CD4+ T cells involve apoptosis induction. In individuals predisposed to autoimmunity, it appears that effector CD4+ T cells resist cell death mechanisms, proliferate, and continue to produce inflammatory molecules that damage the host’s tissues.

The effector T lymphocytes from the T1D-susceptible NOD mice were resistant to cell death signals. The resistance to cell death mapped to a genomic region encompassing the Ctla4 gene, which was defectively expressed on T cells from these mice (229). In human autoimmunity, myelin-specific CD4+ T cells from MS patients provide an example of defective cell death mechanisms (230). These T cells were defective in the FAS–caspase 8 cell death pathway (231) and the galectin-9–TIM-3 cell death pathway (232). It was intriguing that myelin-specific T cells from MS patients with benign disease had a significantly augmented galectin-9–TIM-3 cell death pathway compared to MS patients with active disease (Saresella et al., ECTRIMS 2011, Abstract P324). Moreover, MS disease activity fluctuated seasonally, with a high frequency of new lesions following a period of reduced ambient UV light and vice versa (31). Taken together, the seasonal fluctuations in MS disease activity and disease activity-associated fluctuations in T-cell apoptosis resistance hint at a possible causal relationship between ambient UVB light, vitamin D3 supplies, myelin-specific T-cell susceptibility to cell death, and demyelinating disease activity.

Vitamin D Regulation of Effector T-Cell Apoptosis

A causal relationship between calcitriol, myelin-specific T-cell responsiveness to cell death signals, and demyelinating disease activity has been demonstrated in the animal model of MS (Figure 1B). In animals with EAE, calcitriol treatment in vivo increased the susceptibility of pathogenic myelin-specific CD4+ T cells to extrinsic, CNS-derived apoptotic signals (207). This treatment induced the pro-apoptotic gene encoding caspase-8-associated protein, which is essential for FAS-mediated apoptosis, and repressed cellular inhibitor of apoptosis protein 2 (cIAP-2), an apoptosis inhibitor (209). Effector T-cell death is triggered in the CNS by astrocyte- and neuron-mediated activation of the T-cell FAS death pathway (227, 228, 233235). In the calcitriol treatment study, apoptotic CD4+ T cells were evident in CNS lesions and IFN-γ production ceased by 12 h post treatment; within 1 day, the CD4+ T-cell numbers were 60% reduced, correlating with abatement of clinical disease signs by day 3 (208, 210). These T-cell changes were not observed in the periphery or in vitro. A transient increase in Helios+FoxP3+ Treg cells coincided with the loss of effector T cells suggesting the Treg cells may have played an active role in effector T-cell elimination (216). In the NOD model of T1D, peripheral T cells were resistant to programed cell death signals, and calcitriol treatment restored their sensitivity to these signals (229, 236238). These results support a direct calcitriol action on effector T cells to promote sensitivity to apoptotic signals.

Vitamin D may also influence other cell death pathways in effector CD4+ T cells. Treg cells expressing galectins can trigger effector T-cell death via the TIM3 pathway (226). The N-linked glycans, in particular terminal galactosyl residues, are essential for galectin-9 binding to TIM3 and induction of cell death through calpain and caspase-1 activation (239, 240). Importantly, calcitriol enhanced α-mannosidase N-acetyl-glucosaminyl transferase 1, the rate-limiting enzyme in N-linked glycan synthesis (241), UDP-galactose: beta N-acetyl glucosamine-beta-1,3-galactosyl transferase, and calpain (209). Thus, calcitriol may improve T-cell sensitivity to the galectin-9–TIM3–calpain cell death pathway, which was impaired in MS patients (232). Together, the animal modeling and human in vivo data linking vitamin D and calcitriol with effector T-cell apoptosis suggest that vitamin D may significantly influence the emergence of an autoimmune disease phenotype by increasing effector CD4+ T-cell sensitivity to extrinsic cell death signals.

Vitamin D and CD4+ T-Regulatory Lymphocytes

Treg Lymphocytes in Autoimmune Disease

CD4+ Treg cells are defined functionally by their ability to limit prolonged effector T-cell activation, thereby preventing autoimmune-mediated pathology (242). T lymphocytes that suppress autoimmunity were first demonstrated in mice (243246). Subsequently, their existence was confirmed in humans (247). Several types of Treg cells exist differing in origin, phenotype, and function. The most intensely studied are those for which FoxP3 serves as the definitive transcription factor in lineage specification (248). One FoxP3+CD4+ Treg cell subset arises during thymic development (tTreg), whereas another arises during TCR engagement in the periphery (pTreg) (249). These FoxP3+CD4+ Treg cell subsets are non-functional in male scurfy mice and boys with the multi-organ autoimmune disease immune-dysregulation polyendocrinopathy enteropathy X-linked syndrome (IPEX), due to loss-of-function mutations in the X-linked rodent Foxp3 and human FOXP3 genes, respectively. The FoxP3 protein serves as a biomarker for this Treg lineage.

The T-regulatory cell type 1 (Tr1) cells differ from FoxP3+CD4+ Treg cells in origin, phenotype, and function (250). The CD4+ Tr1 cells are peripheral memory T lymphocytes that are anergic and do not express FoxP3 (251). Selective biomarkers for Tr1 cells in humans and rodents are co-expression of integrin alpha 2 subunit (CD49b, a cell adhesion molecule) and lymphocyte activation gene 3 (LAG-3 or CD223, a negative regulator of effector T-cell function) (252). Upon CD3 stimulation and co-stimulation via the complement regulator CD46 in the presence of IL-2, the CD4+ Tr1 cells rapidly proliferate and secrete large amounts of IL-10 and TGF-β to exert suppressive function (253). The Tr1 cells also release granzyme B to specifically lyse APC of myeloid origin, effectively terminating further effector T-cell activation.

Functional defects in FoxP3+CD4+ Treg cells have been described in multiple autoimmune diseases (254). Such defects were first reported and confirmed in peripheral blood T cells from MS patients (255, 256). Additional evidence emerged for reduced CD4+ Treg cell generation in the thymus of MS patients (257, 258). The impairment correlated with reduced FoxP3 expression in MS patient T cells (259, 260). Studies of FoxP3+CD4+ Treg cell defects in MS have been reviewed (261, 262). Functional defects in FoxP3+CD4+ Treg cells have also been extensively described in T1D (263). Below, we review research suggesting calcitriol enhances the expression of Helios, a positive regulator of FoxP3, and of FoxP3 itself.

Functional defects in IL-10-producing CD4+ Tr1 cells have also been described in many autoimmune diseases (250). An early report demonstrated that T cells from T1D patients showed extreme polarization toward a pro-inflammatory Th1 phenotype, whereas the T cells from the non-diabetic, HLA-matched control subjects showed an extreme bias toward the IL-10-secreting Treg phenotype (264). Functional defects in Tr1 cells have also been extensively studied in MS. Stimulation of MS patient T cells in vitro via CD3 and CD46 elicited very little IL-10 production compared to controls (265). These findings have been confirmed (261, 266, 267) and extended to a primate model of MS (268). Functional defects in IL-10-producing CD4+ Tr1 cells have also been described in asthma and rheumatoid arthritis (269). Below, we review research suggesting calcitriol enhances the expression of IL-10 and controls the Th1–Tr1 switch through modulation of CD46, a positive regulator of IL-10.

Cutaneous UVB Light Exposure and Treg Lymphocytes

Cutaneous UVB light exposure and in situ vitamin D metabolism may have a particularly important role in cutaneous Treg cell development. Cutaneous UVB exposure promotes T-cell-mediated immunity to microbial pathogens; Finsen received the 1903 Nobel Prize in Medicine for phototherapy of skin diseases (270). More recently, the role of cutaneous UVB exposure in peripheral immune tolerance has been recognized (271). The dual role of UVB exposure illustrates how the vitamin D system may have evolved to promote anti-pathogen responses at environmental interfaces, and subsequently to terminate these responses before the tissues sustain immune-mediated damage.

Ground-breaking animal modeling studies revealed how cutaneous UVB exposure promotes peripheral immune tolerance at the molecular level (272). Infection increased keratinocyte expression of receptor-activator of NF-kappaB ligand (RANKL). In transgenic mice with keratinocytes overexpressing RANKL, the keratinocytes signaled receptor-activator of NF-kappaB (RANK)-expressing Langerhans cells to induce IL-10-producing pTreg cells in the skin-draining lymph nodes. These pTreg cells suppressed the contact hypersensitivity response to cutaneous antigens in the original host animal and in adoptive transfer host animals. Furthermore, when skin was grafted from WT mice or Tnfrsf11a-null mice lacking RANKL onto WT host animals, UVB irradiation of the WT but not the Tnfrsf11a-null skin graft prevented graft rejection. These experiments established that keratinocyte RANKL expression was necessary for UVB exposure to induce IL-10-producing pTreg cells specific for the grafted skin to maintain peripheral tolerance.

Importantly, cutaneous UVB exposure promoted pTreg cell development by a calcitriol and VDR-dependent mechanism (273). Keratinocytes have a complete vitamin D3 biosynthetic pathway which when stimulated by infection, injury, or light produced 2–5 nmol/L of calcitriol in situ (159, 274), about 20- to 50-fold higher than calcitriol in the blood plasma of non-pregnant women (275). Furthermore, calcitriol transcriptionally activated the murine (Tnfrsf11a) and human (TNFRSF11A) genes encoding RANKL through highly evolutionarily conserved and functionally active VDREs in the promoter regions, as described in osteoblasts (276278), murine and human T lymphocytes (279, 280), and keratinocytes (127). In the animal model of cutaneous hypersensitivity, topically applied calcitriol mimicked cutaneous UVB exposure by inducing keratinocyte RANKL expression and stimulating IL-10-producing pTreg cells by a mechanism that was VDR-dependent (281, 282). Cell labeling studies demonstrated that the pTregs trafficked between the skin, the skin-draining lymph nodes, and the circulation (283). Calcitriol suppressed T-cell expression of the gut-homing receptors and increased T-cell expression of CC chemokine receptor 10, enabling T-cell migration to the skin-specific chemokine CCL27 produced by keratinocytes (100). In summary, seminal studies of experimental UVB-induced peripheral tolerance in rodents established that a calcitriol- and VDR-dependent pathway exists involving (i) keratinocyte synthesis of calcitriol and stimulation of RANKL expression, (ii) signaling to RANK-expressing Langerhans cells, (iii) movement of the Langerhans cells to the skin-draining lymph nodes, (iv) induction of cutaneous antigen-specific, IL-10-producing pTreg cells, and finally (v) movement of the pTregs to the skin to prevent cutaneous antigen-specific effector T cells from degrading the epidermal barrier.

New research has begun to translate this knowledge to humans. Subjects from the North of Scotland who had skin disease were recruited between December and March for a phototherapy study (284). The subjects had a mean baseline 25-OHD level of 34 nmol/L and 0.5% FoxP3+ pTreg cells as a percentage of blood CD3+ T cells. After 4 weeks of phototherapy, the mean 25-OHD was 78 nmol/L and the FoxP3+ pTreg cell percentage was 1.6%. Another study demonstrated that UVB exposure substantially expanded pTreg cell numbers in the skin, and these skin pTreg cells appeared to have derived from peripheral immune organs (285). The studies in animals and humans implicating calcitriol as a positive regulator of the Tnfsf11a and Tnfrsf11a genes and cutaneous FoxP3+ pTreg cell induction suggest that during periods of light starvation, phototherapy may be a particularly efficient method of influencing the emergence of an autoimmune disease phenotype by increasing CD4+CD25+FoxP3+ Treg cell development and suppressive function.

Vitamin D and the IL-10–IL-10R Pathway

Interleukin-10 is a particularly important anti-inflammatory cytokine that protects the CNS (286), the airway (287), the gastrointestinal tract (288), and other tissues from immune-mediated pathology (289, 290). Disruption of the IL-10–IL-10R signaling pathway results in severe inflammatory disease (291).

Early research in two rodent models suggested that calcitriol promotes the development of IL-10-producing Treg cells. In the EAE model, suppressor T cells inhibited disease by an IL-10-dependent mechanism (292, 293). Calcitriol inhibited EAE not by blocking T-cell priming in the periphery or effector T-cell trafficking into the CNS, but instead by inducing Rag-1 gene-dependent regulatory lymphocytes (182). Subsequent work demonstrated that activation of human and mouse T cells in the presence of dexamethasone and calcitriol in vitro yielded IL-10-producing Tr1 cells (294). Furthermore, vitamin D3 or calcitriol-mediated inhibition of EAE required functional Il10 and Il10R genes, bidirectional IL-10–IL-10R signaling between hematopoietic and non-hematopoietic cells (205), and a functional Vdr gene in CD4+ T cells (184). Calcitriol had no effect on IL-10R expression (205). In the NOD model of T1D, calcitriol increased the frequency of Treg cells in the pancreatic lymph node (295), and global Vdr gene inactivation decreased the frequency of these cells (296). A positive calcitriol influence on T-cell IL-10 synthesis was observed in vitro (297). These animal studies suggested calcitriol may inhibit autoimmune disease at least in part through a VDR-dependent action on CD4+ T cells to induce IL-10-producing Treg cells.

There is limited information on a possible direct link between vitamin D and IL-10 in humans. An early study found a seasonal variation in the cord blood IL-10 level that correlated directly with the 25-OHD for infants born at 51°N (298). The 25-OHD levels were 99% higher and IL-10 was 43% higher in samples obtained in the summer compared to samples obtained in the winter. Investigations of calcitriol effects on IL-10 synthesis by human T cells activated in vitro have yielded conflicting results with reports of both enhancement (172, 221) and inhibition (179). A very interesting recent study examined the effect of adding 25-OHD3 to cultures of lymphocytes from healthy controls and from hereditary vitamin D-dependent rickets (HVDRR) patients with loss-of-function mutations in the VDR gene (189). The 25-OHD3 increased IL-4, IL-10, and IFN-γ production from control lymphocytes, but not patient lymphocytes, demonstrating calcitriol synthesis and VDR dependence of IL-10 induction in vitro.

Vitamin D and Treg Cells in Asthma

Asthma research has yielded important new insights regarding vitamin D and Treg cells (299). One of them relates to the balance between immunity to pathogens and self-tolerance. Calcitriol ingestion by healthy volunteers and calcitriol addition to cultures increased CD4+ T-cell IL-10 secretion and TLR9 expression (300). Adding TLR9 agonists to the cultures decreased T-cell IL-10 production and suppressive function. These findings are significant because they suggest pathogen signaling through TLR9 could suspend T-cell suppressive function temporarily until an infection has been cleared. A second insight relates to calcitriol enhancement of T-cell surface molecules that dampen immune responses. In this case, calcitriol up-regulated CD200 on human peripheral and respiratory tract CD4+ T cells in vitro, and there was a trend toward up-regulation in vivo in healthy, but not asthmatic individuals (301). CD200 is an immunoglobulin superfamily member; it imparts a unidirectional negative signal to suppress innate and adaptive immune responses.

Yet another insight from the asthma research relates to the distinction between CD4+FoxP3IL-10+ Tr1 cells and CD4+FoxP3+IL-10 Treg cells. Systemic vitamin D status correlated directly with airway levels of IL-10 and CD4+FoxP3+ T cells in pediatric asthma patients and in healthy controls (302, 303). In cultures of human peripheral blood T cells, adding calcitriol at moderate levels (10−8 mol/L) increased the frequency of CD4+IL-10+ Tr1 cells, whereas higher calcitriol (10−6 mol/L) increased the frequency of CD4+FoxP3+ T cells. However, there was little co-expression of FoxP3 and IL-10, and IL-10 impaired calcitriol-mediated enhancement of FoxP3. Modulating the cytokine environment to include TGF-β together with calcitriol favored CD4+FoxP3+ T-cell outgrowth (304). The CD4+IL-10+ and CD4+FoxP3+ T-cell populations had equivalent suppressive activity in vitro, although their suppressive mechanisms were IL-10-dependent and IL-10-independent, respectively.

These data are highly significant for two reasons. First, they demonstrate that the vitamin D system supports two phenotypically and functionally distinct Treg cell populations. Secondly, they suggest that as infections are resolved, calcitriol-supported development of CD4+FoxP3IL-10+ Tr1 cells and CD4+FoxP3+IL-10Treg cells may proceed in an ordered sequence to perform slightly different functions, for example attenuating innate and adaptive immune responses by IL-10-dependent and IL-10-independent mechanisms, respectively. Many interesting questions remain regarding the underlying mechanisms of Treg cell immune response regulation in the airway, and the generality of these findings to other interfaces between the environment and the host organism where immunity to pathogens must be balanced with self-tolerance.

Vitamin D and FoxP3+ Treg Lymphocytes

Above, we summarized early animal modeling data suggesting calcitriol promoted the development of IL-10-producing Treg cells before FoxP3 expression was in use as a Treg biomarker. New animal modeling data suggest calcitriol may be a positive regulator of the murine FoxP3 gene itself (185), and of the Ikzf2 gene encoding Helios (216), a transcription factor that binds to the FoxP3 promoter and stimulates its transcription (305307).

CD4+Foxp3+ Treg cell studies in humans are scarce and conflicting. Two reports correlated 25-OHD with circulating CD4+Foxp3+ Treg cell percentages (308, 309). One study correlated the suppressive capacity of the CD25+CD4+ Treg cells with serum 25-OHD levels in MS patients (222), but this association could not be substantiated upon vitamin D3 supplementation (223). A more recent study correlated seasonal increases in 25-OHD and calcitriol in healthy men with increased Treg cell Foxp3 expression, a drop in memory CD4+CD45RO+ T cells, a reduced capacity for T-cell pro-inflammatory cytokine production, and increased CCR4, CCR6, CLA, CCR9, and CCR7 levels for homing to skin, gut, and lymphoid tissue (310). Seemingly contradictory results were reported for infants whose 25-OHD levels and T-cell subsets in cord blood showed no correlations (311). Potential confounders in human studies are differences in phenotypic markers used in the analyses, vitamin D3 status when beginning vitamin D3 supplementation, supplementation protocols (135), use of disease-modifying drugs in patients, and timing of sampling relative to estrogen cycling in women.

Examining human T cells in culture documented an influence of calcitriol on FoxP3 (Figure 1C). Activation of human CD4+CD25 T cells in vitro in the presence of calcitriol resulted in greater numbers of CD4+CD25+FoxP3+ Treg cells (221), particularly for female T cells (172). Addition of calcitriol to human CD4+CD25 T-cell cultures increased CTLA4 (CD152) and FoxP3 protein expression, the latter requiring the presence of IL-2 (218). The increase in CTLA4 and FoxP3 protein expression was confirmed for MS patient T cells (219), and the CTLA4 increase was confirmed for T1D patient T cells (220). The suppressive function of the calcitriol-induced CD4+CD25+FoxP3+ Treg cells was confirmed in many reports (218, 220, 221, 312).

Calcitriol positive regulation of CTLA4 is critically important, because this protein is an essential negative regulator of immune responses (313, 314). Humans who were heterozygous for loss-of-function mutations in the Ctla4 gene developed a highly penetrant autoimmune syndrome with strong similarity to IPEX. Specifically, they displayed loss of CD4+FoxP3+ Treg cell suppressive function, hyperactivation of effector T cells, lymphocyte infiltration of multiple target organs, and fatal autoimmune disease. CTLA4 strips CD80 and CD86 from the APC by endocytosing and degrading these proteins to block their co-stimulation of CD28 on T cells (315). Calcitriol enhancement of this mechanism may account in part for the hormone’s actions as an immune response terminator.

Intriguing new data complemented the rodent data (185) and revealed that calcitriol increased human FOXP3 gene expression by a transcriptional mechanism (217). The human FOXP3 and murine Foxp3 genes share homology in a conserved non-coding sequence (+1714 to +2554 relative to the FOXP3 transcriptional start site) that functions as an enhancer (316). Within this enhancer region, investigators identified three VDREs that enhanced promoter activity in a calcitriol-dependent manner (217). Moreover, T-cell stimulation through CD3 and CD28 in the presence of IL-2 and calcitriol generated CD4+CD25+FoxP3+ Treg cells that inhibited target CD4+ T-cell proliferation by a cell contact- and FOXP3-dependent mechanism. Collectively, the data demonstrating calcitriol enhancement of Ikzf2, FOXP3, and Ctla4 gene expression suggest that vitamin D may significantly influence the emergence of an autoimmune disease phenotype by increasing CD4+CD25+FoxP3+ Treg cells and/or their suppressive function.

Vitamin D, CD46, and IL-10-Producing Tr1 Lymphocytes

Important new research into the function of CD46 has demonstrated the existence of a calcitriol-mediated Th1–Tr1 switch controlling the delicate homeostatic balance between the pro-inflammatory and anti-inflammatory states (Figure 2). CD46 and its murine analog complement receptor 1-related protein Y (Crry) are transmembrane glycoproteins that bind complement fragments C3b and C4b, and function as T-cell costimulatory molecules (269). Temporal processing of two alternatively spliced CD46 cytoplasmic tails, Cyt1 and Cyt2, served as a molecular rheostat controlling activation and de-activation of human T-cell responses (317). Co-engagement of the TCR and CD46 first enhanced Th1 cell effector function (318). However, as IL-2 accumulated, CD46 cytoplasmic domain interactions with the serine–threonine kinase SPAK promoted effector Th1 cell switching to a Tr1 phenotype with low IL-2 production, high IL-10 production, and the capability to suppress bystander T-cell activation. Thus, CD46 served as a molecular link between complement activation and a Th1–Tr1 switch for immune response termination (269). This switch was defective in MS (265, 266), asthma (319), and rheumatoid arthritis patient T cells (318), effectively locking the immune response in an interminable pro-inflammatory state.

FIGURE 2
www.frontiersin.org

Figure 2. A proposed Th1–Tr1 bi-stable biological switch bridging two immune states, pro-inflammatory and anti-inflammatory. (A) A theoretical diagram of two opposing immune states [adapted from a generalized bi-stable switch diagram (302)]. Each circle represents T cells that are in a stable biological state characterized by a particular gene regulatory network and capable of responding to external stimuli in a biologically appropriate manner. Biological barriers impede transitions between the pro- and anti-inflammatorystates. Each biological state is capable of reinforcing its own state, denoted as lines terminating in arrows, and inhibiting the opposing state, denoted as blocked lines. External stimuli, for example pathogens, hormones, cell-cell interactions, and soluble mediators serve as switch inputs to drive a transition. For example, signals 1 and 2 can be organized as “1 or 2”, “1 and 2”, or “1 not 2” to trigger a dynamic transition between the two states. (B) The proposed Th1–Tr1 switch appears to have the qualities of a bi-stable biological switch. The Th1 cytokine IFN-γ and the Tr1 cytokine IL-10 impede the development of Tr1 and Th1 cells, respectively. The autocrine growth factors IL-2 and IL-10 reinforce the survival and proliferation of Th1 and Tr1 cells, respectively. Diminishing pathogen, sensed through TLR molecule engagement (280), maturation of the pathogen-specific immune response, sensed through CD46 molecule engagement by C3b and C4b (299, 300), and accumulating calcitriol in the microenvironment, sensed through the VDR in Th1 cells, provide the external stimuli to alter the dynamics of the system and drive the transition to the Tr1 biological state. Helios may be a biomarker of effector T cells undergoing a transition to a tolerant state (183). Calcitriol increased Helios in CD4+ T cells at the beginning of a calcitriol-induced EAE disease remission (144), and was recently identified as an inhibitor of Il2 gene transcription by an epigenetic silencing mechanism (184).

Calcitriol profoundly amplified the Th1–Tr1 switch in vitro in T cells from healthy donors and patients with MS (219). Calcitriol also amplified this switch in vitro in T cells from T1D patients, imprinting the T cells with an IL-10-producing Tr1 phenotype, and modulating surface expression of chemokine receptors to enable homing to inflamed sites (220, 312). Determining whether this switch functions in vivo is a high priority goal. The emerging data suggest that vitamin D may significantly influence the emergence of an autoimmune disease phenotype by controlling a Th1–Tr1 switch that provides a bridge between two apparently stable immune states, pro-inflammatory and anti-inflammatory. The calcitriol-driven Th1–Tr1 switch is of utmost importance to autoimmune disease, because each state appears to be self-reinforcing and capable of blocking the opposing state (Figure 2). Multistable switches are known to play an important role in hematopoietic cell fate decisions controlled by gene regulatory networks (320), and hormones are known to be of paramount importance in controlling these networks.

Calcitriol enhancement of the transcription factor Helios may also be important to the Th1–Tr1 switch. Calcitriol treatment transiently induced CD4+Helios+ T cells in the CNS of mice with EAE prior to the loss of pathogenic CD4+ T cells from the CNS and the onset of remission (216). New data show Helios was a marker for peripheral T cells that are being driven to tolerance in response to a genuine autoantigen in autoimmune gastritis (321). Helios was identified as a key inhibitor of Il2 gene transcription in Treg cells (322); it repressed the Il2 locus through epigenetic modifications that included histone deacetylation.

Given the importance of pathways downstream of calcitriol signaling to the Th1–Tr1 switch and immune response termination, it is valuable to consider how those pathways could be compromised. CD46 is the receptor for the T lymphotropic human herpesvirus-6; the virions bind CD46 and trigger endocytosis (323). Whether pathogen-mediated CD46 removal would undermine calcitriol action by disabling the Th1–Tr1 switch is not known. Genetically determined defects in IL-10 production and IL-10R signaling would be downstream of a calcitriol-mediated increase in Tr1 cells; both defects were linked to an increased autoimmune disease risk (324329). The IL-10R signals through tyrosine kinase 2 (TYK2) and Janus kinase-1 (330). A novel missense mutation in the TYK2 gene was linked to an increased risk of MS (329). In addition, recombinant hIL-10 did not inhibit proliferation of CD4+ T cells from MS patients; the IL-10R signaling pathway was blocked at the point of STAT3 activation in these cells (267). Defective IL-10R signaling downstream of calcitriol action to induce Tr1 cells could compromise this vitamin D-mediated protective mechanism.

Conclusion and Future Research Directions

This review article sought to summarize and integrate research on vitamin D and CD4+ T-lymphocyte biology in an effort to develop a new understanding of the molecular etiology of autoimmune disease. There is a large, latitude-linked, non-transmissible environmental component that acts at the population level to determine the emergence of an autoimmune phenotype, given an autoimmune disease-susceptible genotype. Diverse and compelling evidence suggests that this environmental component is ambient UVB light exposure catalyzing cutaneous vitamin D3 formation. Importantly, this environmental component exerts a >100-fold influence on the risk of multiple autoimmune diseases, and on disparate populations with distinct genotypes, dietary habits, and exposures to infectious and commensal organisms, larger than the influence of the strongest autoimmune susceptibility locus (MHCII genes; approximately fivefold) or the infectious organism, EBV (approximately two to threefold).

The strength and universality of ambient UVB light and vitamin D3 as autoimmune disease risk factors, the vitamin D hormone’s established role as a transcriptional regulator of gene expression, and the role of CD4+ T lymphocytes as autoimmune disease initiators and suppressors provided the rationale for closely examining calcitriol regulation of CD4+ T-lymphocyte function in this review. The conclusions supported by experimental evidence are presented in Box 1. These conclusions address the Bradford Hill criterion of a plausible biological mechanism that coheres with known biological facts (115). This evidence has contributed to a greater functional and mechanistic understanding of the molecular etiology of autoimmune disease, one of the major challenges in modern immunobiology.

Box 1. Summary of main conclusions regarding vitamin D, CD4+ T lymphocytes, and autoimmune disease.

An autoimmune disease phenotype emerges when modifiable environmental stressors act on a disease-susceptible genotype, and exposure to at least one environmental stressor is increasing.

Vitamin D is probably the environmental factor with the greatest influence on the emergence of an autoimmune disease phenotype given a disease-susceptible genotype.

At high latitudes, there is a period of “light starvation” when vitamin D synthesis is negligible; the higher the latitude, the greater is the light starvation period.

Paracrine calcitriol signaling to CD4+ T cells within tissues is likely the major pathway by which sunlight as an environmental factor exerts its influence on the emergence of an autoimmune disease phenotype.

The VDR–RXR complex regulates gene expression in a cell type-specific manner through VDRE in calcitriol-responsive genes.

The hormonal form of vitamin D may influence thymic selection, which would have a significant influence on autoimmune disease risk acquisition and the timing of autoimmune disease prevention initiatives.

The hormonal form of vitamin D may significantly influence autoimmune disease risk and severity by (i) dampening pathogenic Th17 cell IL-17 synthesis, (ii) increasing effector CD4+ T-cell sensitivity to extrinsic cell death signals, (iii) promoting CD4+CD25+FoxP3+ Treg cell and CD4+IL-10+FoxP3 Tr1 cell development and suppressive function, (iv) amplifying a Th1–Tr1 switch that may bridge two apparently stable immune states, an anti-pathogen pro-inflammatory state and a self-tolerant anti-inflammatory state.

During periods of light starvation, phototherapy to increase FoxP3+ Treg cell development may be a particularly efficient method of influencing autoimmune disease risk and severity.

Important questions remain. Regarding paracrine calcitriol signaling to CD4+ T cells, more work is needed in vivo to understand the cellular sources of calcitriol, the signals and kinetics that induce and terminate calcitriol synthesis and their relationship to infectious organisms. More work is also needed to understand the signals and kinetics that induce and terminate VDR and CYP24A1 gene expression in specific types of CD4+ T cells in vivo, and how gender influences calcitriol responsiveness and calcitriol turnover. Finally, we must learn what factors and mechanisms disrupt paracrine signaling (genetic lesions, drugs, oxidative damage to enzymes, epigenetic dysregulation).

A second and related high priority is probing a possible vitamin D influence on thymic selection. Such an influence would have high significance for our understanding of the molecular etiology of autoimmune disease risk acquisition and our effort to correctly time disease prevention efforts. The data correlating seasonal variation in thymic output inversely with vitamin D status await replication. Assuming reproducibility, mechanistic inquiry will be needed to distinguish potential impacts on thymopoiesis and positive selection from negative selection, and to decipher the calcitriol-regulated genes that influence these processes.

Other questions relate to calcitriol actions on effector CD4+ T cells in vivo. Calcitriol suppression of Il17 transcription and IL-17 protein synthesis has been studied mechanistically in rodents with conflicting results. It will be important to clarify the biochemistry of Il17 transcriptional and post-transcriptional regulation by calcitriol in Th17 cells in vivo in the context of different autoimmune disease settings where physiologically relevant tissue architecture and cell and cytokine microenvironments exist. Other open questions concern the biochemistry of the effector CD4+ T-cell programed cell death pathways that are influenced by calcitriol, and the cellular source and timing of the extrinsic cell death signaling.

Another very high priority is probing how vitamin D promotes CD4+FoxP3+IL-10 Treg and CD4+FoxP3IL-10+ Tr1 cell development, stability, and suppressive function. Unanswered questions relate to the biochemical mechanisms for the proposed calcitriol up-regulation of Ikzf2, Il10, Ctla4, and CD200 gene transcription, and the relationship between VDR and FOXP3, how each gene may impact the other’s expression and the kinetics and stability of the interaction.

Additional questions concern the molecular details of the proposed Th1–Tr1 switch. Whether this switch functions in vivo in the context of different autoimmune diseases is a top priority question. Moreover, questions remain regarding the precise biochemical details of the switch, the signals that trigger it, how the signals are transduced to the nucleus, how the gene expression program is altered by those signals, and exactly how calcitriol amplifies the switch. This molecular knowledge is needed to envision and probe how pathogen-mediated engagement of CD46 might compromise the switch. Mathematical modeling of the switch may be helpful to envision interactions between environmental pathogens, the sun-sensitive hormone, the pathogen-specific immune response, immune response termination, and self-tolerance maintenance.

Sir Austin Bradford Hill suggested minimal criteria for judging the causal nature of relationships between environmental factors and disease (115). Many experts believe the experimental evidence linking vitamin D3 with autoimmune disease risk is sufficiently compelling to satisfy eight of the nine criteria (6). Perhaps, the most important unanswered question derives from Bradford Hill’s ninth criterion, experiment. Would an intervention that substantially elevates circulating 25-OHD3 prevent or change autoimmune disease in light-starved populations with low vitamin D status? Now would be a good time for autoimmune disease researchers to address this unanswered question collaboratively.

Conflict of Interest Statement

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Acknowledgments

We thank Drs. A. Astier and D. Hafler for their thoughtful invitation to contribute this review. We gratefully acknowledge our colleagues for contributing important ideas to our research, particularly Sir D. Acheson, Drs. G. C. Ebers, B. Dittel, K. Spach, J. Spanier, C. Mayne, C. Nelson, I. Amanna, K. Hoag, D. Miller, B. Yandell, J. Gorski, R. Fritz, and W. Woodward. We also thank L. Pedersen and P. Salzwedel. We are especially grateful to P. O’Looney and C. Kalt of the National Multiple Sclerosis Society for unwavering encouragement, and T. Wolfe and P. Powers for establishing the “Multiple Sclerosis Research Fund in Biochemistry” at the University of Wisconsin Foundation. We are indebted to the University of Wisconsin Graduate School Research Committee research grant #13008, a HATCH McIntyre Stennis Award (MSN119798, PRJ18KV), and a National Multiple Sclerosis Society award (RG4076A5/1) for support.

Abbreviations

25-OHD3, 25-hydroxyvitamin D3; APC, antigen-presenting cells; calcitriol, 1α,25-dihydroxyvitamin D3; CNS, central nervous system; CTLA4, cytotoxic T-lymphocyte antigen 4; DC, dendritic cell; EAE, experimental autoimmune encephalomyelitis; EAU, experimental autoimmune uveitis; EBV, Epstein-Barr virus; IFN, interferon; IL, interleukin; MHC, major histocompatibility complex; MRI, magnetic resonance imaging; MS, multiple sclerosis; NOD, non-obese diabetic; RANK, receptor-activator of NF-kappaB; RANKL, receptor-activator of NF-kappaB ligand; T1D, type 1 diabetes; TCR, T-cell receptor; Th, T helper; TLR, Toll-like receptor; Tr1, T-regulatory cell type 1; Treg, T-regulatory; TYK2, tyrosine kinase 2; UV, ultraviolet; VDR, vitamin D receptor; VDRE, vitamin D response element.

References

1. Blumberg RS, Dittel B, Hafler D, von Herrath M, Nestle FO. Unraveling the autoimmune translational research process layer by layer. Nat Med (2012) 18:35–41. doi:10.1038/nm.2632

CrossRef Full Text | Google Scholar

2. Kingwell E, Marriott JJ, Jette N, Pringsheim T, Makhani N, Morrow SA, et al. Incidence and prevalence of multiple sclerosis in Europe: a systematic review. BMC Neurol (2013) 13:128. doi:10.1186/1471-2377-13-128

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

3. Evans C, Beland SG, Kulaga S, Wolfson C, Kingwell E, Marriott J, et al. Incidence and prevalence of multiple sclerosis in the Americas: a systematic review. Neuroepidemiology (2013) 40:195–210. doi:10.1159/000342779

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

4. Patterson C, Guariguata L, Dahlquist G, Soltesz G, Ogle G, Silink M. Diabetes in the young – a global view and worldwide estimates of numbers of children with type 1 diabetes. Diabetes Res Clin Pract (2014) 103:161–75. doi:10.1016/j.diabres.2013.11.005

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

5. Singleton AB. Genetics. A unified process for neurological disease. Science (2014) 343:497–8. doi:10.1126/science.1250172

CrossRef Full Text | Google Scholar

6. Hayes CE, Nelson CD, Spanier JA. Vitamin D and autoimmunity. In: Gombart AF, editor. Vitamin D: Oxidation, Immunity, and Aging. Boca Raton: Taylor & Francis Group, CRC Press (2012). p. 239–306.

Google Scholar

7. Antico A, Tampoia M, Tozzoli R, Bizzaro N. Can supplementation with vitamin D reduce the risk or modify the course of autoimmune diseases? A systematic review of the literature. Autoimmun Rev (2012) 12:127–36. doi:10.1016/j.autrev.2012.07.007

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

8. Friese MA, Schattling B, Fugger L. Mechanisms of neurodegeneration and axonal dysfunction in multiple sclerosis. Nat Rev Neurol (2014) 10:225–38. doi:10.1038/nrneurol.2014.37

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

9. Atkinson MA, Eisenbarth GS, Michels AW. Type 1 diabetes. Lancet (2014) 383:69–82. doi:10.1016/S0140-6736(13)60591-7

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

10. Herold KC, Vignali DA, Cooke A, Bluestone JA. Type 1 diabetes: translating mechanistic observations into effective clinical outcomes. Nat Rev Immunol (2013) 13:243–56. doi:10.1038/nri3422

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

11. Chao MJ, Ramagopalan SV, Herrera BM, Orton SM, Handunnetthi L, Lincoln MR, et al. MHC transmission: insights into gender bias in MS susceptibility. Neurology (2011) 76:242–6. doi:10.1212/WNL.0b013e318207b060

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

12. Isobe N, Damotte V, Lo Re V, Ban M, Pappas D, Guillot-Noel L, et al. Genetic burden in multiple sclerosis families. Genes Immun (2013) 14:434–40. doi:10.1038/gene.2013.37

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

13. Steck AK, Rewers MJ. Genetics of type 1 diabetes. Clin Chem (2011) 57:176–85. doi:10.1373/clinchem.2010.148221

CrossRef Full Text | Google Scholar

14. Baranzini SE, Mudge J, van Velkinburgh JC, Khankhanian P, Khrebtukova I, Miller NA, et al. Genome, epigenome and RNA sequences of monozygotic twins discordant for multiple sclerosis. Nature (2010) 464:1351–6. doi:10.1038/nature08990

CrossRef Full Text | Google Scholar

15. Stefan M, Zhang W, Concepcion E, Yi Z, Tomer Y. DNA methylation profiles in type 1 diabetes twins point to strong epigenetic effects on etiology. J Autoimmun (2014) 50:33–7. doi:10.1016/j.jaut.2013.10.001

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

16. Halfvarson J. Genetics in twins with Crohn’s disease: less pronounced than previously believed? Inflamm Bowel Dis (2011) 17:6–12. doi:10.1002/ibd.21295

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

17. Petersen BS, Spehlmann ME, Raedler A, Stade B, Thomsen I, Rabionet R, et al. Whole genome and exome sequencing of monozygotic twins discordant for Crohn’s disease. BMC Genomics (2014) 15:564. doi:10.1186/1471-2164-15-564

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

18. Furukawa H, Oka S, Matsui T, Hashimoto A, Arinuma Y, Komiya A, et al. Genome, epigenome and transcriptome analyses of a pair of monozygotic twins discordant for systemic lupus erythematosus. Hum Immunol (2013) 74:170–5. doi:10.1016/j.humimm.2012.11.007

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

19. Handel AE, Giovannoni G, Ebers GC, Ramagopalan SV. Environmental factors and their timing in adult-onset multiple sclerosis. Nat Rev Neurol (2010) 6:156–66. doi:10.1038/nrneurol.2010.1

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

20. Ebers G. Interactions of environment and genes in multiple sclerosis. J Neurol Sci (2013) 334:161–3. doi:10.1016/j.jns.2013.08.018

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

21. Acheson ED, Bachrach CA, Wright FM. Some comments on the relationship of the distribution of multiple sclerosis to latitude, solar radiation and other variables. Acta Psychiatr Scand Suppl (1960) 35(Suppl 147):132–47. doi:10.1111/j.1600-0447.1960.tb08674.x

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

22. Simpson S Jr, Blizzard L, Otahal P, Van der Mei I, Taylor B. Latitude is significantly associated with the prevalence of multiple sclerosis: a meta-analysis. J Neurol Neurosurg Psychiatry (2011) 82:1132–41. doi:10.1136/jnnp.2011.240432

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

23. Ebers GC. Environmental factors and multiple sclerosis. Lancet Neurol (2008) 7:268–77. doi:10.1016/S1474-4422(08)70042-5

CrossRef Full Text | Google Scholar

24. Dyment DA, Yee IM, Ebers GC, Sadovnick AD, G. Canadian Collaborative Study. Multiple sclerosis in stepsiblings: recurrence risk and ascertainment. J Neurol Neurosurg Psychiatry (2006) 77:258–9. doi:10.1136/jnnp.2005.063008

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

25. Orton SM, Herrera BM, Yee IM, Valdar W, Ramagopalan SV, Sadovnick AD, et al. Sex ratio of multiple sclerosis in Canada: a longitudinal study. Lancet Neurol (2006) 5:932–6. doi:10.1016/S1474-4422(06)70581-6

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

26. Orton SM, Ramagopalan SV, Brocklebank D, Herrera BM, Dyment DA I, Yee M, et al. Effect of immigration on multiple sclerosis sex ratio in Canada: the Canadian collaborative study. J Neurol Neurosurg Psychiatry (2010) 81:31–6. doi:10.1136/jnnp.2008.162784

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

27. Sadovnick AD. European Charcot foundation lecture: the natural history of multiple sclerosis and gender. J Neurol Sci (2009) 286:1–5. doi:10.1016/j.jns.2009.09.005

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

28. McKenzie RL, Liley JB, Bjorn LO. UV radiation: balancing risks and benefits. Photochem Photobiol (2009) 85:88–98. doi:10.1111/j.1751-1097.2008.00400.x

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

29. Embry AF, Snowdon LR, Vieth R. Vitamin D and seasonal fluctuations of gadolinium-enhancing magnetic resonance imaging lesions in multiple sclerosis. Ann Neurol (2000) 48:271–2. doi:10.1002/1531-8249(200008)48:2<271::AID-ANA28>3.3.CO;2-F

CrossRef Full Text | Google Scholar

30. Soilu-Hanninen M, Laaksonen M, Laitinen I, Eralinna JP, Lilius EM, Mononen I. A longitudinal study of serum 25-hydroxyvitamin D and intact parathyroid hormone levels indicate the importance of vitamin D and calcium homeostasis regulation in multiple sclerosis. J Neurol Neurosurg Psychiatry (2008) 79:152–7. doi:10.1136/jnnp.2006.105320

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

31. Meier DS, Balashov KE, Healy B, Weiner HL, Guttmann CR. Seasonal prevalence of MS disease activity. Neurology (2010) 75:799–806. doi:10.1212/WNL.0b013e3181f0734c

CrossRef Full Text | Google Scholar

32. van der Mei IA, Ponsonby AL, Dwyer T, Blizzard L, Simmons R, Taylor BV, et al. Past exposure to sun, skin phenotype, and risk of multiple sclerosis: case-control study. BMJ (2003) 327:316. doi:10.1136/bmj.327.7410.316

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

33. Kampman MT, Wilsgaard T, Mellgren SI. Outdoor activities and diet in childhood and adolescence relate to MS risk above the Arctic circle. J Neurol (2007) 254:471–7. doi:10.1007/s00415-006-0395-5

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

34. Islam T, Gauderman WJ, Cozen W, Mack TM. Childhood sun exposure influences risk of multiple sclerosis in monozygotic twins. Neurology (2007) 69:381–8. doi:10.1212/01.wnl.0000268266.50850.48

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

35. McDowell TY, Amr S, Culpepper WJ, Langenberg P, Royal W, Bever C, et al. Sun exposure, vitamin D and age at disease onset in relapsing multiple sclerosis. Neuroepidemiology (2011) 36:39–45. doi:10.1159/000322512

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

36. Goldacre MJ, Seagroatt V, Yeates D, Acheson ED. Skin cancer in people with multiple sclerosis: a record linkage study. J Epidemiol Community Health (2004) 58:142–4. doi:10.1136/jech.58.2.142

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

37. Mohr SB, Garland CF, Gorham ED, Garland FC. The association between ultraviolet B irradiance, vitamin D status and incidence rates of type 1 diabetes in 51 regions worldwide. Diabetologia (2008) 51:1391–8. doi:10.1007/s00125-008-1061-5

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

38. Moltchanova EV, Schreier N, Lammi N, Karvonen M. Seasonal variation of diagnosis of Type 1 diabetes mellitus in children worldwide. Diabet Med (2009) 26:673–8. doi:10.1111/j.1464-5491.2009.02743.x

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

39. Staples JA, Ponsonby AL, Lim LL, McMichael AJ. Ecologic analysis of some immune-related disorders, including type 1 diabetes, in Australia: latitude, regional ultraviolet radiation, and disease prevalence. Environ Health Perspect (2003) 111:518–23. doi:10.1289/ehp.5941

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

40. Godar DE. UV doses worldwide. Photochem Photobiol (2005) 81:736–49. doi:10.1111/j.1751-1097.2005.tb01438.x

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

41. Sloka S, Grant M, Newhook LA. Time series analysis of ultraviolet B radiation and type 1 diabetes in Newfoundland. Pediatr Diabetes (2008) 9:81–6. doi:10.1111/j.1399-5448.2007.00310.x

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

42. Ascherio A. Environmental factors in multiple sclerosis. Expert Rev Neurother (2013) 13:3–9. doi:10.1586/14737175.2013.865866

CrossRef Full Text | Google Scholar

43. Lossius A, Riise T, Pugliatti M, Bjornevik K, Casetta I, Drulovic J, et al. Season of infectious mononucleosis and risk of multiple sclerosis at different latitudes; the EnvIMS Study. Mult Scler (2014) 20:669–74. doi:10.1177/1352458513505693

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

44. Ascherio A, Munger KL. Epstein-Barr virus infection and multiple sclerosis: a review. J Neuroimmune Pharmacol (2010) 5:271–7. doi:10.1007/s11481-010-9201-3

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

45. Disanto G, Meier U, Giovannoni G, Ramagopalan SV. Vitamin D: a link between Epstein-Barr virus and multiple sclerosis development? Expert Rev Neurother (2011) 11:1221–4. doi:10.1586/ern.11.97

CrossRef Full Text | Google Scholar

46. Hayes CE, Cantorna MT, DeLuca HF. Vitamin D and multiple sclerosis. Proc Soc Exp Biol Med (1997) 216:21–7. doi:10.3181/00379727-216-44153A

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

47. Smolders J, Damoiseaux J. Vitamin D as a T-cell modulator in multiple sclerosis. Vitam Horm (2011) 86:401–28. doi:10.1016/B978-0-12-386960-9.00018-6

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

48. D’Aurizio F, Villalta D, Metus P, Doretto P, Tozzoli R. Is vitamin D a player or not in the pathophysiology of autoimmune thyroid diseases? Autoimmun Rev (2014). doi:10.1016/j.autrev.2014.10.008

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

49. Lopez ER, Regulla K, Pani MA, Krause M, Usadel KH, Badenhoop K. CYP27B1 polymorphisms variants are associated with type 1 diabetes mellitus in Germans. J Steroid Biochem Mol Biol (2004) 89-90:155–7. doi:10.1016/j.jsbmb.2004.03.095

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

50. Bailey R, Cooper JD, Zeitels L, Smyth DJ, Yang JH, Walker NM, et al. Association of the vitamin D metabolism gene CYP27B1 with type 1 diabetes. Diabetes (2007) 56:2616–21. doi:10.2337/db07-0652

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

51. Ramos-Lopez E, Bruck P, Jansen T, Herwig J, Badenhoop K. CYP2R1 (vitamin D 25-hydroxylase) gene is associated with susceptibility to type 1 diabetes and vitamin D levels in Germans. Diabetes Metab Res Rev (2007) 23:631–6. doi:10.1002/dmrr.719

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

52. Cooper JD, Smyth DJ, Walker NM, Stevens H, Burren OS, Wallace C, et al. Inherited variation in vitamin d genes is associated with predisposition to autoimmune disease type 1 diabetes. Diabetes (2011) 60:1624–31. doi:10.2337/db10-1656

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

53. Hussein AG, Mohamed RH, Alghobashy AA. Synergism of CYP2R1 and CYP27B1 polymorphisms and susceptibility to type 1 diabetes in Egyptian children. Cell Immunol (2012) 279:42–5. doi:10.1016/j.cellimm.2012.08.006

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

54. Frederiksen BN, Kroehl M, Fingerlin TE, Wong R, Steck AK, Rewers M, et al. Association between vitamin D metabolism gene polymorphisms and risk of islet autoimmunity and progression to type 1 diabetes: the diabetes autoimmunity study in the young (DAISY). J Clin Endocrinol Metab (2013) 98:E1845–51. doi:10.1210/jc.2013-2256

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

55. Fichna M, Zurawek M, Januszkiewicz-Lewandowska D, Gryczynska M, Fichna P, Sowinski J, et al. Association of the CYP27B1 C(-1260)A polymorphism with autoimmune Addison’s disease. Exp Clin Endocrinol Diabetes (2010) 118:544–9. doi:10.1055/s-0029-1241206

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

56. Lopez ER, Zwermann O, Segni M, Meyer G, Reincke M, Seissler J, et al. A promoter polymorphism of the CYP27B1 gene is associated with Addison’s disease, Hashimoto’s thyroiditis, Graves’ disease and type 1 diabetes mellitus in Germans. Eur J Endocrinol (2004) 151:193–7. doi:10.1530/eje.0.1510193

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

57. Torkildsen O, Knappskog PM, Nyland HI, Myhr KM. Vitamin D-dependent rickets as a possible risk factor for multiple sclerosis. Arch Neurol (2008) 65:809–11. doi:10.1001/archneur.65.6.809

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

58. Ramagopalan SV, Dyment DA, Cader MZ, Morrison KM, Disanto G, Morahan JM, et al. Rare variants in the CYP27B1 gene are associated with multiple sclerosis. Ann Neurol (2011) 70:881–6. doi:10.1002/ana.22678

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

59. Alcina A, Fedetz M, Fernandez O, Saiz A, Izquierdo G, Lucas M, et al. Identification of a functional variant in the KIF5A-CYP27B1-METTL1-FAM119B locus associated with multiple sclerosis. J Med Genet (2013) 50:25–33. doi:10.1136/jmedgenet-2012-101085

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

60. Ross JP, Bernales CQ, Lee JD, Sadovnick AD, Traboulsee AL, Vilarino-Guell C. Analysis of CYP27B1 in multiple sclerosis. J Neuroimmunol (2014) 266:64–6. doi:10.1016/j.jneuroim.2013.11.006

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

61. Sawcer S, Hellenthal G, Pirinen M, Spencer CC, Patsopoulos NA, Moutsianas L, et al. Genetic risk and a primary role for cell-mediated immune mechanisms in multiple sclerosis. Nature (2011) 476:214–9. doi:10.1038/nature10251

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

62. Pociot F, Akolkar B, Concannon P, Erlich HA, Julier C, Morahan G, et al. Genetics of type 1 diabetes: what’s next? Diabetes (2010) 59:1561–71. doi:10.2337/db10-0076

CrossRef Full Text | Google Scholar

63. Barizzone N, Pauwels I, Luciano B, Franckaert D, Guerini FR, Cosemans L, et al. No evidence for a role of rare CYP27B1 functional variations in multiple sclerosis. Ann Neurol (2013) 73:433–7. doi:10.1002/ana.23834

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

64. Ban M, Caillier S I, Mero L, Myhr KM, Celius EG, Aarseth J, et al. No evidence of association between mutant alleles of the CYP27B1 gene and multiple sclerosis. Ann Neurol (2013) 73:430–2. doi:10.1002/ana.23833

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

65. Reinthaler E, Machetanz G, Hotzy C, Reindl M, Fazekas F, Kristoferitsch W, et al. No evidence for a role of rare CYP27B1 variants in Austrian multiple sclerosis patients. Mult Scler (2014) 20:391–2. doi:10.1177/1352458513498130

CrossRef Full Text | Google Scholar

66. Smolders J, Peelen E, Thewissen M, Menheere P, Cohen Tervaert JW, Hupperts R, et al. The relevance of vitamin D receptor gene polymorphisms for vitamin D research in multiple sclerosis. Autoimmun Rev (2009) 8:621–6. doi:10.1016/j.autrev.2009.02.009

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

67. Berlanga-Taylor AJ, Disanto G, Ebers GC, Ramagopalan SV. Vitamin D-gene interactions in multiple sclerosis. J Neurol Sci (2011) 311:32–6. doi:10.1016/j.jns.2011.08.041

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

68. Huang J, Xie ZF. Polymorphisms in the vitamin D receptor gene and multiple sclerosis risk: a meta-analysis of case-control studies. J Neurol Sci (2012) 313:79–85. doi:10.1016/j.jns.2011.09.024

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

69. Niino M, Fukazawa T, Yabe I, Kikuchi S, Sasaki H, Tashiro K. Vitamin D receptor gene polymorphism in multiple sclerosis and the association with HLA class II alleles. J Neurol Sci (2000) 177:65–71. doi:10.1016/S0022-510X(00)00336-1

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

70. Ramagopalan SV, Maugeri NJ, Handunnetthi L, Lincoln MR, Orton SM, Dyment DA, et al. Expression of the multiple sclerosis-associated MHC class II Allele HLA-DRB1*1501 is regulated by vitamin D. PLoS Genet (2009) 5:e1000369. doi:10.1371/journal.pgen.1000369

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

71. Gross C, Krishnan AV, Malloy PJ, Eccleshall TR, Zhao XY, Feldman D. The vitamin D receptor gene start codon polymorphism: a functional analysis of FokI variants. J Bone Miner Res (1998) 13:1691–9. doi:10.1359/jbmr.1998.13.11.1691

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

72. Jurutka PW, Remus LS, Whitfield GK, Thompson PD, Hsieh JC, Zitzer H, et al. The polymorphic N terminus in human vitamin D receptor isoforms influences transcriptional activity by modulating interaction with transcription factor IIB. Mol Endocrinol (2000) 14:401–20. doi:10.1210/mend.14.3.0435

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

73. Partridge JM, Weatherby SJ, Woolmore JA, Highland DJ, Fryer AA, Mann CL, et al. Susceptibility and outcome in MS: associations with polymorphisms in pigmentation-related genes. Neurology (2004) 62:2323–5. doi:10.1212/WNL.62.12.2323

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

74. Orton SM, Morris AP, Herrera BM, Ramagopalan SV, Lincoln MR, Chao MJ, et al. Evidence for genetic regulation of vitamin D status in twins with multiple sclerosis. Am J Clin Nutr (2008) 88:441–7.

Pubmed Abstract | Pubmed Full Text | Google Scholar

75. Mamutse G, Woolmore J, Pye E, Partridge J, Boggild M, Young C, et al. Vitamin D receptor gene polymorphism is associated with reduced disability in multiple sclerosis. Mult Scler (2008) 14:1280–3. doi:10.1177/1352458508094643

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

76. Smolders J, Damoiseaux J, Menheere P, Tervaert JW, Hupperts R. Fok-I vitamin D receptor gene polymorphism (rs10735810) and vitamin D metabolism in multiple sclerosis. J Neuroimmunol (2009) 207:117–21. doi:10.1016/j.jneuroim.2008.12.011

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

77. Arai H, Miyamoto K, Taketani Y, Yamamoto H, Iemori Y, Morita K, et al. A vitamin D receptor gene polymorphism in the translation initiation codon: effect on protein activity and relation to bone mineral density in Japanese women. J Bone Miner Res (1997) 12:915–21. doi:10.1359/jbmr.1997.12.6.915

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

78. Dickinson J, Perera D, van der Mei A, Ponsonby AL, Polanowski A, Thomson R, et al. Past environmental sun exposure and risk of multiple sclerosis: a role for the Cdx-2 Vitamin D receptor variant in this interaction. Mult Scler (2009) 15:563–70. doi:10.1177/1352458509102459

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

79. Ponsonby AL, Pezic A, Ellis J, Morley R, Cameron F, Carlin J, et al. Variation in associations between allelic variants of the vitamin D receptor gene and onset of type 1 diabetes mellitus by ambient winter ultraviolet radiation levels: a meta-regression analysis. Am J Epidemiol (2008) 168:358–65. doi:10.1093/aje/kwn142

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

80. Barrett JC, Clayton DG, Concannon P, Akolkar B, Cooper JD, Erlich HA, et al. Genome-wide association study and meta-analysis find that over 40 loci affect risk of type 1 diabetes. Nat Genet (2009) 41:703–7. doi:10.1038/ng.381

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

81. Israni N, Goswami R, Kumar A, Rani R. Interaction of vitamin D receptor with HLA DRB1 0301 in type 1 diabetes patients from North India. PLoS One (2009) 4:e8023. doi:10.1371/journal.pone.0008023

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

82. Mellanby M, Pattison CL. The action of vitamin D in preventing the spread and promoting the arrest of caries in children. Br Med J (1928) 2:1079–82. doi:10.1136/bmj.2.3545.1079

CrossRef Full Text | Google Scholar

83. Craelius W. Comparative epidemiology of multiple sclerosis and dental caries. J Epidemiol Community Health (1978) 32:155–65. doi:10.1136/jech.32.3.155

CrossRef Full Text | Google Scholar

84. Spach KM, Hayes CE. Vitamin D3 confers protection from autoimmune encephalomyelitis only in female mice. J Immunol (2005) 175:4119–26. doi:10.4049/jimmunol.175.6.4119

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

85. Munger KL, Zhang SM, O’Reilly E, Hernan MA, Olek MJ, Willett WC, et al. Vitamin D intake and incidence of multiple sclerosis. Neurology (2004) 62:60–5. doi:10.1212/01.WNL.0000101723.79681.38

CrossRef Full Text | Google Scholar

86. Munger KL, Levin LI, Hollis BW, Howard NS, Ascherio A. Serum 25-hydroxyvitamin D levels and risk of multiple sclerosis. JAMA (2006) 296:2832–8. doi:10.1001/jama.296.23.2832

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

87. Salzer J, Hallmans G, Nystrom M, Stenlund H, Wadell G, Sundstrom P. Vitamin D as a protective factor in multiple sclerosis. Neurology (2012) 79:2140–5. doi:10.1212/WNL.0b013e3182752ea8

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

88. Simpson S Jr, Taylor B, Blizzard L, Ponsonby AL, Pittas F, Tremlett H, et al. Higher 25-hydroxyvitamin D is associated with lower relapse risk in multiple sclerosis. Ann Neurol (2010) 68:193–203. doi:10.1002/ana.22043

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

89. Soilu-Hanninen M, Airas L, Mononen I, Heikkila A, Viljanen M, Hanninen A. 25-Hydroxyvitamin D levels in serum at the onset of multiple sclerosis. Mult Scler (2005) 11:266–71. doi:10.1191/1352458505ms1157oa

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

90. Mowry EM, Krupp LB, Milazzo M, Chabas D, Strober JB, Belman AL, et al. Vitamin D status is associated with relapse rate in pediatric-onset multiple sclerosis. Ann Neurol (2010) 67:618–24. doi:10.1002/ana.21972

CrossRef Full Text | Google Scholar

91. Banwell B, Bar-Or A, Arnold DL, Sadovnick D, Narayanan S, McGowan M, et al. Clinical, environmental, and genetic determinants of multiple sclerosis in children with acute demyelination: a prospective national cohort study. Lancet Neurol (2011) 10:436–45. doi:10.1016/S1474-4422(11)70045-X

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

92. Runia TF, Hop WC, de Rijke YB, Buljevac D, Hintzen RQ. Lower serum vitamin D levels are associated with a higher relapse risk in multiple sclerosis. Neurology (2012) 79:261–6. doi:10.1212/WNL.0b013e31825fdec7

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

93. Loken-Amsrud KI, Holmoy T, Bakke SJ, Beiske AG, Bjerve KS, Bjornara BT, et al. Vitamin D and disease activity in multiple sclerosis before and during interferon-beta treatment. Neurology (2012) 79:267–73. doi:10.1212/WNL.0b013e31825fdf01

CrossRef Full Text | Google Scholar

94. Ascherio A, Munger KL, White R, Kochert K, Simon KC, Polman CH, et al. Vitamin D as an early predictor of multiple sclerosis activity and progression. JAMA Neurol (2014) 71:306–14. doi:10.1001/jamaneurol.2013.5993

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

95. Becklund BR, Severson KS, Vang SV, DeLuca HF. UV radiation suppresses experimental autoimmune encephalomyelitis independent of vitamin D production. Proc Natl Acad Sci U S A (2010) 107:6418–23. doi:10.1073/pnas.1001119107

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

96. Burton JM, Kimball S, Vieth R, Bar-Or A, Dosch HM, Cheung R, et al. A phase I/II dose-escalation trial of vitamin D3 and calcium in multiple sclerosis. Neurology (2010) 74:1852–9. doi:10.1212/WNL.0b013e3181e1cec2

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

97. Derakhshandi H, Etemadifar M, Feizi A, Abtahi SH, Minagar A, Abtahi MA, et al. Preventive effect of vitamin D3 supplementation on conversion of optic neuritis to clinically definite multiple sclerosis: a double blind, randomized, placebo-controlled pilot clinical trial. Acta Neurol Belg (2013) 113:257–63. doi:10.1007/s13760-012-0166-2

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

98. Stein MS, Liu Y, Gray OM, Baker JE, Kolbe SC, Ditchfield MR, et al. A randomized trial of high-dose vitamin D2 in relapsing-remitting multiple sclerosis. Neurology (2011) 77:1611–8. doi:10.1212/WNL.0b013e3182343274

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

99. Grimaldi L, Barkhof F, Beelke M, Burton J, Holmoy T, Hupperts R, et al. A randomized trial of high-dose vitamin D2 in relapsing-remitting multiple sclerosis. Neurology (2012) 78:841. doi:10.1212/01.wnl.0000413180.13413.ce

CrossRef Full Text | Google Scholar

100. Sigmundsdottir H, Pan J, Debes GF, Alt C, Habtezion A, Soler D, et al. DCs metabolize sunlight-induced vitamin D3 to ‘program’ T cell attraction to the epidermal chemokine CCL27. Nat Immunol (2007) 8:285–93. doi:10.1038/ni1433

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

101. Bjelakovic G, Gluud LL, Nikolova D, Whitfield K, Wetterslev J, Simonetti RG, et al. Vitamin D supplementation for prevention of mortality in adults. Cochrane Database Syst Rev (2011) 1(7):CD007470. doi:10.1002/14651858.CD007470.pub2

CrossRef Full Text | Google Scholar

102. Tripkovic L, Lambert H, Hart K, Smith CP, Bucca G, Penson S, et al. Comparison of vitamin D2 and vitamin D3 supplementation in raising serum 25-hydroxyvitamin D status: a systematic review and meta-analysis. Am J Clin Nutr (2012) 95:1357–64. doi:10.3945/ajcn.111.031070

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

103. Smolders J, Hupperts R, Barkhof F, Grimaldi LM, Holmoy T, Killestein J, et al. Efficacy of vitamin D3 as add-on therapy in patients with relapsing-remitting multiple sclerosis receiving subcutaneous interferon beta-1a: a phase II, multicenter, double-blind, randomized, placebo-controlled trial. J Neurol Sci (2011) 311:44–9. doi:10.1016/j.jns.2011.04.013

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

104. Hypponen E. Vitamin D and increasing incidence of type 1 diabetes-evidence for an association? Diabetes Obes Metab (2010) 12:737–43. doi:10.1111/j.1463-1326.2010.01211.x

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

105. Mathieu C. Vitamin D and diabetes: the devil is in the D-tails. Diabetologia (2010) 53:1545–8. doi:10.1007/s00125-010-1791-z

CrossRef Full Text | Google Scholar

106. Vitamin D supplement in early childhood and risk for type I (insulin-dependent) diabetes mellitus. The EURODIAB substudy 2 study group. Diabetologia (1999) 42:51–4.

Pubmed Abstract | Pubmed Full Text | Google Scholar

107. Zipitis CS, Akobeng AK. Vitamin D supplementation in early childhood and risk of type 1 diabetes: a systematic review and meta-analysis. Arch Dis Child (2008) 93:512–7. doi:10.1136/adc.2007.128579

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

108. Holick MF. Environmental factors that influence the cutaneous production of vitamin D. Am J Clin Nutr (1995) 61:638S–45S.

Pubmed Abstract | Pubmed Full Text | Google Scholar

109. Mohr SB, Garland FC, Garland CF, Gorham ED, Ricordi C. Is there a role of vitamin D deficiency in type 1 diabetes of children? Am J Prev Med (2010) 39:189–90. doi:10.1016/j.amepre.2010.03.023

CrossRef Full Text | Google Scholar

110. Hypponen E, Laara E, Reunanen A, Jarvelin MR, Virtanen SM. Intake of vitamin D and risk of type 1 diabetes: a birth-cohort study. Lancet (2001) 358:1500–3. doi:10.1016/S0140-6736(01)06580-1

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

111. Takiishi T, Ding L, Baeke F, Spagnuolo I, Sebastiani G, Laureys J, et al. Dietary supplementation with high doses of regular vitamin D3 safely reduces diabetes incidence in NOD mice when given early and long term. Diabetes (2014) 63:2026–36. doi:10.2337/db13-1559

CrossRef Full Text | Google Scholar

112. Gedik O, Akalin S. Effects of vitamin D deficiency and repletion on insulin and glucagon secretion in man. Diabetologia (1986) 29:142–5. doi:10.1007/BF02427083

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

113. Aljabri KS, Bokhari SA, Khan MJ. Glycemic changes after vitamin D supplementation in patients with type 1 diabetes mellitus and vitamin D deficiency. Ann Saudi Med (2010) 30:454–8. doi:10.4103/0256-4947.72265

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

114. Nwosu BU, Maranda L. The effects of vitamin D supplementation on hepatic dysfunction, vitamin D status, and glycemic control in children and adolescents with vitamin D deficiency and either type 1 or type 2 diabetes mellitus. PLoS One (2014) 9:e99646. doi:10.1371/journal.pone.0099646

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

115. Hill AB. The environment and disease: association or causation? Proc R Soc Med (1965) 58:295–300.

Google Scholar

116. Wucherpfennig KW, Sethi D. T cell receptor recognition of self and foreign antigens in the induction of autoimmunity. Semin Immunol (2011) 23:84–91. doi:10.1016/j.smim.2011.01.007

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

117. Baxter AG. The origin and application of experimental autoimmune encephalomyelitis. Nat Rev Immunol (2007) 7:904–12. doi:10.1038/nri2190

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

118. Sibley RK, Sutherland DE, Goetz F, Michael AF. Recurrent diabetes mellitus in the pancreas iso- and allograft. A light and electron microscopic and immunohistochemical analysis of four cases. Lab Invest (1985) 53:132–44.

Pubmed Abstract | Pubmed Full Text | Google Scholar

119. Boitard C. T-lymphocyte recognition of beta cells in type 1 diabetes: clinical perspectives. Diabetes Metab (2013) 39:459–66. doi:10.1016/j.diabet.2013.08.001

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

120. Hafler DA, Weiner HL. T cells in multiple sclerosis and inflammatory central nervous system diseases. Immunol Rev (1987) 100:307–32. doi:10.1111/j.1600-065X.1987.tb00537.x

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

121. Wucherpfennig KW, Hafler DA. A review of T-cell receptors in multiple sclerosis: clonal expansion and persistence of human T-cells specific for an immunodominant myelin basic protein peptide. Ann N Y Acad Sci (1995) 756:241–58. doi:10.1111/j.1749-6632.1995.tb44522.x

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

122. Severson C, Hafler DA. T-cells in multiple sclerosis. Results Probl Cell Differ (2010) 51:75–98. doi:10.1007/400_2009_9012

CrossRef Full Text | Google Scholar

123. Gandhi KS, McKay FC, Cox M, Riveros C, Armstrong N, Heard RN, et al. The multiple sclerosis whole blood mRNA transcriptome and genetic associations indicate dysregulation of specific T cell pathways in pathogenesis. Hum Mol Genet (2010) 19:2134–43. doi:10.1093/hmg/ddq090

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

124. Ottoboni L, Keenan BT, Tamayo P, Kuchroo M, Mesirov JP, Buckle GJ, et al. An RNA profile identifies two subsets of multiple sclerosis patients differing in disease activity. Sci Transl Med (2012) 4:153ra131. doi:10.1126/scitranslmed.3004186

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

125. Raj T, Rothamel K, Mostafavi S, Ye C, Lee MN, Replogle JM, et al. Polarization of the effects of autoimmune and neurodegenerative risk alleles in leukocytes. Science (2014) 344:519–23. doi:10.1126/science.1249547

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

126. Ye CJ, Feng T, Kwon HK, Raj T, Wilson MT, Asinovski N, et al. Intersection of population variation and autoimmunity genetics in human T cell activation. Science (2014) 345:1254665. doi:10.1126/science.1254665

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

127. Haussler MR, Whitfield GK, Kaneko I, Haussler CA, Hsieh D, Hsieh JC, et al. Molecular mechanisms of vitamin D action. Calcif Tissue Int (2013) 92:77–98. doi:10.1007/s00223-012-9619-0

CrossRef Full Text | Google Scholar

128. Holick MF. Vitamin D: a millennium perspective. J Cell Biochem (2003) 88:296–307. doi:10.1002/jcb.10338

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

129. Whitfield GK, Dang HT, Schluter SF, Bernstein RM, Bunag T, Manzon LA, et al. Cloning of a functional vitamin D receptor from the lamprey (Petromyzon marinus), an ancient vertebrate lacking a calcified skeleton and teeth. Endocrinology (2003) 144:2704–16. doi:10.1210/en.2002-221101

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

130. Nystad AE, Wergeland S, Aksnes L, Myhr KM, Bo L, Torkildsen O. Effect of high-dose 1.25 dihydroxyvitamin D on remyelination in the cuprizone model. APMIS (2014) 122(12):1178–86. doi:10.1111/apm.12281

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

131. Wolden-Kirk H, Overbergh L, Christesen HT, Brusgaard K, Mathieu C. Vitamin D and diabetes: its importance for beta cell and immune function. Mol Cell Endocrinol (2011) 347:106–20. doi:10.1016/j.mce.2011.08.016

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

132. Wang TT, Dabbas B, Laperriere D, Bitton AJ, Soualhine H, Tavera-Mendoza LE, et al. Direct and indirect induction by 1,25-dihydroxyvitamin D3 of the NOD2/CARD15-defensin beta2 innate immune pathway defective in Crohn disease. J Biol Chem (2010) 285:2227–31. doi:10.1074/jbc.C109.071225

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

133. Prosser DE, Jones G. Enzymes involved in the activation and inactivation of vitamin D. Trends Biochem Sci (2004) 29:664–73. doi:10.1016/j.tibs.2004.10.005

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

134. Jones G, Prosser DE, Kaufmann M. 25-Hydroxyvitamin D-24-hydroxylase (CYP24A1): its important role in the degradation of vitamin D. Arch Biochem Biophys (2012) 523:9–18. doi:10.1016/j.abb.2011.11.003

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

135. Hollis BW, Wagner CL. The role of the parent compound vitamin D with respect to metabolism and function: why clinical dose intervals can affect clinical outcomes. J Clin Endocrinol Metab (2013) 98(12):4619–28. doi:10.1210/jc.2013-2653

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

136. Adams JS, Rafison B, Witzel S, Reyes RE, Shieh A, Chun R, et al. Regulation of the extrarenal CYP27B1-hydroxylase. J Steroid Biochem Mol Biol (2014) 144 Pt A:22–7. doi:10.1016/j.jsbmb.2013.12.009

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

137. Hine TJ, Roberts NB. Seasonal variation in serum 25-hydroxy vitamin D3 does not affect 1,25-dihydroxy vitamin D. Ann Clin Biochem (1994) 31(Pt 1):31–4. doi:10.1177/000456329403100105

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

138. Webb AR, Kline L, Holick MF. Influence of season and latitude on the cutaneous synthesis of vitamin D3: exposure to winter sunlight in Boston and Edmonton will not promote vitamin D3 synthesis in human skin. J Clin Endocrinol Metab (1988) 67:373–78. doi:10.1210/jcem-67-2-373

CrossRef Full Text | Google Scholar

139. Holick MF, Chen TC, Lu Z, Sauter E. Vitamin D and skin physiology: a D-lightful story. J Bone Miner Res (2007) 22(Suppl 2):V28–33. doi:10.1359/jbmr.07s211

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

140. Hossein-nezhad A, Holick MF. Vitamin D for health: a global perspective. Mayo Clin Proc (2013) 88:720–55. doi:10.1016/j.mayocp.2013.05.011

CrossRef Full Text | Google Scholar

141. Bener A, Al-Ali M, Hoffmann GF. High prevalence of vitamin D deficiency in young children in a highly sunny humid country: a global health problem. Minerva Pediatr (2009) 61:15–22.

Pubmed Abstract | Pubmed Full Text | Google Scholar

142. Adams JS, Hewison M. Update in vitamin D. J Clin Endocrinol Metab (2010) 95:471–8. doi:10.1210/jc.2009-1773

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

143. Grant WB, Peiris AN. Possible role of serum 25-hydroxyvitamin D in black-white health disparities in the United States. J Am Med Dir Assoc (2010) 11:617–28. doi:10.1016/j.jamda.2010.03.013

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

144. Ahn J, Yu K, Stolzenberg-Solomon R, Simon KC, McCullough ML, Gallicchio L, et al. Genome-wide association study of circulating vitamin D levels. Hum Mol Genet (2010) 19:2739–45. doi:10.1093/hmg/ddq155

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

145. Wang TJ, Zhang F, Richards JB, Kestenbaum B, van Meurs JB, Berry D, et al. Common genetic determinants of vitamin D insufficiency: a genome-wide association study. Lancet (2010) 376:180–8. doi:10.1016/S0140-6736(10)60588-0

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

146. Powe CE, Evans MK, Wenger J, Zonderman AB, Berg AH, Nalls M, et al. Vitamin D-binding protein and vitamin D status of black Americans and white Americans. N Engl J Med (2013) 369:1991–2000. doi:10.1056/NEJMoa1306357

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

147. Speeckaert M, Huang G, Delanghe JR, Taes YE. Biological and clinical aspects of the vitamin D binding protein (Gc-globulin) and its polymorphism. Clin Chim Acta (2006) 372:33–42. doi:10.1016/j.cca.2006.03.011

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

148. Batai K, Murphy AB, Shah E, Ruden M, Newsome J, Agate S, et al. Common vitamin D pathway gene variants reveal contrasting effects on serum vitamin D levels in African Americans and European Americans. Hum Genet (2014) 133:1395–405. doi:10.1007/s00439-014-1472-y

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

149. Murphy AB, Kelley B, Nyame YA I, Martin K, Smith DJ, Castaneda L, et al. Predictors of serum vitamin D levels in African American and European American men in Chicago. Am J Mens Health (2012) 6:420–6. doi:10.1177/1557988312437240

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

150. Hewison M, Zehnder D, Bland R, Stewart PM. 1alpha-Hydroxylase and the action of vitamin D. J Mol Endocrinol (2000) 25:141–8. doi:10.1677/jme.0.0250141

CrossRef Full Text | Google Scholar

151. Adams JS, Gacad MA. Characterization of 1 alpha-hydroxylation of vitamin D3 sterols by cultured alveolar macrophages from patients with sarcoidosis. J Exp Med (1985) 161:755–65. doi:10.1084/jem.161.4.755

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

152. Bhalla AK, Amento EP, Clemens TL, Holick MF, Krane SM. Specific high-affinity receptors for 1,25-dihydroxyvitamin D3 in human peripheral blood mononuclear cells: presence in monocytes and induction in T lymphocytes following activation. J Clin Endocrinol Metab (1983) 57:1308–10. doi:10.1210/jcem-57-6-1308

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

153. Provvedini DM, Tsoukas CD, Deftos LJ, Manolagas SC. 1,25-dihydroxyvitamin D3 receptors in human leukocytes. Science (1983) 221:1181–3. doi:10.1126/science.6310748

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

154. Jeffery LE, Wood AM, Qureshi OS, Hou TZ, Gardner D, Briggs Z, et al. Availability of 25-hydroxyvitamin D(3) to APCs controls the balance between regulatory and inflammatory T cell responses. J Immunol (2012) 189:5155–64. doi:10.4049/jimmunol.1200786

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

155. Stoffels K, Overbergh L, Giulietti A, Verlinden L, Bouillon R, Mathieu C. Immune regulation of 25-hydroxyvitamin-D3-1alpha-hydroxylase in human monocytes. J Bone Miner Res (2006) 21:37–47. doi:10.1359/JBMR.050908

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

156. Krutzik SR, Hewison M, Liu PT, Robles JA, Stenger S, Adams JS, et al. IL-15 links TLR2/1-induced macrophage differentiation to the vitamin D-dependent antimicrobial pathway. J Immunol (2008) 181:7115–20. doi:10.4049/jimmunol.181.10.7115

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

157. Hewison M. Extrarenal CYP27B1 and vitamin D physiology. In: Gombart AF, editor. Vitamin D. Oxidative Stress, Immunity, and Aging. Boca Raton, FL: CRC Press, Taylor & Francis Group (2013). p. 99–127.

Google Scholar

158. Fu GK, Lin D, Zhang MY, Bikle DD, Shackleton CH, Miller WL, et al. Cloning of human 25-hydroxyvitamin D-1 alpha-hydroxylase and mutations causing vitamin D-dependent rickets type 1. Mol Endocrinol (1997) 11:1961–70. doi:10.1210/mend.11.13.0035

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

159. Bikle DD, Nemanic MK, Gee E, Elias P. 1,25-Dihydroxyvitamin D3 production by human keratinocytes. Kinetics and regulation. J Clin Invest (1986) 78:557–66. doi:10.1172/JCI112609

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

160. Hansdottir S, Monick MM, Hinde SL, Lovan N, Look DC, Hunninghake GW. Respiratory epithelial cells convert inactive vitamin D to its active form: potential effects on host defense. J Immunol (2008) 181:7090–9. doi:10.4049/jimmunol.181.10.7090

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

161. Liu N, Nguyen L, Chun RF, Lagishetty V, Ren S, Wu S, et al. Altered endocrine and autocrine metabolism of vitamin D in a mouse model of gastrointestinal inflammation. Endocrinology (2008) 149:4799–808. doi:10.1210/en.2008-0060

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

162. Cross HS, Nittke T, Kallay E. Colonic vitamin D metabolism: implications for the pathogenesis of inflammatory bowel disease and colorectal cancer. Mol Cell Endocrinol (2011) 347(1–2):70–9. doi:10.1016/j.mce.2011.07.022

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

163. Anderson PH, Hendrix I, Sawyer RK, Zarrinkalam R, Manvis J, Sarvestani GT, et al. Co-expression of C’YP27B1 enzyme with the 1.5 kb CYP27B1 promoter-luciferase transgene in the mouse. Mol Cell Endocrinol (2008) 285:1–9. doi:10.1016/j.mce.2007.12.018

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

164. O’Brien KO, Li S, Cao C, Kent T, Young BV, Queenan RA, et al. Placental CYP27B1 and CYP24A1 expression in human placental tissue and their association with maternal and neonatal calciotropic hormones. J Clin Endocrinol Metab (2014) 99:1348–56. doi:10.1210/jc.2013-1366

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

165. Overbergh L, Decallonne B, Valckx D, Verstuyf A, Depovere J, Laureys J, et al. Identification and immune regulation of 25-hydroxyvitamin D-1-alpha-hydroxylase in murine macrophages. Clin Exp Immunol (2000) 120:139–46. doi:10.1046/j.1365-2249.2000.01204.x

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

166. Bell NH, Stern PH, Pantzer E, Sinha TK, DeLuca HF. Evidence that increased circulating 1 alpha, 25-dihydroxyvitamin D is the probable cause for abnormal calcium metabolism in sarcoidosis. J Clin Invest (1979) 64:218–25. doi:10.1172/JCI109442

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

167. Zerwekh JE, Pak CY, Kaplan RA, McGuire JL, Upchurch K, Breslau N, et al. Pathogenetic role of 1 alpha,25-dihydroxyvitamin D in sarcoidosis and absorptive hypercalciuria: different response to prednisolone therapy. J Clin Endocrinol Metab (1980) 51:381–6. doi:10.1210/jcem-51-2-381

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

168. Castro M, King TS, Kunselman SJ, Cabana MD, Denlinger L, Holguin F, et al. Effect of vitamin D3 on asthma treatment failures in adults with symptomatic asthma and lower vitamin D levels: the VIDA randomized clinical trial. JAMA (2014) 311:2083–91. doi:10.1001/jama.2014.5052

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

169. Novakovic B, Sibson M, Ng HK, Manuelpillai U, Rakyan V, Down T, et al. Placenta-specific methylation of the vitamin D 24-hydroxylase gene: implications for feedback autoregulation of active vitamin D levels at the fetomaternal interface. J Biol Chem (2009) 284:14838–48. doi:10.1074/jbc.M809542200

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

170. Luo W, Karpf AR, Deeb KK, Muindi JR, Morrison CD, Johnson CS, et al. Epigenetic regulation of vitamin D 24-hydroxylase/CYP24A1 in human prostate cancer. Cancer Res (2010) 70:5953–62. doi:10.1158/0008-5472.CAN-10-0617

CrossRef Full Text | Google Scholar

171. Spanier JA, Nashold FE, Olson JK, Hayes CE. The ifng gene is essential for vdr gene expression and vitamin d3-mediated reduction of the pathogenic T cell burden in the central nervous system in experimental autoimmune encephalomyelitis, a multiple sclerosis model. J Immunol (2012) 189:3188–97. doi:10.4049/jimmunol.1102925

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

172. Correale J, Ysrraelit MC, Gaitan MI. Gender differences in 1,25 dihydroxyvitamin D3 immunomodulatory effects in multiple sclerosis patients and healthy subjects. J Immunol (2010) 185:4948–58. doi:10.4049/jimmunol.1000588

CrossRef Full Text | Google Scholar

173. Orton SM, Wald L, Confavreux C, Vukusic S, Krohn JP, Ramagopalan SV, et al. Association of UV radiation with multiple sclerosis prevalence and sex ratio in France. Neurology (2011) 76:425–31. doi:10.1212/WNL.0b013e31820a0a9f

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

174. Carlberg C. Genome-wide (over)view on the actions of vitamin D. Front Physiol (2014) 5:167. doi:10.3389/fphys.2014.00167

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

175. Carlberg C, Campbell MJ. Vitamin D receptor signaling mechanisms: integrated actions of a well-defined transcription factor. Steroids (2013) 78:127–36. doi:10.1016/j.steroids.2012.10.019

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

176. Wang Y, Becklund BR, DeLuca HF. Identification of a highly specific and versatile vitamin D receptor antibody. Arch Biochem Biophys (2010) 494:166–77. doi:10.1016/j.abb.2009.11.029

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

177. Wang Y, Zhu J, DeLuca HF. Where is the vitamin D receptor? Arch Biochem Biophys (2012) 523:123–33. doi:10.1016/j.abb.2012.04.001

CrossRef Full Text | Google Scholar

178. Brennan A, Katz DR, Nunn JD, Barker S, Hewison M, Fraher LJ, et al. Dendritic cells from human tissues express receptors for the immunoregulatory vitamin D3 metabolite, dihydroxycholecalciferol. Immunology (1987) 61:457–61.

Pubmed Abstract | Pubmed Full Text | Google Scholar

179. Baeke F, Korf H, Overbergh L, van Etten E, Verstuyf A, Gysemans C, et al. Human T lymphocytes are direct targets of 1,25-dihydroxyvitamin D3 in the immune system. J Steroid Biochem Mol Biol (2010) 121:221–7. doi:10.1016/j.jsbmb.2010.03.037

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

180. Veldman CM, Cantorna MT, DeLuca HF. Expression of 1,25-dihydroxyvitamin D(3) receptor in the immune system. Arch Biochem Biophys (2000) 374:334–8. doi:10.1006/abbi.1999.1605

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

181. von Essen MR, Kongsbak M, Schjerling P, Olgaard K, Odum N, Geisler C. Vitamin D controls T cell antigen receptor signaling and activation of human T cells. Nat Immunol (2010) 11:344–9. doi:10.1038/ni.1851

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

182. Nashold FE, Hoag KA, Goverman J, Hayes CE. Rag-1-dependent cells are necessary for 1,25-dihydroxyvitamin D(3) prevention of experimental autoimmune encephalomyelitis. J Neuroimmunol (2001) 119:16–29. doi:10.1016/S0165-5728(01)00360-5

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

183. Chang SH, Chung Y, Dong C. Vitamin D suppresses Th17 cytokine production by inducing C/EBP homologous protein (CHOP) expression. J Biol Chem (2010) 285:38751–5. doi:10.1074/jbc.C110.185777

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

184. Mayne CG, Spanier JA, Relland LM, Williams CB, Hayes CE. 1,25-Dihydroxyvitamin D3 acts directly on the T lymphocyte vitamin D receptor to inhibit experimental autoimmune encephalomyelitis. Eur J Immunol (2011) 41:822–32. doi:10.1002/eji.201040632

CrossRef Full Text | Google Scholar

185. Joshi S, Pantalena LC, Liu XK, Gaffen SL, Liu H, Rohowsky-Kochan C, et al. 1,25-dihydroxyvitamin D(3) ameliorates Th17 autoimmunity via transcriptional modulation of interleukin-17A. Mol Cell Biol (2011) 31:3653–69. doi:10.1128/MCB.05020-11

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

186. Nashold FE, Spach KM, Spanier JA, Hayes CE. Estrogen controls vitamin D3-mediated resistance to experimental autoimmune encephalomyelitis by controlling vitamin D3 metabolism and receptor expression. J Immunol (2009) 183:3672–81. doi:10.4049/jimmunol.0901351

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

187. Chang JH, Cha HR, Lee DS, Seo KY, Kweon MN. 1,25-Dihydroxyvitamin D3 inhibits the differentiation and migration of T(H)17 cells to protect against experimental autoimmune encephalomyelitis. PLoS One (2010) 5:e12925. doi:10.1371/journal.pone.0012925

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

188. Guo C, Gombart AF. The antibiotic effects of vitamin D. Endocr Metab Immune Disord Drug Targets (2014) 14:255–66. doi:10.2174/1871530314666140709085159

CrossRef Full Text | Google Scholar

189. Tiosano D, Wildbaum G, Gepstein V, Verbitsky O, Weisman Y, Karin N, et al. The role of vitamin D receptor in innate and adaptive immunity: a study in hereditary vitamin D-resistant rickets patients. J Clin Endocrinol Metab (2013) 98:1685–93. doi:10.1210/jc.2012-3858

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

190. Klein L, Kyewski B, Allen PM, Hogquist KA. Positive and negative selection of the T cell repertoire: what thymocytes see (and don’t see). Nat Rev Immunol (2014) 14:377–91. doi:10.1038/nri3667

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

191. Wucherpfennig KW, Call MJ, Deng L, Mariuzza R. Structural alterations in peptide-MHC recognition by self-reactive T cell receptors. Curr Opin Immunol (2009) 21:590–5. doi:10.1016/j.coi.2009.07.008

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

192. Lang HL, Jacobsen H, Ikemizu S, Andersson C, Harlos K, Madsen L, et al. A functional and structural basis for TCR cross-reactivity in multiple sclerosis. Nat Immunol (2002) 3:940–3. doi:10.1038/ni835

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

193. Disanto G, Watson CT, Meier UC, Ebers GC, Giovannoni G, Ramagopalan SV. Month of birth and thymic output. JAMA Neurol (2013) 70:527–8. doi:10.1001/jamaneurol.2013.2116

CrossRef Full Text | Google Scholar

194. Sewell AK. Why must T cells be cross-reactive? Nat Rev Immunol (2012) 12:669–77. doi:10.1038/nri3279

CrossRef Full Text | Google Scholar

195. Birnbaum ME, Mendoza JL, Sethi DK, Dong S, Glanville J, Dobbins J, et al. Deconstructing the peptide-MHC specificity of T cell recognition. Cell (2014) 157:1073–87. doi:10.1016/j.cell.2014.03.047

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

196. Marrack P, Scott-Browne J, MacLeod MK. Terminating the immune response. Immunol Rev (2010) 236:5–10. doi:10.1111/j.1600-065X.2010.00928.x

CrossRef Full Text | Google Scholar

197. Vignali DA, Collison LW, Workman CJ. How regulatory T cells work. Nat Rev Immunol (2008) 8:523–32. doi:10.1038/nri2343

CrossRef Full Text | Google Scholar

198. Robinson AP, Harp CT, Noronha A, Miller SD. The experimental autoimmune encephalomyelitis (EAE) model of MS: utility for understanding disease pathophysiology and treatment. Handbook Clin Neurol (2014) 122:173–89. doi:10.1016/B978-0-444-52001-2.00008-X

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

199. Ben-Nun A, Kaushansky N, Kawakami N, Krishnamoorthy G, Berer K, Liblau R, et al. From classic to spontaneous and humanized models of multiple sclerosis: impact on understanding pathogenesis and drug development. J Autoimmun (2014) 54:33–50. doi:10.1016/j.jaut.2014.06.004

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

200. Driver JP, Serreze DV, Chen YG. Mouse models for the study of autoimmune type 1 diabetes: a NOD to similarities and differences to human disease. Semin Immunopathol (2011) 33:67–87. doi:10.1007/s00281-010-0204-1

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

201. Mathieu C, Waer M, Laureys J, Rutgeerts O, Bouillon R. Prevention of autoimmune diabetes in NOD mice by 1,25 dihydroxyvitamin D3. Diabetologia (1994) 37:552–8. doi:10.1007/BF00403372

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

202. Branisteanu DD, Waer M, Sobis H, Marcelis S, Vandeputte M, Bouillon R. Prevention of murine experimental allergic encephalomyelitis: cooperative effects of cyclosporine and 1 alpha, 25-(OH)2D3. J Neuroimmunol (1995) 61:151–60. doi:10.1016/0165-5728(95)00076-E

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

203. Cantorna MT, Hayes CE, DeLuca HF. 1,25-dihydroxyvitamin D3 reversibly blocks the progression of relapsing encephalomyelitis, a model of multiple sclerosis. Proc Natl Acad Sci U S A (1996) 93:7861–4. doi:10.1073/pnas.93.15.7861

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

204. Cantorna MT, Woodward WD, Hayes CE, DeLuca HF. 1,25-dihydroxyvitamin D3 is a positive regulator for the two anti-encephalitogenic cytokines TGF-beta 1 and IL-4. J Immunol (1998) 160:5314–9.

Google Scholar

205. Spach KM, Nashold FE, Dittel BN, Hayes CE. IL-10 signaling is essential for 1,25-dihydroxyvitamin D3-mediated inhibition of experimental autoimmune encephalomyelitis. J Immunol (2006) 177:6030–7. doi:10.4049/jimmunol.177.9.6030

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

206. Mattner F, Smiroldo S, Galbiati F, Muller M, Di Lucia P, Poliani PL, et al. Inhibition of Th1 development and treatment of chronic-relapsing experimental allergic encephalomyelitis by a non-hypercalcemic analogue of 1,25-dihydroxyvitamin D(3). Eur J Immunol (2000) 30:498–508. doi:10.1002/1521-4141(200002)30:2<498::AID-IMMU498>3.3.CO;2-H

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

207. Nataf S, Garcion E, Darcy F, Chabannes D, Muller JY, Brachet P. 1,25 Dihydroxyvitamin D3 exerts regional effects in the central nervous system during experimental allergic encephalomyelitis. J Neuropathol Exp Neurol (1996) 55:904–14. doi:10.1097/00005072-199608000-00006

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

208. Nashold FE, Miller DJ, Hayes CE. 1,25-dihydroxyvitamin D3 treatment decreases macrophage accumulation in the CNS of mice with experimental autoimmune encephalomyelitis. J Neuroimmunol (2000) 103:171–9. doi:10.1016/S0165-5728(99)00247-7

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

209. Spach KM, Pedersen LB, Nashold FE, Kayo T, Yandell BS, Prolla TA, et al. Gene expression analysis suggests that 1,25-dihydroxyvitamin D3 reverses experimental autoimmune encephalomyelitis by stimulating inflammatory cell apoptosis. Physiol Genomics (2004) 18:141–51. doi:10.1152/physiolgenomics.00003.2004

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

210. Pedersen LB, Nashold FE, Spach KM, Hayes CE. 1,25-dihydroxyvitamin D3 reverses experimental autoimmune encephalomyelitis by inhibiting chemokine synthesis and monocyte trafficking. J Neurosci Res (2007) 85:2480–90. doi:10.1002/jnr.21382

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

211. Muller K, Gram J, Bollerslev J, Diamant M, Barington T, Hansen MB, et al. Down-regulation of monocyte functions by treatment of healthy adults with 1 alpha,25 dihydroxyvitamin D3. Int J Immunopharmacol (1991) 13:525–30. doi:10.1016/0192-0561(91)90072-F

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

212. Mahon BD, Gordon SA, Cruz J, Cosman F, Cantorna MT. Cytokine profile in patients with multiple sclerosis following vitamin D supplementation. J Neuroimmunol (2003) 134:128–32. doi:10.1016/S0165-5728(02)00396-X

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

213. Drozdenko G, Heine G, Worm M. Oral vitamin D increases the frequencies of CD38+ human B cells and ameliorates IL-17-producing T cells. Exp Dermatol (2014) 23:107–12. doi:10.1111/exd.12300

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

214. Tang J, Zhou R, Luger D, Zhu W, Silver PB, Grajewski RS, et al. Calcitriol suppresses antiretinal autoimmunity through inhibitory effects on the Th17 effector response. J Immunol (2009) 182:4624–32. doi:10.4049/jimmunol.0801543

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

215. Bruce D, Yu S, Ooi JH, Cantorna MT. Converging pathways lead to overproduction of IL-17 in the absence of vitamin D signaling. Int Immunol (2011) 23:519–28. doi:10.1093/intimm/dxr045

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

216. Nashold FE, Nelson CD, Brown LM, Hayes CE. One calcitriol dose transiently increases Helios+FoxP3+ T cells and ameliorates autoimmune demyelinating disease. J Neuroimmunol (2013) 263:64–74. doi:10.1016/j.jneuroim.2013.07.016

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

217. Kang SW, Kim SH, Lee N, Lee WW, Hwang KA, Shin MS, et al. 1,25-Dihyroxyvitamin D3 promotes FOXP3 expression via binding to vitamin D response elements in its conserved noncoding sequence region. J Immunol (2012) 188:5276–82. doi:10.4049/jimmunol.1101211

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

218. Jeffery LE, Burke F, Mura M, Zheng Y, Qureshi OS, Hewison M, et al. 1,25-dihydroxyvitamin D3 and IL-2 combine to inhibit T cell production of inflammatory cytokines and promote development of regulatory T cells expressing CTLA-4 and FoxP3. J Immunol (2009) 183:5458–67. doi:10.4049/jimmunol.0803217

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

219. Kickler K, Ni Choileain S, Williams A, Richards A, Astier AL. Calcitriol modulates the CD46 pathway in T cells. PLoS One (2012) 7:e48486. doi:10.1371/journal.pone.0048486

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

220. Van Belle TL, Vanherwegen AS, Feyaerts D, De Clercq P, Verstuyf A, Korf H, et al. 1,25-dihydroxyvitamin D3 and its analog TX527 promote a stable regulatory T cell phenotype in T cells from type 1 diabetes patients. PLoS One (2014) 9:e109194. doi:10.1371/journal.pone.0109194

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

221. Correale J, Ysrraelit MC, Gaitan MI. Immunomodulatory effects of Vitamin D in multiple sclerosis. Brain (2009) 132:1146–60. doi:10.1093/brain/awp033

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

222. Smolders J, Thewissen M, Peelen E, Menheere P, Cohen Tervaert JW, Damoiseaux J, et al. Vitamin D status is positively correlated with regulatory T cell function in patients with multiple sclerosis. PLoS One (2009) 4:e6635. doi:10.1371/journal.pone.0006635

CrossRef Full Text | Google Scholar

223. Smolders J, Peelen E, Thewissen M, Cohen Tervaert JW, Menheere P, Hupperts R, et al. Safety and T cell modulating effects of high dose vitamin D3 supplementation in multiple sclerosis. PLoS One (2010) 5:e15235. doi:10.1371/journal.pone.0015235

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

224. Cantorna MT, Humpal-Winter J, DeLuca HF. In vivo upregulation of interleukin-4 is one mechanism underlying the immunoregulatory effects of 1,25-dihydroxyvitamin D(3). Arch Biochem Biophys (2000) 377:135–8. doi:10.1006/abbi.2000.1765

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

225. Keating P, Munim A, Hartmann JX. Effect of vitamin D on T-helper type 9 polarized human memory cells in chronic persistent asthma. Ann Allergy Asthma Immunol (2014) 112:154–62. doi:10.1016/j.anai.2013.11.015

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

226. Wang F, Wan L, Zhang C, Zheng X, Li J, Chen ZK. Tim-3-Galectin-9 pathway involves the suppression induced by CD4+CD25+ regulatory T cells. Immunobiology (2009) 214:342–9. doi:10.1016/j.imbio.2008.10.007

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

227. Flugel A, Schwaiger FW, Neumann H, Medana I, Willem M, Wekerle H, et al. Neuronal FasL induces cell death of encephalitogenic T lymphocytes. Brain Pathol (2000) 10:353–64. doi:10.1111/j.1750-3639.2000.tb00267.x

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

228. Kohji T, Matsumoto Y. Coexpression of Fas/FasL and Bax on brain and infiltrating T cells in the central nervous system is closely associated with apoptotic cell death during autoimmune encephalomyelitis. J Neuroimmunol (2000) 106:165–71. doi:10.1016/S0165-5728(00)00238-1

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

229. Colucci F, Bergman ML, Penha-Goncalves C, Cilio CM, Holmberg D. Apoptosis resistance of nonobese diabetic peripheral lymphocytes linked to the Idd5 diabetes susceptibility region. Proc Natl Acad Sci U S A (1997) 94:8670–4. doi:10.1073/pnas.94.16.8670

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

230. Comi C, Fleetwood T, Dianzani U. The role of T cell apoptosis in nervous system autoimmunity. Autoimmun Rev (2012) 12:150–6. doi:10.1016/j.autrev.2011.08.011

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

231. Bieganowska KD, Ausubel LJ, Modabber Y, Slovik E, Messersmith W, Hafler DA. Direct ex vivo analysis of activated, Fas-sensitive autoreactive T cells in human autoimmune disease. J Exp Med (1997) 185:1585–94. doi:10.1084/jem.185.9.1585

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

232. Saresella M, Marventano I, Speciale L, Ruzzante S, Trabattoni D, Della Bella S, et al. Programmed cell death of myelin basic protein-specific T lymphocytes is reduced in patients with acute multiple sclerosis. J Neuroimmunol (2005) 166:173–9. doi:10.1016/j.jneuroim.2005.05.010

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

233. Kohji T, Tanuma N, Aikawa Y, Kawazoe Y, Suzuki Y, Kohyama K, et al. Interaction between apoptotic cells and reactive brain cells in the central nervous system of rats with autoimmune encephalomyelitis. J Neuroimmunol (1998) 82:168–74. doi:10.1016/S0165-5728(97)00198-7

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

234. Bechmann I, Mor G, Nilsen J, Eliza M, Nitsch R, Naftolin F. FasL (CD95L, Apo1L) is expressed in the normal rat and human brain: evidence for the existence of an immunological brain barrier. Glia (1999) 27:62–74. doi:10.1002/(SICI)1098-1136(199907)27:1<62::AID-GLIA7>3.0.CO;2-S

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

235. Bechmann I, Steiner B, Gimsa U, Mor G, Wolf S, Beyer M, et al. Astrocyte-induced T cell elimination is CD95 ligand dependent. J Neuroimmunol (2002) 132:60–5. doi:10.1016/S0165-5728(02)00311-9

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

236. Decallonne B, Mathieu C. Defective activation-induced cell death in NOD T lymphocytes: 1,25-dihydroxyvitamin D3 restores defect. Ann N Y Acad Sci (2003) 1005:176–7. doi:10.1196/annals.1288.021

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

237. Decallonne B, van Etten E, Giulietti A, Casteels K, Overbergh L, Bouillon R, et al. Defect in activation-induced cell death in non-obese diabetic (NOD) T lymphocytes. J Autoimmun (2003) 20:219–26. doi:10.1016/S0896-8411(03)00025-8

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

238. Decallonne B, van Etten E, Overbergh L, Valckx D, Bouillon R, Mathieu C. 1Alpha,25-dihydroxyvitamin D3 restores thymocyte apoptosis sensitivity in non-obese diabetic (NOD) mice through dendritic cells. J Autoimmun (2005) 24:281–9. doi:10.1016/j.jaut.2005.03.007

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

239. Zhu C, Anderson AC, Schubart A, Xiong H, Imitola J, Khoury SJ, et al. The Tim-3 ligand galectin-9 negatively regulates T helper type 1 immunity. Nat Immunol (2005) 6:1245–52. doi:10.1038/ni1271

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

240. Toscano MA, Bianco GA, Ilarregui JM, Croci DO, Correale J, Hernandez JD, et al. Differential glycosylation of T(H)1, T(H)2 and T(H)-17 effector cells selectively regulates susceptibility to cell death. Nat Immunol (2007) 8(8):825–34. doi:10.1038/ni1482

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

241. Mkhikian H, Grigorian A, Li CF, Chen HL, Newton B, Zhou RW, et al. Genetics and the environment converge to dysregulate N-glycosylation in multiple sclerosis. Nat Commun (2011) 2:334. doi:10.1038/ncomms1333

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

242. Wing K, Sakaguchi S. Regulatory T cells exert checks and balances on self tolerance and autoimmunity. Nat Immunol (2010) 11:7–13. doi:10.1038/ni.1818

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

243. Swierkosz JE, Swanborg RH. Suppressor cell control of unresponsiveness to experimental allergic encephalomyelitis. J Immunol (1975) 115:631–3.

Pubmed Abstract | Pubmed Full Text | Google Scholar

244. Karpus WJ, Swanborg RH. CD4+ suppressor cells differentially affect the production of IFN-gamma by effector cells of experimental autoimmune encephalomyelitis. J Immunol (1989) 143:3492–7.

Pubmed Abstract | Pubmed Full Text | Google Scholar

245. Sakaguchi S, Sakaguchi N, Asano M, Itoh M, Toda M. Immunologic self-tolerance maintained by activated T cells expressing IL-2 receptor alpha-chains (CD25). Breakdown of a single mechanism of self-tolerance causes various autoimmune diseases. J Immunol (1995) 155:1151–64.

Google Scholar

246. Brunkow ME, Jeffery EW, Hjerrild KA, Paeper B, Clark LB, Yasayko SA, et al. Disruption of a new forkhead/winged-helix protein, scurfin, results in the fatal lymphoproliferative disorder of the scurfy mouse. Nat Genet (2001) 27:68–73. doi:10.1038/83784

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

247. Ochs HD, Gambineri E, Torgerson TR. IPEX, FOXP3 and regulatory T-cells: a model for autoimmunity. Immunol Res (2007) 38:112–21. doi:10.1007/s12026-007-0022-2

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

248. Ohkura N, Kitagawa Y, Sakaguchi S. Development and maintenance of regulatory T cells. Immunity (2013) 38:414–23. doi:10.1016/j.immuni.2013.03.002

CrossRef Full Text | Google Scholar

249. Abbas AK, Benoist C, Bluestone JA, Campbell DJ, Ghosh S, Hori S, et al. Regulatory T cells: recommendations to simplify the nomenclature. Nat Immunol (2013) 14:307–8. doi:10.1038/ni.2554

CrossRef Full Text | Google Scholar

250. Roncarolo MG, Gregori S, Bacchetta R, Battaglia M. Tr1 cells and the counter-regulation of immunity: natural mechanisms and therapeutic applications. Curr Top Microbiol Immunol (2014) 380:39–68. doi:10.1007/978-3-662-43492-5_3

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

251. Vieira PL, Christensen JR, Minaee S, O’Neill EJ, Barrat FJ, Boonstra A, et al. IL-10-secreting regulatory T cells do not express Foxp3 but have comparable regulatory function to naturally occurring CD4+CD25+ regulatory T cells. J Immunol (2004) 172:5986–93. doi:10.4049/jimmunol.172.10.5986

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

252. Gagliani N, Magnani CF, Huber S, Gianolini ME, Pala M, Licona-Limon P, et al. Coexpression of CD49b and LAG-3 identifies human and mouse T regulatory type 1 cells. Nat Med (2013) 19:739–46. doi:10.1038/nm.3179

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

253. Kemper C, Chan AC, Green JM, Brett KA, Murphy KM, Atkinson JP. Activation of human CD4+ cells with CD3 and CD46 induces a T-regulatory cell 1 phenotype. Nature (2003) 421:388–92. doi:10.1038/nature01315

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

254. Buckner JH. Mechanisms of impaired regulation by CD4(+)CD25(+) FOXP3(+) regulatory T cells in human autoimmune diseases. Nat Rev Immunol (2010) 10:849–59. doi:10.1038/nri2889

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

255. Viglietta V, Baecher-Allan C, Weiner HL, Hafler DA. Loss of functional suppression by CD4+CD25+ regulatory T cells in patients with multiple sclerosis. J Exp Med (2004) 199:971–9. doi:10.1084/jem.20031579

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

256. Haas J, Hug A, Viehover A, Fritzsching B, Falk CS, Filser A, et al. Reduced suppressive effect of CD4+CD25high regulatory T cells on the T cell immune response against myelin oligodendrocyte glycoprotein in patients with multiple sclerosis. Eur J Immunol (2005) 35:3343–52. doi:10.1002/eji.200526065

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

257. Haas J, Fritzsching B, Trubswetter P, Korporal M, Milkova L, Fritz B, et al. Prevalence of newly generated naive regulatory T cells (Treg) is critical for Treg suppressive function and determines Treg dysfunction in multiple sclerosis. J Immunol (2007) 179:1322–30. doi:10.4049/jimmunol.179.2.1322

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

258. Venken K, Hellings N, Broekmans T, Hensen K, Rummens JL, Stinissen P. Natural naive CD4+CD25+CD127low regulatory T cell (Treg) development and function are disturbed in multiple sclerosis patients: recovery of memory Treg homeostasis during disease progression. J Immunol (2008) 180:6411–20. doi:10.4049/jimmunol.180.9.6411

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

259. Huan J, Culbertson N, Spencer L, Bartholomew R, Burrows GG, Chou YK, et al. Decreased FOXP3 levels in multiple sclerosis patients. J Neurosci Res (2005) 81:45–52. doi:10.1002/jnr.20522

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

260. Venken K, Hellings N, Thewissen M, Somers V, Hensen K, Rummens JL, et al. Compromised CD4+ CD25(high) regulatory T-cell function in patients with relapsing-remitting multiple sclerosis is correlated with a reduced frequency of FOXP3-positive cells and reduced FOXP3 expression at the single-cell level. Immunology (2008) 123:79–89. doi:10.1111/j.1365-2567.2007.02690.x

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

261. Costantino CM, Baecher-Allan C, Hafler DA. Multiple sclerosis and regulatory T cells. J Clin Immunol (2008) 28:697–706. doi:10.1007/s10875-008-9236-x

CrossRef Full Text | Google Scholar

262. Venken K, Hellings N, Liblau R, Stinissen P. Disturbed regulatory T cell homeostasis in multiple sclerosis. Trends Mol Med (2010) 16:58–68. doi:10.1016/j.molmed.2009.12.003

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

263. Bluestone JA, Tang Q, Sedwick CE. T regulatory cells in autoimmune diabetes: past challenges, future prospects. J Clin Immunol (2008) 28:677–84. doi:10.1007/s10875-008-9242-z

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

264. Arif S, Tree TI, Astill TP, Tremble JM, Bishop AJ, Dayan CM, et al. Autoreactive T cell responses show proinflammatory polarization in diabetes but a regulatory phenotype in health. J Clin Invest (2004) 113:451–63. doi:10.1172/JCI19585

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

265. Astier AL, Meiffren G, Freeman S, Hafler DA. Alterations in CD46-mediated Tr1 regulatory T cells in patients with multiple sclerosis. J Clin Invest (2006) 116:3252–7. doi:10.1172/JCI29251

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

266. Astier AL. T-cell regulation by CD46 and its relevance in multiple sclerosis. Immunology (2008) 124:149–54. doi:10.1111/j.1365-2567.2008.02821.x

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

267. Martinez-Forero I, Garcia-Munoz R, Martinez-Pasamar S, Inoges S, Lopez-Diaz de Cerio A, Palacios R, et al. . IL-10 suppressor activity and ex vivo Tr1 cell function are impaired in multiple sclerosis. Eur J Immunol (2008) 38:576–86. doi:10.1002/eji.200737271

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

268. Ma A, Xiong Z, Hu Y, Qi S, Song L, Dun H, et al. Dysfunction of IL-10-producing type 1 regulatory T cells and CD4(+)CD25(+) regulatory T cells in a mimic model of human multiple sclerosis in Cynomolgus monkeys. Int Immunopharmacol (2009) 9:599–608. doi:10.1016/j.intimp.2009.01.034

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

269. Ni Choileain S, Astier AL. CD46 processing: a means of expression. Immunobiology (2012) 217:169–75. doi:10.1016/j.imbio.2011.06.003

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

270. Grzybowski A, Pietrzak K. From patient to discoverer – Niels Ryberg Finsen (1860-1904) – the founder of phototherapy in dermatology. Clin Dermatol (2012) 30:451–5. doi:10.1016/j.clindermatol.2011.11.019

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

271. Loser K, Beissert S. Regulatory T cells: banned cells for decades. J Invest Dermatol (2012) 132:864–71. doi:10.1038/jid.2011.375

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

272. Loser K, Mehling A, Loeser S, Apelt J, Kuhn A, Grabbe S, et al. Epidermal RANKL controls regulatory T-cell numbers via activation of dendritic cells. Nat Med (2006) 12:1372–9. doi:10.1038/nm1518

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

273. Loser K, Beissert S. Regulation of cutaneous immunity by the environment: an important role for UV irradiation and vitamin D. Int Immunopharmacol (2009) 9:587–9. doi:10.1016/j.intimp.2009.01.024

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

274. Lehmann B, Sauter W, Knuschke P, Dressler S, Meurer M. Demonstration of UVB-induced synthesis of 1 alpha,25-dihydroxyvitamin D3 (calcitriol) in human skin by microdialysis. Arch Dermatol Res (2003) 295:24–8. doi:10.1007/s00403-003-0387-6

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

275. Elenkov IJ, Wilder RL, Bakalov VK, Link AA, Dimitrov MA, Fisher S, et al. IL-12, TNF-alpha, and hormonal changes during late pregnancy and early postpartum: implications for autoimmune disease activity during these times. J Clin Endocrinol Metab (2001) 86:4933–8. doi:10.1210/jcem.86.10.7905

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

276. Kitazawa R, Kitazawa S. Vitamin D(3) augments osteoclastogenesis via vitamin D-responsive element of mouse RANKL gene promoter. Biochem Biophys Res Commun (2002) 290:650–5. doi:10.1006/bbrc.2001.6251

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

277. Kitazawa S, Kajimoto K, Kondo T, Kitazawa R. Vitamin D3 supports osteoclastogenesis via functional vitamin D response element of human RANKL gene promoter. J Cell Biochem (2003) 89:771–7. doi:10.1002/jcb.10567

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

278. Kim S, Yamazaki M, Zella LA, Meyer MB, Fretz JA, Shevde NK, et al. Multiple enhancer regions located at significant distances upstream of the transcriptional start site mediate RANKL gene expression in response to 1,25-dihydroxyvitamin D3. J Steroid Biochem Mol Biol (2007) 103:430–4. doi:10.1016/j.jsbmb.2006.12.020

CrossRef Full Text | Google Scholar

279. Bishop KA, Coy HM, Nerenz RD, Meyer MB, Pike JW. Mouse Rankl expression is regulated in T cells by c-Fos through a cluster of distal regulatory enhancers designated the T cell control region. J Biol Chem (2011) 286:20880–91. doi:10.1074/jbc.M111.231548

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

280. Bishop KA, Wang X, Coy HM, Meyer MB, Gumperz JE, Pike JW. Transcriptional regulation of the human TNFSF11 gene in t cells via a cell type-selective set of distal enhancers. J Cell Biochem (2014) 116(2):320–30. doi:10.1002/jcb.24974

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

281. Gorman S, Kuritzky LA, Judge MA, Dixon KM, McGlade JP, Mason RS, et al. Topically applied 1,25-dihydroxyvitamin D3 enhances the suppressive activity of CD4+CD25+ cells in the draining lymph nodes. J Immunol (2007) 179:6273–83. doi:10.4049/jimmunol.179.9.6273

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

282. Ghoreishi M, Bach P, Obst J, Komba M, Fleet JC, Dutz JP. Expansion of antigen-specific regulatory T cells with the topical vitamin d analog calcipotriol. J Immunol (2009) 182:6071–8. doi:10.4049/jimmunol.0804064

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

283. Tomura M, Honda T, Tanizaki H, Otsuka A, Egawa G, Tokura Y, et al. Activated regulatory T cells are the major T cell type emigrating from the skin during a cutaneous immune response in mice. J Clin Invest (2010) 120:883–93. doi:10.1172/JCI40926

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

284. Milliken SV, Wassall H, Lewis BJ, Logie J, Barker RN, Macdonald H, et al. Effects of ultraviolet light on human serum 25-hydroxyvitamin D and systemic immune function. J Allergy Clin Immunol (2012) 129:1554–61. doi:10.1016/j.jaci.2012.03.001

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

285. Yamazaki S, Nishioka A, Kasuya S, Ohkura N, Hemmi H, Kaisho T, et al. Homeostasis of thymus-derived Foxp3+ regulatory T cells is controlled by ultraviolet B exposure in the skin. J Immunol (2014) 193(11):5488–97. doi:10.4049/jimmunol.1400985

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

286. Strle K, Zhou JH, Shen WH, Broussard SR, Johnson RW, Freund GG, et al. Interleukin-10 in the brain. Crit Rev Immunol (2001) 21:427–49. doi:10.1615/CritRevImmunol.v21.i5.20

CrossRef Full Text | Google Scholar

287. Ogawa Y, Duru EA, Ameredes BT. Role of IL-10 in the resolution of airway inflammation. Curr Mol Med (2008) 8:437–45. doi:10.2174/156652408785160907

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

288. Engelhardt KR, Grimbacher B. IL-10 in humans: lessons from the gut, IL-10/IL-10 receptor deficiencies, and IL-10 polymorphisms. Curr Top Microbiol Immunol (2014) 380:1–18. doi:10.1007/978-3-662-43492-5_1

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

289. Rubtsov YP, Rasmussen JP, Chi EY, Fontenot J, Castelli L, Ye X, et al. Regulatory T cell-derived interleukin-10 limits inflammation at environmental interfaces. Immunity (2008) 28:546–58. doi:10.1016/j.immuni.2008.02.017

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

290. Gabrysova L, Howes A, Saraiva M, O’Garra A. The regulation of IL-10 expression. Curr Top Microbiol Immunol (2014) 380:157–90. doi:10.1007/978-3-662-43492-5_8

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

291. Walter MR. The molecular basis of IL-10 function: from receptor structure to the onset of signaling. Curr Top Microbiol Immunol (2014) 380:191–212. doi:10.1007/978-3-662-43492-5_9

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

292. Stohlman SA, Pei L, Cua DJ, Li Z, Hinton DR. Activation of regulatory cells suppresses experimental allergic encephalomyelitis via secretion of IL-10. J Immunol (1999) 163:6338–44.

Pubmed Abstract | Pubmed Full Text | Google Scholar

293. Burkhart C, Liu GY, Anderton SM, Metzler B, Wraith DC. Peptide-induced T cell regulation of experimental autoimmune encephalomyelitis: a role for IL-10. Int Immunol (1999) 11:1625–34. doi:10.1093/intimm/11.10.1625

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

294. Barrat FJ, Cua DJ, Boonstra A, Richards DF, Crain C, Savelkoul HF, et al. In vitro generation of interleukin 10-producing regulatory CD4(+) T cells is induced by immunosuppressive drugs and inhibited by T helper type 1 (Th1)- and Th2-inducing cytokines. J Exp Med (2002) 195:603–16. doi:10.1084/jem.20011629

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

295. Gregori S, Giarratana N, Smiroldo S, Uskokovic M, Adorini L. A 1alpha,25-dihydroxyvitamin D(3) analog enhances regulatory T-cells and arrests autoimmune diabetes in NOD mice. Diabetes (2002) 51:1367–74. doi:10.2337/diabetes.51.5.1367

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

296. Gysemans C, van Etten E, Overbergh L, Giulietti A, Eelen G, Waer M, et al. Unaltered diabetes presentation in NOD mice lacking the vitamin D receptor. Diabetes (2008) 57:269–75. doi:10.2337/db07-1095

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

297. Palmer MT, Lee YK, Maynard CL, Oliver JR, Bikle DD, Jetten AM, et al. Lineage-specific effects of 1,25-dihydroxyvitamin D(3) on the development of effector CD4 T cells. J Biol Chem (2011) 286:997–1004. doi:10.1074/jbc.M110.163790

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

298. Zittermann A, Dembinski J, Stehle P. Low vitamin D status is associated with low cord blood levels of the immunosuppressive cytokine interleukin-10. Pediatr Allergy Immunol (2004) 15:242–6. doi:10.1111/j.1399-3038.2004.00140.x

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

299. Mann EH, Chambers ES, Pfeffer PE, Hawrylowicz CM. Immunoregulatory mechanisms of vitamin D relevant to respiratory health and asthma. Ann N Y Acad Sci (2014) 1317:57–69. doi:10.1111/nyas.12410

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

300. Urry Z, Xystrakis E, Richards DF, McDonald J, Sattar Z, Cousins DJ, et al. Ligation of TLR9 induced on human IL-10-secreting Tregs by 1alpha,25-dihydroxyvitamin D3 abrogates regulatory function. J Clin Invest (2009) 119:387–98. doi:10.1172/JCI32354

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

301. Dimeloe S, Richards DF, Urry ZL, Gupta A, Stratigou V, Farooque S, et al. 1alpha,25-dihydroxyvitamin D3 promotes CD200 expression by human peripheral and airway-resident T cells. Thorax (2012) 67:574–81. doi:10.1136/thoraxjnl-2011-200651

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

302. Urry Z, Chambers ES, Xystrakis E, Dimeloe S, Richards DF, Gabrysova L, et al. The role of 1alpha,25-dihydroxyvitamin D3 and cytokines in the promotion of distinct Foxp3+ and IL-10+ CD4+ T cells. Eur J Immunol (2012) 42:2697–708. doi:10.1002/eji.201242370

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

303. Gupta A, Dimeloe S, Richards DF, Chambers ES, Black C, Urry Z, et al. Defective IL-10 expression and in vitro steroid-induced IL-17A in paediatric severe therapy-resistant asthma. Thorax (2014) 69:508–15. doi:10.1136/thoraxjnl-2013-203421

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

304. Chambers ES, Suwannasaen D, Mann EH, Urry Z, Richards DF, Lertmemongkolchai G, et al. 1alpha,25-dihydroxyvitamin D3 in combination with transforming growth factor-beta increases the frequency of Foxp3(+) regulatory T cells through preferential expansion and usage of interleukin-2. Immunology (2014) 143:52–60. doi:10.1111/imm.12289

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

305. Thornton AM, Korty PE, Tran DQ, Wohlfert EA, Murray PE, Belkaid Y, et al. Expression of Helios, an Ikaros transcription factor family member, differentiates thymic-derived from peripherally induced Foxp3+ T regulatory cells. J Immunol (2010) 184:3433–41. doi:10.4049/jimmunol.0904028

CrossRef Full Text | Google Scholar

306. Getnet D, Grosso JF, Goldberg MV, Harris TJ, Yen HR, Bruno TC, et al. A role for the transcription factor Helios in human CD4(+)CD25(+) regulatory T cells. Mol Immunol (2010) 47:1595–600. doi:10.1016/j.molimm.2010.02.001

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

307. Zabransky DJ, Nirschl CJ, Durham NM, Park BV, Ceccato CM, Bruno TC, et al. Phenotypic and functional properties of Helios+ regulatory T cells. PLoS One (2012) 7:e34547. doi:10.1371/journal.pone.0034547

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

308. Royal W III, Mia Y, Li H, Naunton K. Peripheral blood regulatory T cell measurements correlate with serum vitamin D levels in patients with multiple sclerosis. J Neuroimmunol (2009) 213:135–41. doi:10.1016/j.jneuroim.2009.05.012

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

309. Smolders J, Menheere P, Thewissen M, Peelen E, Cohen Tervaert JW, Hupperts R, et al. Regulatory T cell function correlates with serum 25-hydroxyvitamin D, but not with 1,25-dihydroxyvitamin D, parathyroid hormone and calcium levels in patients with relapsing remitting multiple sclerosis. J Steroid Biochem Mol Biol (2010) 121:243–6. doi:10.1016/j.jsbmb.2010.03.001

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

310. Khoo AL, Koenen HJ, Chai LY, Sweep FC, Netea MG, van der Ven AJ, et al. Seasonal variation in vitamin D(3) levels is paralleled by changes in the peripheral blood human T cell compartment. PLoS One (2012) 7:e29250. doi:10.1371/journal.pone.0029250

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

311. Guven A, Ecevit A, Sozer O, Tarcan A, Tarcan A, Ozbek N. Correlation between the cord vitamin D levels and regulatory T cells in newborn infants. Eur J Pediatr (2012) 171:1161–6. doi:10.1007/s00431-012-1688-6

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

312. Baeke F, Korf H, Overbergh L, Verstuyf A, Thorrez L, Van Lommel L, et al. The vitamin D analog, TX527, promotes a human CD4+CD25highCD127low regulatory T cell profile and induces a migratory signature specific for homing to sites of inflammation. J Immunol (2011) 186:132–42. doi:10.4049/jimmunol.1000695

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

313. Schubert D, Bode C, Kenefeck R, Hou TZ, Wing JB, Kennedy A, et al. Autosomal dominant immune dysregulation syndrome in humans with CTLA4 mutations. Nat Med (2014) 20(12):1410–6. doi:10.1038/nm.3746

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

314. Kuehn HS, Ouyang W, Lo B, Deenick EK, Niemela JE, Avery DT, et al. Immune dysregulation in human subjects with heterozygous germline mutations in CTLA4. Science (2014) 345:1623–7. doi:10.1126/science.1255904

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

315. Qureshi OS, Zheng Y, Nakamura K, Attridge K, Manzotti C, Schmidt EM, et al. Trans-endocytosis of CD80 and CD86: a molecular basis for the cell-extrinsic function of CTLA-4. Science (2011) 332:600–3. doi:10.1126/science.1202947

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

316. Tone Y, Furuuchi K, Kojima Y, Tykocinski ML, Greene MI, Tone M. Smad3 and NFAT cooperate to induce Foxp3 expression through its enhancer. Nat Immunol (2008) 9:194–202. doi:10.1038/ni1549

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

317. Ni Choileain S, Weyand NJ, Neumann C, Thomas J, So M, Astier AL. The dynamic processing of CD46 intracellular domains provides a molecular rheostat for T cell activation. PLoS One (2011) 6:e16287. doi:10.1371/journal.pone.0016287

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

318. Cardone J, Le Friec G, Vantourout P, Roberts A, Fuchs A, Jackson I, et al. Complement regulator CD46 temporally regulates cytokine production by conventional and unconventional T cells. Nat Immunol (2010) 11:862–71. doi:10.1038/ni.1917

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

319. Xu YQ, Gao YD, Yang J, Guo W. A defect of CD4+CD25+ regulatory T cells in inducing interleukin-10 production from CD4+ T cells under CD46 costimulation in asthma patients. J Asthma (2010) 47:367–73. doi:10.3109/02770903.2010.481340

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

320. Ghaffarizadeh A, Flann NS, Podgorski GJ. Multistable switches and their role in cellular differentiation networks. BMC Bioinformatics (2014) 15(Suppl 7):S7. doi:10.1186/1471-2105-15-S7-S7

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

321. Ross EM, Bourges D, Hogan TV, Gleeson PA, van Driel IR. Helios defines T cells being driven to tolerance in the periphery and thymus. Eur J Immunol (2014) 44:2048–58. doi:10.1002/eji.201343999

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

322. Baine I, Basu S, Ames R, Sellers RS, Macian F. Helios induces epigenetic silencing of IL2 gene expression in regulatory T cells. J Immunol (2013) 190:1008–16. doi:10.4049/jimmunol.1200792

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

323. Maeki T, Mori Y. Features of human herpesvirus-6A and -6B entry. Adv Virol (2012) 2012:384069. doi:10.1155/2012/384069

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

324. de Jong BA, Schrijver HM, Huizinga TW, Bollen EL, Polman CH, Uitdehaag BM, et al. Innate production of interleukin-10 and tumor necrosis factor affects the risk of multiple sclerosis. Ann Neurol (2000) 48:641–6. doi:10.1002/1531-8249(200010)48:4<641::AID-ANA11>3.0.CO;2-Z

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

325. Martinez Doncel A, Rubio A, Arroyo R, de las Heras V, Martin C, Fernandez-Arquero M, et al. Interleukin-10 polymorphisms in Spanish multiple sclerosis patients. J Neuroimmunol (2002) 131:168–72. doi:10.1016/S0165-5728(02)00248-5

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

326. Almeras L, Meresse B, Seze J, De Lefranc D, Dubucquoi S, Fajardy I, et al. Interleukin-10 promoter polymorphism in multiple sclerosis: association with disease progression. Eur Cytokine Netw (2002) 13:200–6.

Pubmed Abstract | Pubmed Full Text | Google Scholar

327. Luomala M, Lehtimaki T, Huhtala H, Ukkonen M, Koivula T, Hurme M, et al. Promoter polymorphism of IL-10 and severity of multiple sclerosis. Acta Neurol Scand (2003) 108:396–400. doi:10.1034/j.1600-0404.2003.00165.x

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

328. Myhr KM, Vagnes KS, Maroy TH, Aarseth JH, Nyland HI, Vedeler CA. Interleukin-10 promoter polymorphisms in patients with multiple sclerosis. J Neurol Sci (2002) 202:93–7. doi:10.1016/S0022-510X(02)00246-0

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

329. Dyment DA, Cader MZ, Chao MJ, Lincoln MR, Morrison KM, Disanto G, et al. Exome sequencing identifies a novel multiple sclerosis susceptibility variant in the TYK2 gene. Neurology (2012) 79:406–11. doi:10.1212/WNL.0b013e3182616fc4

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

330. Moore KW, de Waal Malefyt R, Coffman RL, O’Garra A. Interleukin-10 and the interleukin-10 receptor. Annu Rev Immunol (2001) 19:683–765. doi:10.1146/annurev.immunol.19.1.683

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text | Google Scholar

Keywords: vitamin D, CD4-positive T lymphocytes, autoimmune diseases, multiple sclerosis, type 1 diabetes

Citation: Hayes CE, Hubler SL, Moore JR, Barta LE, Praska CE and Nashold FE (2015) Vitamin D actions on CD4+ T cells in autoimmune disease. Front. Immunol. 6:100. doi: 10.3389/fimmu.2015.00100

Received: 05 December 2014; Accepted: 23 February 2015;
Published online: 18 March 2015.

Edited by:

Anne L. Astier, University of Edinburgh, UK

Reviewed by:

Yehuda Julyus Shoenfeld, Sheba Medical Center, Israel
Jan Damoiseaux, Maastricht University Medical Center, Netherlands

Copyright: © 2015 Hayes, Hubler, Moore, Barta, Praska and Nashold. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) or licensor are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Colleen Elizabeth Hayes, Department of Biochemistry, College of Agricultural and Life Sciences, University of Wisconsin-Madison, 433 Babcock Drive, Madison, WI 53706, USA e-mail: hayes@biochem.wisc.edu

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.