Next Article in Journal
Types of Breast Cancer Surgery and Breast Reconstruction
Previous Article in Journal
Inorganic Nanomedicine—Mediated Ferroptosis: A Synergistic Approach to Combined Cancer Therapies and Immunotherapy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Vitamin D in Cancer Prevention and Treatment: A Review of Epidemiological, Preclinical, and Cellular Studies

by
Siva Dallavalasa
1,†,
SubbaRao V. Tulimilli
1,†,
Vidya G. Bettada
1,
Medha Karnik
1,
Chinnappa A. Uthaiah
1,
Preethi G. Anantharaju
1,
Suma M. Nataraj
1,
Rajalakshmi Ramashetty
2,
Olga A. Sukocheva
3,*,
Edmund Tse
3,
Paramahans V. Salimath
4,‡ and
SubbaRao V. Madhunapantula
1,5,*
1
Center of Excellence in Molecular Biology and Regenerative Medicine (CEMR) Laboratory (DST-FIST Supported Center and ICMR Collaborating Center of Excellence—ICMR-CCoE), Department of Biochemistry (DST-FIST Supported Department), JSS Medical College, JSS Academy of Higher Education & Research (JSS AHER), Mysuru 570015, Karnataka, India
2
Department of Physiology, JSS Medical College, JSS Academy of Higher Education & Research (JSS AHER), Mysuru 570015, Karnataka, India
3
Department of Hepatology, Royal Adelaide Hospital, Port Rd., Adelaide, SA 5000, Australia
4
JSS Academy of Higher Education & Research (JSS AHER), Mysuru 570015, Karnataka, India
5
Special Interest Group in Cancer Biology and Cancer Stem Cells (SIG-CBCSC), JSS Medical College, JSS Academy of Higher Education & Research (JSS AHER), Mysuru 570015, Karnataka, India
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Superannuated from JSS Academy of Higher Education & Research, Mysuru 570015, Karnataka, India.
Cancers 2024, 16(18), 3211; https://doi.org/10.3390/cancers16183211
Submission received: 8 August 2024 / Revised: 12 September 2024 / Accepted: 17 September 2024 / Published: 20 September 2024
(This article belongs to the Special Issue Emerging Trends in Global Cancer Epidemiology: 2nd Edition)

Abstract

:

Simple Summary

Inhibition of human cancers has previously been linked to the administration of vitamin D. Studies have shown that increased cancer incidence is associated with decreased vitamin D. The anticancer activity of vitamin D has been confirmed by several in vitro and in vivo studies. Vitamin D inhibits the growth of cancer cells by (a) the induction of apoptosis, (b) decreasing the metastatic spread, (c) arresting the cells at the G0/G1 (or) G2/M phase in the cell cycle, and (d) downregulating proliferation signals. Supplementation of vitamin D slows down the growth of xenografted tumors in mice. Hence, vitamin D could be considered a potential cancer chemotherapeutic agent.

Abstract

Background: Inhibition of human carcinomas has previously been linked to vitamin D due to its effects on cancer cell proliferation, migration, angiogenesis, and apoptosis induction. The anticancer activity of vitamin D has been confirmed by several studies, which have shown that increased cancer incidence is associated with decreased vitamin D and that dietary supplementation of vitamin D slows down the growth of xenografted tumors in mice. Vitamin D inhibits the growth of cancer cells by the induction of apoptosis as well as by arresting the cells at the G0/G1 (or) G2/M phase of the cell cycle. Aim and Key Scientific Concepts of the Review: The purpose of this article is to thoroughly review the existing information and discuss and debate to conclude whether vitamin D could be used as an agent to prevent/treat cancers. The existing empirical data have demonstrated that vitamin D can also work in the absence of vitamin D receptors (VDRs), indicating the presence of multiple mechanisms of action for this sunshine vitamin. Polymorphism in the VDR is known to play a key role in tumor cell metastasis and drug resistance. Although there is evidence that vitamin D has both therapeutic and cancer-preventive properties, numerous uncertainties and concerns regarding its use in cancer treatment still exist. These include (a) increased calcium levels in individuals receiving therapeutic doses of vitamin D to suppress the growth of cancer cells; (b) hyperglycemia induction in certain vitamin D-treated study participants; (c) a dearth of evidence showing preventive or therapeutic benefits of cancer in clinical trials; (d) very weak support from proof-of-principle studies; and (e) the inability of vitamin D alone to treat advanced cancers. Addressing these concerns, more potent and less toxic vitamin D analogs have been created, and these are presently undergoing clinical trial evaluation. To provide key information regarding the functions of vitamin D and VDRs, this review provided details of significant advancements in the functional analysis of vitamin D and its analogs and VDR polymorphisms associated with cancers.

1. Introduction

Vitamin D, also called the “sunshine” vitamin, is formed when skin is exposed to sunlight. It is well known that vitamin D keeps the serum calcium levels within the physiological range, i.e., 8.5 to 10.2 mg/dL, and preserves bone health [1]. Any concentration of vitamin D less than 30 ng/mL of blood is considered vitamin D insufficiency. To maintain vitamin D levels within the normal range (30–50 ng/mL blood), the Endocrine Society recommends 400–1000 international units (IU; 1 IU = 0.025 µg of vitamin D) every day for infants under the age of one year, 600–1000 IU for children and adolescents, and 1500–2000 IU for adults [2,3]. Earlier studies have demonstrated that vitamin D deficiency is a global health concern [4]. Recent findings in public health research have also demonstrated that one of the most prevalent conditions seen in cancer patients is vitamin D insufficiency [5]. Furthermore, meta-analyses of randomized trials revealed a robust correlation between vitamin D sufficiency and a lower cancer death rate [6]. Vitamin D supplementation, however, does not appear to have any effect on tumor development, cancer-induced mortality, and tumor cells’ susceptibility to radiation. Additionally, several studies evaluating the impact of vitamin D supplementation on cancer incidence and death have yielded poor results, suggesting that more studies are necessary to validate vitamin D’s potential as a cancer preventive and therapeutic [7]. Due to contradictory findings about vitamin D’s role in cancer prevention and therapy [8], an in-depth review is necessary to evaluate vitamin D’s uses and functions. Although it has been shown that vitamin D can inhibit the development of cancer even in cell lines or tissues that express a mutant version of the VDR, additional investigations are needed to determine whether or not VDR expression is a prerequisite for vitamin D to have anticancer activity [9]. Recent studies have demonstrated the prolonged antiproliferative potential and increased biological activity of vitamin D analogs [10]. Therefore, this review also covers recent research on vitamin D analogs.

1.1. VDR Structure and Function: A Short Overview

The gene encoding for a VDR is located on the long arm of chromosome 12. The VDR gene is made up of a promoter region, regulatory regions 1a–1f, and exons 2–9. The VDR protein is made up of six domains, viz., A–F (Figure 1) [11]. The nuclear localization region (represented in RED) of the VDR protein guides the receptor into the nucleus [12,13,14,15]. When calcitriol binds to the hormone-binding domain (represented in GREEN color), PKC phosphorylates serine 51 in the VDR DNA-binding domain (represented in YELLOW color), and CKII phosphorylates serine 208 in the hinge region [16,17]. The phosphorylated VDR complex can dimerize with RXR via dimerization domains (represented in BLUE color) and form the calcitriol–VDR–RXR complex. This tripartite complex binds to the VDRE and influences the expression of target genes. The conformational changes in VDR also influence gene expression linked to the separation of the co-repressor SMRT. The interaction of the VDR with the activation function 2 (AF2) transactivation domain (represented in LIGHT GREY color) and stimulatory coactivators (such as steroid receptor coactivators (SRCs)) is made possible by the dissociation of SMRT. The nuclear coactivator-62-kDa-Ski-interacting protein (NCoA62–SKIP) complex, in conjunction with VDR-interacting proteins, enhances the transcriptional activation of VDR target genes [11].

1.2. Vitamin D Metabolism and Mechanism: Short Overview

The physiologically active form of vitamin D, calcitriol (1α,25(OH)2D3), is synthesized in a well-controlled multistep process [18]. The two main isoforms of vitamin D, vitamin D2 (ergocalciferol) and vitamin D3 (cholecalciferol), are produced from ergosterol by UV-B radiation. In humans, UV-B radiation can also produce 7-dehydrocholesterol in the epidermis of skin [19,20]. The prohormone vitamin D binds to its nuclear receptors to regulate a variety of physiological processes after being metabolized into a physiologically active substance [9]. In the blood, vitamin D is circulated in the form of 25-hydroxycholecalciferol [25(OH)D3, calcidiol or circulating vitamin D], which is produced by the metabolism of cholecalciferol in the liver by 25-hydroxylase. This hydroxylase is encoded by the CYP27A1 gene [21,22]. The physiologically active 1α,25 hydroxycholecalciferol [1,25(OH)2D3, calcitriol or active form of vitamin D] is produced in the kidney by 25-hydroxyvitamin D3-1α-hydroxylase, which is encoded by the CYP27B1 gene [21,22]. Calcitriol enters the bloodstream, where it binds to the vitamin D-binding protein (VDBP) and travels to the kidney, bone, and gut, among other target organs, to control the uptake, mobilization, and, finally, reabsorption of calcium and phosphate [19]. The hydroxylation at position 24 by the cytochrome P-450 enzyme generates 24,25(OH)2 D3 and 1α,24,25(OH)2 D3 [11]. 24-hydroxylase is encoded by the CYP24A1 gene. An increase in 24,25(OH)2 D3 causes the production of 1α,25(OH)2 D3, which in turn regulates the levels of calcitriol. Increases in Ca2+, inorganic phosphate, and calcitriol itself may inhibit 1α,25(OH)2D3 production [11,23,24,25]. Parathyroid hormone (PTH) induces CYP27B1 expression, while 1α,25(OH)2D3 represses it [23,26]. In contrast to CYP27B1, 1α,25(OH)2D3 substantially induces the expression of CYP24A [27] (Figure 2).
Calcitriol binds to either membrane-bound (non-genomic pathway) or cytosolic (genomic pathway) VDRs in target tissues, thereby inducing the expression of downstream target genes that are involved in controlling several biological processes. The VDR belongs to the family of ligand-activated transcription factors that are nuclear receptors [9,28,29,30]. In the genomic pathway, calcitriol binds with the cytosolic VDR to create a complex that stimulates protein kinase C (PKC) and casein kinase II (CKII) to phosphorylate the VDR on serine 51 and serine 208, respectively. A complex formed by the heterodimerization of the VDR with the RXR is translocated into the nucleus [31]. The translocated calcitriol–VDR–RXR complex regulates mRNA expression by binding to the VDRE and recruits transcriptional modulators to the promoter region of target genes [32]. Through direct protein–protein interaction, calcitriol binds to membrane-bound VDRs, or 1,25D-MARRS, in the non-genomic pathway. This causes acute changes in cell signaling pathways, including calcium and mitogen-activated protein kinase (MAPK) signaling [28,33] (Figure 2).

1.3. Is Vitamin D a Good Cancer Prevention Agent?

Vitamin D sensitizes cancer cells to the chemotherapeutic agent 5-fluorouracil (5-FU) by downregulating the expression of the antiapoptotic protein survivin and thymidylate synthase, a key enzyme involved in the biosynthesis of DNA. Additionally, in vitro studies have demonstrated that vitamin D promotes the differentiation of cells by increasing the expression of several cell adhesion components that are required for maintaining the epithelial phenotype along with proteins associated with the actin cytoskeleton and intermediate filaments. Treatment with vitamin D has been shown to significantly slow down tumor development and boost cancer cells’ susceptibility to chemotherapeutic drugs [34]. Epidemiological studies played a major role in the initial scientific discussion on vitamin D. The early studies were conducted in the United States, which provided evidence indicative of a north–south gradient in the risk of specific cancers. This finding has prompted the hypothesis of a protective influence of vitamin D on the risk of cancer at various sites. For instance, it has been demonstrated that vitamin D has a critical role in reducing the incidence of colorectal cancer [35]. Several observational studies have also reported a negative correlation between serum vitamin D levels and the risk of developing several cancers, including breast, colorectal, kidney, lung, and pancreatic [36]. Additionally, studies have reported a correlation between vitamin D deficiency and increased mortality due to several prevalent malignancies [37].
According to a meta-analysis, consuming enough vitamin D significantly lowers the risk and mortality of cancer [38]. Circulating vitamin D correlated well with patient survival, indicating that it may be a potential prognostic marker in advanced pancreatic cancer [39]. Yejin Kim et al. demonstrated that serum vitamin D could be used as a predictive marker for screening people at risk for CRC [40]. According to a case–control study of 102 participants, methylation levels of important CpG sites in VDRs, CYP24A1, and CYP2R1 are negatively correlated with the incidence of colorectal cancer (CRC) [41]. The CYP24A1 gene, which is located on chromosome 20, encodes for the 24-OHase. The enzyme 24-OHase is involved in the catabolism of 25(OH)D and 1,25(OH)2D and is responsible for the regulation of the concentration of these two vitamin D metabolites in the circulation. Studies have demonstrated that CYP24A1 is induced by 25(OH)D, 1,25(OH)2D, and FGF23. The CYP2R1 gene, which is located on chromosome 11, codes for an enzyme whose biochemical properties are consistent with a D 25-hydroxylase. The CYP2R1 gene transcript has characteristic sequence features associated with cytochromes P450 of the endoplasmic reticulum.
In a different study, Xuezhao Chen et al. used Mendelian randomization to examine the relationship between serum vitamin D levels and obesity as risk factors for basal cell carcinoma (BCC). This study found that a high BMI may increase the risk of BCC at the genetic level, but vitamin D has no effect on BCC [42].
Numerous epidemiological studies have also documented the association between vitamin D and different cancer types [43]. A cohort study of 182 children with neuroblastoma reported a lack of correlation between vitamin D deficiency and high-risk neuroblastoma [44]. Another study, which analyzed the function of vitamin D addition on telomere length, reported that administering monthly doses of vitamin D to older patients (age > 60) did not affect telomere length [45]. According to a cohort study involving 476 women with breast cancer, the patients with adequate vitamin D levels had smaller and lower-grade tumors than those with insufficient vitamin D (Table 1) [46]. Although these observational studies reported that serum vitamin D levels correlate with a lower incidence of cancers, they include various confounding factors like outdoor activity, adiposity, and overall nutrition status [47]. Hence, further evidence from RCTs is required to confirm these findings.
Several clinical trials testing the impact of vitamin D on cancer incidence, progression, and mortality have supported the benefits of vitamin D treatment and addressed some of its associated concerns [49]. Clinical trials conducted previously showed the antiproliferative effect of calcitriol in acute myeloid leukemia (AML) patients [50]. Several trials have successfully demonstrated a positive correlation between higher vitamin D levels in cancer patients and increased disease-free survival, lower recurrence risk, and reduced mortality rate [51]. One such study initiated in Norway, a country with minimal sun exposure, showed a better survival rate in cancer patients with higher vitamin D levels [52]. This study monitored vitamin D levels in a total of 658 patients suffering from breast (n = 251), colon (n = 52), lung (n = 210), and lymph (n = 145) carcinomas. The participants were divided into first (<2.5 mg/dL), second (2.5–3.4 mg/dL), third (3.4–4.5 mg/dL), and fourth (>4.5 mg/dL) quartiles based on circulating vitamin D. Death due to cancer was significantly lower (HR 0.36 95% CI 0.27, 0.51) in patients within the fourth quartile (>4.5 mg/dL) when compared to patients in the first and second quartiles. These results support a positive relationship between circulating 25-OHD and better survival in cancer patients [52]. Since the half-life of the active form of vitamin D (calcitriol) in systemic circulation is very short (15 h compared to 15 days for calcidiol), and the amount of circulating calcitriol is much lower (0.029 ng/mL to 0.083 ng/mL compared to 30 to 50 ng/mL of calcidiol). The levels of circulating vitamin D (calcidiol) rather than calcitriol are measured to determine the vitamin D status.
Robsahm, Trude Eid, et al. made an attempt to understand the effect of vitamin D in pre- and post-cancer diagnosis [53]. The study was conducted with 556 participants, which included breast cancer (n = 202), lung cancer (n = 193), lymphoma (n = 124), and colon cancer (n = 37) cases. Serum levels of 25-(OH)D were assessed before and after cancer incidence, and participants were divided into low (<2.5 mg/dL) and high (≥2.5 mg/dL) categories. The median 25-OHD levels were 3.5 and 3.4 mg/dL for pre- and post-diagnosis, respectively. Patients who had both samples at high (≥3.4 mg/dL) levels had a 59% lower mortality risk than the patients whose vitamin D levels were <2.5 mg/dL. Compared to individuals whose serum vitamin D levels were steady even after diagnosis, lower levels were linked to an increased risk of death [53].
A comprehensive meta-analysis comprising 23 case–control and 10 prospective cohort studies revealed a correlation between exposure and outcome, referring to circulating total 25(OH)D and CRC, respectively, in the overall population [54]. Whereas circulating 25(OH)D and 1,25(OH)2D were considered for both men and women in the case–control studies, only circulating 25(OH)D data were assessed for both men and women in prospective cohort studies [54]. Results of the case–control-based analysis showed a 39% lower risk of CRC in populations with serum 25(OH)D >33 ng/mL when compared to 12 ng/mL ((95% CI): 0.61 (0.52; 0.71); 11 studies) and a 20% lower risk of CRC in prospective studies (HR (95% CI): 0.80 (0.66; 0.97); 6 studies). However, these results were only observed in the female population and failed to show any significance in the male population [54]. Another separate analysis evaluated vitamin D levels in pooled data from 17 cohorts, which consisted of 5706 CRC patients and 7107 controls [55]. Based on circulating 25(OH)D levels, the study participants were divided into low-level (2.8– <3.4 mg/dL), deficient (1.6 mg/dL), and high-level (4.1–4.8 and 4.8– <5.5 mg/dL) groups. The deficient group had a 31% higher risk of CRC (RR = 1.31, 95% CI = 1.05 to 1.62). The higher level of vitamin D was linked to a lower risk of CRC. Interestingly, an inverse correlation of vitamin D level with CRC risk was observed only among the female participants but not in male participants [55].

Vitamin D Supplementation and Cancer

Rather than focusing on cancer incidence, several observational studies were carried out to examine any potential associations between vitamin D supplementation and cancer mortality [56]. For instance, a study conducted by Poole et al. demonstrated no significant changes in the mortality rate of women upon receiving 1 year of supplementation of vitamin D [57]. A retrospective study, which included a total of 308 breast cancer patients on Trastuzumab chemotherapy, showed that supplementation of <10,000 IU/week of vitamin D along with Trastuzumab improved the disease-free survival rate in HER2+ non-metastatic breast cancer patients when compared to those patients who did not receive the weekly vitamin D dosage [58]. In a separate study, the intake of vitamin D as a component of diet or as a supplement was studied in 434 head and neck cancer patients. Analysis of the data showed no significant effect on the mortality rate; however, lower disease recurrence was found in individuals with higher vitamin D levels [59].
A case–control study conducted in the US has evaluated the impact of vitamins and fatty acids on glioblastoma patients [60]. A total of 470 patients were considered, of which 77% of them were on alternative therapies. The addition of vitamin D has reduced the mortality rate of patients when compared to those who were not taking the supplement. However, this association was not observed after applying multivariate adjustments for factors like the KPS (Karnofsky Performance Scale) and tumor removal [60]. The use of de novo vitamin D among women between the ages of 50 and 80 years who were diagnosed with breast cancer was associated with a 20% decrease in the mortality rate [61]. Furthermore, a decrease of 49% in mortality rate was reported in patients who were put on vitamin D supplements soon after breast cancer diagnosis [61]. Another study measured the effect of vitamin D intake on the early inception of CRC [62]. A total of 111 women aged below 50 years were included in the study. The results of this study showed a significantly lowered risk of early onset of CRC with increased vitamin D intake (HR for ≥450 IU/day vs <300 IU/day, 0.49; 95% CI, 0.26–0.93; P for trend = 0.01). Furthermore, this association was evident when there was an increase in the dietary intake of vitamin D (HR per 400 IU/day increase, 0.34; 95% CI, 0.15–0.79) rather than vitamin D supplementation (HR per 400 IU/day increase, 0.77; 95% CI, 0.37–1.62), demonstrating that dietary intake is more beneficial compared to the supplements in reducing the risk of CRC onset [62]
Numerous RCTs have been carried out to evaluate the impact of vitamin D in the prevention of cancer; among them, only three RCTs published by Lappe et al. in 2007 [63], Bolland et al. in 2011 [64], and Lappe et al. in 2017 [65] have corroborated the findings of observational studies and documented a decrease in cancer risk in those using vitamin D supplements [66]. Nevertheless, the analysis in all three RCTs was limited to those who had not used calcium or vitamin D supplements before enrolling in the study [66]. Furthermore, these RCTs overlooked the fact that the duration of vitamin D (25(OH)D) in the blood, rather than the vitamin D dosage itself, influences the health outcome [67]. Taking these limitations into account, the biggest RCT, known as VITAL (NCT 01169259), examined the potential of a daily dose of 2000 IU vitamin D3 to reduce cancer risk over five years. The RCT included 25,871 participants. The findings of the VITAL trial revealed that vitamin D addition did not affect the incidence of invasive cancer in post-menopausal women [68]. In addition, the study findings showed that 25-hydroxyvitamin D levels were not associated with subsequent breast cancer risk [69]. A meta-analysis by Keum N., et al. 2019 reported that vitamin D supplementation (circulating levels of 25(OH)D around 3.0–7.4 mg/dL) significantly reduced total cancer mortality but failed to reduce total cancer incidence [70]. Virtanen J.K., et al., 2022 reported that vitamin D3 supplementation (1600 IU/day or 3200 IU/day) did not lower invasive cancer incidence among older adults, possibly due to sufficient vitamin D at baseline (4.1 mg/dL) in a majority of study participants [71]. Following these publications, arguments on vitamin D’s potential to prevent cancer have acquired even more traction. One of the potential reasons for not observing a positive correlation between vitamin D and a lower risk of cancer is that the vitamin D levels at baseline are too high in the participants to show the benefit of supplementation. In conclusion, further in-depth studies are warranted to find out whether taking supplements of vitamin D can help prolong the protection against cancers.

1.4. Darker Side of the Sunshine Vitamin

Although several epidemiological studies have shown a negative correlation between serum vitamin D levels and the incidence of cancer, little attention has been paid to the use of elevated vitamin D dosages in the treatment of cancer and its potential health consequences. According to research by Garland and colleagues, maintaining blood levels of vitamin D metabolites—which are essential to decrease the risk of breast and colon cancers—requires daily consumption of vitamin D in the range of 4000–8000 IU [72]. On the other hand, prolonged use of vitamin D supplements at a dosage exceeding 400–600 IU (i.e., RDA) raised blood vitamin D levels and resulted in hypervitaminosis. Although hypervitaminosis is more likely to occur when someone takes more than 10,000 IU of vitamin D per day, it is a possibility that excessive doses of vitamin D might be consumed either due to misuse of over-the-counter supplements or erroneous prescriptions [73]. Hypercalcemia, or an excess of calcium in the blood, is one of the main effects of hypervitaminosis. Excessive calcium in the blood can cause problems with the kidneys and bones, including kidney stones. Therefore, to prevent hypervitaminosis, individuals utilizing vitamin D supplements should measure their serum vitamin D. Innovative approaches are warranted to mitigate these side effects while maintaining the therapeutic level of vitamin D in serum. Using vitamin D derivatives, which do not raise blood calcium but encourage the death of cancer cells, is one possible strategy. The studies that tested the effects of vitamin D analogs on cancer incidence and treatment will be described in Section 2.

1.5. A Promising Potential of Vitamin D Supplement: Anticancer Effects

People who obtain enough vitamin D have lower cancer incidence and fewer cancer-related deaths. This has been shown in several epidemiological and observational studies [43]. Recent findings have provided support for this epidemiological observation by demonstrating greater death rates among those who received very little natural light in comparison to those who lived in higher-latitude locations, which are linked to higher levels of naturally synthesized vitamin D in the skin [74]. Studies have also shown that sunlight protects against the incidence of different cancers, including skin, prostate, colorectal, breast, and ovarian cancers [75]. Circulating vitamin D was shown to be correlated with a decreased prevalence of prostate and colorectal malignancies [76]. The first evidence of vitamin D growth inhibitory effects in tumor cells was shown by Colston and colleagues by demonstrating a pivotal role of VDR in malignant melanoma [77]. Subsequently, several other studies validated vitamin D’s antitumor properties [78,79,80,81,82,83,84,85,86,87,88]. Vitamin D and its analogs show their antitumor properties by (a) inhibiting cell proliferation, (b) inducing proapoptotic genes, (c) reducing angiogenesis, and (d) blocking the spread of cancerous cells [89]. Recent findings highlighted the growth-inhibitory properties of vitamin D in several cancerous cells and demonstrated the differentiation of leukemia cells (HL60) to macrophage lineage cells [90]. Several mechanisms and effectors, such as prostaglandins, COX-2, 15-PGDH, and EP2, were reported to be involved in the vitamin D-mediated inhibition of prostate cancer cell proliferation [91,92]. The combination of vitamin D with gemcitabine showed the altered gene expression of pancreatic stellate cells and minimized the tumor volume [93]. The proliferation of breast cancer cells was inhibited by vitamin D via selective downregulation of estrogen receptor alpha (ERα) in the malignant cells and blockade of aromatase expression in breast adipose tissue [94]. To mitigate the toxic effects, such as apathy, drowsiness, depression, psychosis, polydipsia, anorexia, constipation, peptic ulcers, hypercalciuria, polyuria, polydipsia, dehydration, nephrocalcinosis, and renal failure, which are associated with the administration of vitamin D, analogs of vitamin D are now being investigated. Vitamin D analogs are reported to be beneficial to health in preclinical and clinical studies [95]. For instance, several studies have demonstrated that vitamin D and analogs can trigger apoptosis in cancer cells [96]. Notably, the observed cancer cell death was induced through the (1) inhibition of angiogenesis and metastatic potential, (2) upregulation of proapoptotic gene expression, (3) activation of antitumoral immunity, and (4) chemotherapy sensitization [95]. Vitamin D-mediated inhibition of angiogenesis prevents the release of several growth- and survival-promoting factors and deprives cancer cells of oxygen and nutrients [97]. Mechanistically, the calcitriol–VDR complex induces conformational changes in the receptor structure, thereby promoting the binding of the VDR to the retinoic acid receptor (RXR). Interaction of VDRs with RXRs activates VDRE in the promoter region of various target genes, thereby controlling cell division, differentiation, proliferation, and metastasis [98].

1.6. Cancer Risk Reduction by Vitamin D Metabolite Calcitriol

Calcitriol activates genes that control stress response, DNA repair, immune responses, and various transcriptional factors responsible for the regulation of cellular processes [99]. These additional functions of calcitriol, which are not necessarily linked to the regulation of calcium homeostasis, represent a potential role of this vitamin D metabolite in the prevention and treatment of cancers [100]. The ability of calcitriol to resolve inflammation may be exploited for the inhibition of angiogenesis and cancer cell proliferation [101]. For instance, an enhanced antiproliferative effect was reported in ER+ breast cancer cells by calcitriol. Mechanistically, calcitriol downregulated aromatase transcription and blocked the synthesis of estrogen [102].
The activity of calcitriol is self-regulated by CYP24A1, which encodes an enzyme that catalyzes the degradation of both calcitriol and 25-hydroxycholecalciferol, thereby reducing the calcemic effects compared to vitamin D [103]. Notably, the presence of enzyme alpha 1-hydroxylase (encoded by CYP27B1) was reported in normal and malignant breast tissues [102]. The enzyme converts circulating prohormone 25(OH)D3 to the active hormone calcitriol in the breast tissue [102]. Therefore, the fortification of vitamin D in the diet helps to enhance the levels of substrate for CYP27B1, increase the production of calcitriol at the local tissue level, and inhibit the progression of breast cancer [102]. Accordingly, local production of 1, 25(OH)2D3 by alpha 1-hydroxylase was noted in breast cancer tissues [104]. Increased 1,25(OH)2D3 production due to dysregulated extra-renal alpha 1-hydroxylase expression was reported in B-cell lymphoma and breast, colon, and prostate cancers [105]. Increased expression (21-fold) of alpha 1-hydroxylase was also observed in less differentiated CRC tissues compared to normal adjacent tissue. However, highly differentiated CRC tissues did not show significant changes in alpha 1-hydroxylase levels, indicating that the enzyme expression might vary in different stages of cancer [106]. Interestingly, a case study showed that the abnormal synthesis of the enzyme depends on the paracrine activity of cancer-associated macrophages [107]. In conclusion, the local expression of alpha 1-hydroxylase may play an important role in tumorigenesis of several cancers, which warrants further investigations.

2. Vitamin D Analogs for Cancer Treatment

Even though vitamin D has been reported to have anticancer properties, some studies have documented negative consequences, such as hypercalcemia and an imbalance in the regulation of bone metabolism due to long-term usage of vitamin D. Hypercalcemia leads to clinical manifestations and symptoms of toxicity of long-term vitamin D usage, hence limiting its potential use as a supplemental chemotherapeutic drug for extended durations. Several vitamin D derivatives were produced to tackle various issues related to vitamin D-based therapy. The structure and anticancer effects of vitamin D and its analogs are shown in Table 2. Derivatives of vitamin D have relatively low calcemic efficacy and are being investigated in clinical studies for possible anticancer applications. When compared with the naturally occurring vitamin D, several of these analogs are more selective VDR ligands [108]. In the following sections, we describe vitamin D analogs and assess their prospective application in clinics.

2.1. EB-1089 (Seocalcitol)

Compared to vitamin D, EB-1089, a synthetic analog of vitamin D, has 50–100 times greater binding capacity to VDRs and a stronger ability to suppress cancer [111]. By initiating apoptosis in breast cancer cells (MCF-7) in vitro, EB-1089 caused tumor regression [111]. The anticancer effects of EB-1089 were mediated by Bcl-2/Bax activation [112]. Moreover, EB-1089 did not stimulate hyperglycemia but induced the expression of cell cycle inhibitors p-21 and p-27 [113,114]. EB-1089 also suppressed tumor growth via increased levels of microRNA (miR-498) while downregulating the expression of the hTERT gene in ovarian cancers [115]. EB-1089 regulated the activity and expression of several transcription factors in pancreatic cancers [116]. It has been observed that EB-1089 inhibits the growth of breast cancer cells both in vitro and in vivo [113,117]. The agent was shown to block the growth of H&N SCCs [114], HCCs [118], and ovarian [88,115], pancreatic [116], and non-small cell lung [115] cancers.

2.2. HY-11

The vitamin D-based agent demonstrated effective anti-leukemic activity [119]. Its ability to retard HL-60 cancer cell growth was attributed to G1 phase cell cycle arrest in a dose-dependent way. The antiproliferative effect of this compound was also mediated by the induction of apoptosis through the caspase 3 pathway and increased expression of transforming growth factor beta (TGFβ) [119]. However, the effects of this agent in vivo were not tested.

2.3. Tacalcitol

Coded as PRI-2191 or Tacalcitol, the 1α,24(R)(OH)2D3 analog of vitamin D was tested in CRCs in combination with standard chemotherapeutic drugs [109]. The anticancer activity of PRI-2191 in combination with 5-FU (5-fluorouracil) was observed in HT29 CRC cells. The agent upregulated CDKN1A expression through VDRs, leading to an ultimate decrease in the expression of thymidylate synthase [120]. Treatment with Tacalcitol sensitized HT29 cells to 5-FU therapy, supporting the high potential of this agent as a treatment for gastrointestinal malignancies [120].

2.4. Inecalcitol

Inecalcitol (TX-522) is an epi-analog of calcitriol. It is also known as Hyrigenics paris [121,122]. Inecalcitol inhibited the growth of squamous cell and prostate cancers by inducing caspase-3, 8/10-mediated apoptosis [122]. P-27 and P-21 are the proapoptotic effectors that were activated in prostate cancer by inecalcitol. In prostate cancer cells, inecalcitol reduced the expression of Pim-1 and Ets variant-1 [123]. Inecalcitol demonstrated advanced growth-inhibitory and VDR-binding properties compared to vitamin D in animal models. However, its effects on the level of blood calcium in humans remain unknown [123,124]. Administration of inecalcitol (80 µg/day up to 3 days) inhibited the growth of SCC [124]. Another study demonstrated that inecalcitol (1300 μg/kg up to 3 days) also inhibited the proliferation of SCC and did not induce hyperglycemia [123]. When administered in combination with docetaxel, inecalcitol was well tolerated in individuals with prostate cancer; nevertheless, hypercalcemia caused dose-limiting toxicity [123]. In a phase-2 clinical study, individuals with myeloid leukemia are being treated with inecalcitol in addition to decitabine (NCT02802267).

2.5. TX527

TX527 was shown to activate the VDR/RXR complex at lower doses than vitamin D [109]. The antiproliferative effect of this agent was tested in Kaposi’s Sarcoma-associated herpesvirus GPCR-transformed endothelial cells (SVEC-vGPCR) [125]. TX527 effects in SVEC-vGPCR cells were marked by the inhibition of NF-κB expression (similar to that of vitamin D) and increased expression of IκBα. Activated IκBα decreased the localization of NF-κB in the nucleus and, thus, prevented its gene-activating effects. After receiving TX527, there was a decrease in the synthesis of inflammatory cytokines (IL-6, CCL2, and CCL20). Therefore, the agent demonstrated effective antiproliferative and anti-inflammatory properties [125], although further clinical testing is warranted.

2.6. Paricalcitol

The anticancer activities of Paricalcitol were tested in different cancer cells [126,127,128]. The agent stimulated leukemia cell HL-60 differentiation (maturation), which was marked by the elevated expression of the CD11b marker and decreased colony formation (prolonged effect after 10-day treatment) [126]. Incubation of HL-60 cells with Paricalcitol (72 h) increased apoptosis and cell cycle arrest at the G0/G1 phase [126]. Paricalcitol inhibited the growth of gastric cancer cells (AGS, MKN45, and SNU719) by inducing apoptosis, which was marked by activation of caspase 3 and decreased Bcl-2 protein expression [127]. An in vivo CRC growth study tested the combined treatment effect of 5-FU and Paricalcitol [128]. The data showed decreased tumor growth in rats that received the combination therapy compared to the monotherapy-treated animals. The combined drug treatment triggered Wnt/β-catenin, NF-kB, and COX-2 signaling cascades, which are involved in the initiation and progression of CRCs [128].

2.7. Doxercalciferol

Doxercalciferol (1α,25-(OH)2D2) is a synthetic vitamin D2 analog, which can be metabolized to form an active form of vitamin D2 in vivo. The agent was tested in combination with an arsenic compound KML001 in acute lymphoid leukemia models. The combined application demonstrated a synergistic effect and increased the number of cells in the late apoptotic phase [129]. However, the agent was not tested in other solid tumor models.

2.8. Maxacalcitol

Maxacalcitol is a non-calcemic analog of vitamin D3. The anticancer effect of this agent was studied in pancreatic cancer cells. In vitro, the analysis showed G1 phase cell cycle arrest, while in vivo, the BxPC-3 cell xenograft model showed retarded tumor growth without hypercalcemia [130]. Further analysis is warranted.

2.9. Calcipotriol

It was discovered that calcipotriol, also known as PRI-2201, inhibits the development of tumors in animal models of pancreatic cancer [93,131]. Calcipotriol administration increased stromal remodeling and promoted the uptake of the anticancer drug gemcitabine. Accordingly, tumor regression was detected in vivo in animals treated with a combination of calcipotriol. Moreover, calcipotriol alone can also inhibit tumor growth via decreased Wnt/catenin signaling [131]. The agent was not tested against other cancers.

2.10. BGP-13

BGP-13 is a synthetic calcipotriene-based analog of vitamin D3. The anticancer activity of BGP-13 was tested in pancreatic (LNCaP), breast (MCF-7), and colon cancer (HT29) cell lines [132]. The agent can induce apoptosis and cell cycle arrest at the G0/G1 phase. Apoptosis was activated by BGP-13 through the caspase-3 pathway. The agent stimulated the expression of VDRs in MCF-7 and LNCaP cells. HT29 xenografts in nude mice were also sensitive to the growth-inhibiting effects of BGP-13 [132].

2.11. PRI-2205

PRI-2205 was tested for its anticancer activities in vitro and in vivo [133]. Combined treatment with cisplatin showed a promising cytostatic potential of this agent. In combination with tamoxifen, PRI-2205 induced G2/M phase cell cycle arrest in MCF-7 breast cancer cells [134]. Another study reported a strong reduction in mRNA levels of estrogen receptors (ERs) in MCF-7 cells treated with a combination of PRI-2205 and anastrozole (an aromatase inhibitor) [133]. The combined treatment showed the reduced expression of other genes involved in the ER signaling, including estrogen-related receptor alpha (ESRRα), estrogen-related receptor gamma (ESRRγ), and Gonadotropin-releasing hormone (GnRH1) [133]. Therefore, the agent provoked substantial interest as a potential regulator of gene expression in cancer cells.

2.12. PRI-1906

PRI-1906 stimulated an increased mRNA expression of CYP24A1 in patient-derived ovarian cancer cells. The agent increased the level of nuclear VDRs, which led to decreased cell viability [135]. Further testing of this agent in different cancer models is warranted.

2.13. BXL-01-0126

BXL-01-0126 belongs to the C20 Gemini class of vitamin D3 analogs. It exhibits better anticancer activity than vitamin D3, although similar hypercalcemic properties were registered [136]. In AML cells, the agent increased cAMP levels in a dose-dependent manner. Similar results have been observed in vivo using a xenograft model. The agent’s ability to regulate cAMP levels in AML patients may protect them from microbial infections during chemotherapy [136].

2.14. BXL0124

BXL0124 has demonstrated anticancer activity in different cancer models in vivo, including breast cancer xenograft models [137,138,139]. BX1024 blocked the local invasion of tumor cells in the preclinical testing [140]. However, the mechanism of BXL0124’s anti-metastatic effects remains unclear and warrants additional investigations.

2.15. Gemini0097

Gemini0097 belongs to the family of the C20 class of vitamin D3 analogs. It is a less toxic and more efficient anticancer substance than vitamin D3 [141]. The agent was tested in vivo using an ER-positive breast cancer model with chemically induced tumors (N-methyl-N-nitrosourea (NMU)-induced cancer). Gemini0097 was the most potent agent among the tested substances and caused up to 60% growth inhibition of NMU-induced tumors. The agent also increased the expression of p21 (CDK inhibitor) and insulin-like growth factor binding protein 3 (IGFBP3) in ER+ and ER tumors [142].

2.16. MART-10

MART-10 is a potent anticancer agent without any serious side effects. The agent was tested in MDA-MB-231 (triple-negative breast cancer (TNBC) cell line) [143] and decreased the migratory and invasive potential of these cells. These effects were mediated by the upregulated E-cadherin and downregulated N-cadherin proteins. MMP9 downregulation was also observed upon treatment with MART-10. These results support the promising potential of this substance for the treatment of TNBC [143]. However, further investigation is required to ensure its effectiveness in vivo.

2.17. 1,25-Dihydroxyvitamin D3-3-Bromoacetate

The anticancer effect of 1,25(OH)2D3-3-BE was evaluated in the human renal tumor (A498 and Caki1) xenograft model. The agent was a more effective inhibitor of tumor growth compared to vitamin D3 in that study [144]. It also effectively inhibited cell cycle progression via enhanced caspase activity, decreased expression of cyclin A, and the inhibition of Akt phosphorylation [144].

2.18. Ro26-2198

Ro26-2198 delayed the onset of colitis in an azoxymethane (AOM)-induced CRC model in mice. The anticancer property of this agent in CRC was attributed to the decreased expression of c-Myc, COX2, and pERK (oncogenes) in vivo. In vitro models confirmed the observed mechanism of Ro26-2198 effects. Moreover, the agent decreased the expression of IL-1B, suggesting a possible role in the regulation of inflammation [145].

2.19. EM1

EM1, the more stable analog of vitamin D, demonstrated antiproliferative properties in vitro in a variety of cancer cell types [146,147]. The administration of EM1 demonstrated antitumor activity in animal models with various cancer xenografts, including breast carcinomas. The administration of EM1 inhibited metastasis by E-cadherin expression [148].

3. Molecular Mechanisms of Tumor Growth Inhibition by Vitamin D

3.1. Molecular Mechanisms of Vitamin D-Induced Apoptosis

Morphologic changes associated with apoptosis include cell shrinkage, blebbing of the plasma membrane, release of mitochondrial cytochrome c, fragmentation of the cellular DNA into multiples of 180 bp, and, finally, the breaking of the cell into small apoptotic bodies that will be cleared by nearby cells through phagocytosis. The activation of caspases (cysteine aspartate-specific proteases) is known to cause these morphologic alterations [149]. Apoptosis was indicated as an important mechanism of anticancer effects induced by vitamin D and its analogs [150]. However, there are several conflicting results on vitamin D-induced apoptosis. For instance, one study demonstrated that vitamin D induces apoptosis via a caspase-dependent mechanism [151], while others reported activation of caspase-independent cascades [152,153]. In MCF-7 cells, vitamin D induces apoptosis via a caspase-independent mechanism by activating mitochondrial malfunction, translocating Bax, and generating reactive oxygen species (ROS) [153]. But a separate study reported that vitamin D3 induces apoptosis in breast cancer cell lines (MCF-7 and in MDA-MB-231) by triggering Caspase-3/7 [154]. The vitamin D analog EB-1089 induced apoptosis in MCF-7 cells via elevating intracellular calcium levels, which activated μ-calpain [155]. The activation of autophagy (as evidenced by enhanced beclin-1 expression) by vitamin D was also reported in certain studies, which was further linked to the upregulation of apoptosis [156]. By increasing the proapoptotic BAK gene and downregulating the antiapoptotic proteins Bcl-2 and IAP, vitamin D induces apoptosis in CRC cells [157]. Furthermore, vitamin D can induce apoptosis in prostate cancer cell lines (LNCaP and ALVA-31) by activating mitochondria-related apoptotic pathways [151]. A downregulation of telomerase (hTERT), an activation of caspase-dependent MEK cleavage, and the overexpression of p53 are the other key mechanisms involved in vitamin D-induced apoptosis [83,158].

3.2. Antiproliferative Mechanisms of Vitamin D

Deregulated proliferation is one of the hallmarks of cancer cells [159]. Cancer cells not only generate specific growth factor production in stromal cells but also alter growth factor signaling pathways [159]. Notably, vitamin D-induced effects were shown to trigger numerous signaling mechanisms that stimulate growth inhibition and apoptosis. Previous research has demonstrated that vitamin D activates p-21waf1/cip1 and induces G0/G1 cell cycle arrest, which results in antitumor actions in breast and prostate cancer cells [160]. VDRs and VDRE directly target P-21waf1/cip1 [161,162]. By increasing the expression of p21waf/cip1 and p-27kip1, as well as by lowering the levels of cyclins and cyclin-dependent kinases (CDKs), vitamin D treatment inhibits pancreatic cancer cells [163]. Another study discovered that vitamin D administration causes squamous cell carcinomas to produce more p-27kip1 and p21wf/cip1, which in turn causes G0/G1 phase cell cycle arrest [164].
Vitamin D treatment inhibited the growth of ovarian cancer cells by inducing G2/M phase arrest, which is triggered by transcriptional activity that is independent of p53 [165]. Furthermore, vitamin D administration could cause cell cycle arrest by lowering ERK1/2 expression and activity [166]. The vitamin D analog EB-1089 also showed growth-inhibitory potential. The effect was mediated by the induction of PTEN, which inhibited the Akt pathway in thyroid carcinoma cells [167]. In animal models of thyroid carcinogenesis, vitamin D inhibited PIK3/Akt-associated cell proliferation. [168]. Vitamin D and its analog also demonstrated antitumor activity in kidney cancer cells, which was mediated by the inhibition of Akt and its target gene caspase-9 [144]. Furthermore, vitamin D treatment arrested the growth of endothelial cells by inhibition of NF-kB activation, a mechanism independent of PI3K/Akt and MAPK pathways [169]. Growth-inhibitory effects of vitamin D treatment in MCF-7 cells were mediated by Src-tyrosine kinase and tyrosine phosphatase activation and the phosphorylation of ERK1/2 [170]. Another study reported that vitamin D inhibits the proliferation of breast cancer cells via activating IGFBP3, a target of the VDR and a functional activator of VDRE [171]. Vitamin D treatment elevates TGF-β production, which is another potent regulator of proliferation and apoptosis in breast cancer cells [172]. Vitamin D treatment retarded the proliferation of CRC cell line SW-480-ADH via upregulation of microRNA-22 (miR-22) expression [173]. Increased levels of let-7a-2 (anticancer miR) were reported in A-549 cells [174]. The effects were associated with the activation of VDRE [174]. However, many of the reported effects were observed in vitro and require further clinical confirmation.

3.3. Vitamin D Inhibits Key Events in the Metastatic Spread of Cancer Cells

Metastasis is a complex process by which cancer cells spread from the original tumor to form a new tumor in other parts of the body. Cell adhesion, invasion, migration, and re-establishment are the key steps involved in metastasis [175]. The spread of cancer to distant organs is responsible for almost 90% of cancer-related fatalities [176]. Shreds of evidence reported that vitamin D and its analogs modulate several crucial steps of metastasis. For instance, it has been demonstrated that EB-1089 activates fibronectin in thyroid cancer cells in a PTEN-dependent manner, thereby restoring cell adhesion [177]. In squamous cell carcinoma (SCC), vitamin D was shown to inhibit tumor cell motility via modulating E-cadherin expression [178]. Vitamin D inhibited the invasion and metastasis of cancer cells by downregulating the secretion of MMP-2 and MMP-9 [178]. The invasion of prostate cancer was reduced by vitamin D through the reduction in MMP-9 and cathepsins secretion, while MMP-1 (a tissue inhibitor) was activated [179].
The metastatic potential of prostate cancer cells was reduced by vitamin D, which decreased the adhesion and rolling activity of cells in the vasculature [180,181]. In Lewis lung carcinoma cell lines (LLC and LLC-LN7), vitamin D treatment blocked the invasion and migration by decreasing the production of GM-CSF, which was associated with reduced activity of PKA [182,183]. Other mechanisms of inhibition were associated with the downregulation of Dickkopf 4, a protein-coding gene, which inhibits the Wnt/β-catenin signaling and MAPK pathways [184]. By inhibiting SNAI1, Slug, and vimentin as well as the EMT, the vitamin D analog MART-10 reduced the migration of pancreatic cells BXPC-3 and PANC [185]. Interestingly, vitamin D facilitated cell motility in normal cells by the specific activation of PI3K [186].
Vitamin D treatment also inhibited the migration and invasion of cancer cells in animal models [187]. In a melanoma B-16 mouse model, vitamin D (daily at a dose of 0.5 µg/kg for 28 days) decreased both experimental and spontaneous pulmonary metastases [187]. Vitamin D analog 22-oxa decreased the cancer colony formation in the lungs by negatively regulating fibroblast-induced angiogenesis [188]. In mice models of breast cancer, another vitamin D analog, EB-1089, had inhibitory effects on bone metastasis [189]. Larger studies are warranted to confirm the observed anti-metastatic effects of vitamin D analogs.

3.4. Vitamin D Induces Cancer Cell Differentiation

Cancer cells have lost the ability to differentiate. This characteristic allows the malignant cells to grow faster and spread quicker compared to normal/differentiated cells [190]. Acute myeloid leukemia (HL-60) cells treated with vitamin D were seen to exhibit a mature monocyte phenotype, suggesting that vitamin D can induce cell differentiation [191]. The ability to differentiate upon exposure to vitamin D was marked by cell cycle arrest and elevated p-21waf1/cip-1 and p-27kip-1 [191]. In human monocytic leukemia cells, vitamin D promoted differentiation through the inhibition of p-38 MAP-kinase or activation of PI3K [192,193]. However, another study reported the involvement of JNK and MAPK signaling cascades in vitamin D-mediated differentiation of HL-60 and U-937 cells [194]. Vitamin D activated ERK1/2 in HL-60 cells, which in turn upregulated the expression of C/EBPβ and c-JUN [195]. Enhanced nuclear localization of C/EBPβ-2 and C/EBPβ-3 promoted the expression of Ras-1, a well-known differentiation trigger [196].
Earlier reports stated that vitamin D treatment promotes epithelial carcinoma differentiation through the activation of alkaline phosphatase and AP-1 protein [197]. Vitamin D inhibited cell migration by increasing the expression of E-cadherin and decreased the levels of vimentin, SNAI1, and ZEB1 [198]. For example, vitamin D suppressed the expression of mesenchymal markers (N-cadherin, p-cadherin, and integrins α-6 and β-4) but increased the expression of claudin-7 and β5 FAK in MDA-MB-453 breast cancer cells [199]. Vitamin D analog 1α-(OH)D5 showed the differentiation of breast cancer cell line T47D by increasing lipid production and casein expression [200]. Vitamin D’s ability to stimulate differentiation in MCF-7 and T47D mammary cancer cell lines was mediated not only by cell cycle arrest but also by reducing the ability to produce anchorage-independent colonies [200,201]. Vitamin D activated the pro-differentiation-related gene ICB-1, which subsequently increased the expression of E-cadherin in MDA-MB-231 breast cancer cells [202]. Vitamin D was also found to upregulate the expression of androgen receptor (AR) and PSA proteins, which can regulate the differentiation process [203]. However, the role of vitamin D analogs in the regulation of cell differentiation remains unclear and warrants further investigation in vivo and in clinical studies.

3.5. Vitamin D Inhibits Angiogenesis and Constrains Tumor Growth

According to earlier observations, solid tumors are incapable of expanding their diameter beyond 2.0 mm if they are unable to induce the growth of new blood vessels [204]. For malignant cells to survive and proliferate, an abundant supply of nutrients and oxygen is necessary, for which tumor cells promote the formation of blood vessels (angiogenesis). The process is crucial for tumor progression, metastasis, and the development of drug resistance [205,206,207,208]. Vitamin D was shown to influence angiogenesis via multiple mechanisms. For instance, TX-527 (vitamin D analog) downregulated the expression of G protein-coupled receptors (GPCRs), which led to the inhibition of murine endothelial angiogenesis [209]. The agent significantly inhibited vascular GPCR-related tumor progression in vivo [209]. Vitamin D treatment suppressed the expression and signaling of vascular endothelial growth factor (VEGF), thus triggering apoptosis and inhibition of vascular cell elongation [210]. In SCC models, the proliferation of tumor-derived endothelium cells was also inhibited by other vitamin D analogs, including EB-1089, -6760, and -7553 [206].
Notably, apoptosis and cell cycle arrest in the G0/G1 phase were detected in tumor-derived endothelial cells (TDECs) exposed to vitamin D, while no significant death-promoting effects were observed in specific matrigel-derived endothelial cells [208]. Vitamin D treatment resulted in the increased expression of p-27kip-1, while Akt and ERK1/2 activities were downregulated [166]. Interestingly, no inhibition of vascular cell growth was observed in the experiments with embryonic chorioallantoic membranes (CAMs) exposed to vitamin D. However, numerous reports on cell-based and animal models have supported the anti-angiogenic effects of vitamin D and its analogs [211]. For instance, vitamin D blocked vascular growth in xenografted breast cancer models [210]. Topical administration of vitamin D also demonstrated anti-angiogenic effects in a mouse model with suture-induced corneal inflammation [212]. Treatment with vitamin D and its analog 22-oxa-1,25D3 decreased MMP-2, MMP-9, and VEGF levels [213]. In prostate cancer tumor models, vitamin D treatment was reported to downregulate MMP-9 and IL-8 levels, as well as inhibit the migration and tube formation in human umbilical vein endothelial cells (HUVECs) [214] (Figure 3).

4. VDR Gene Polymorphism in Cancers

Numerous biological processes, such as immunological responses, bone metabolism, specific cell proliferation, and differentiation in healthy tissues, are regulated by vitamin D. It has been demonstrated that during carcinogenesis, vitamin D regulates the genes involved in inhibiting the development, migration, adhesion, and angiogenesis of cancer cells [77,187,215,216,217,218,219,220,221]. Considering that VDRs mediate vitamin D biological activities, variations in VDRs can modulate the function of vitamin D [222,223]. Notably, the VDR-coding gene is known to contain about 200 single nucleotide polymorphs (SNPs) [224]. Polymorphic VDRs may have an impact on a person’s vitamin D levels. The alteration in VDR signaling is caused by SNPs in the VDR gene [225,226]. Fok1 (rs10735810), Bsm1 (rs1544410) [227,228], Apa1 (rs7975232), Taq1 (rs731236), and Cdx2 (rs11568820) [229] genes are among the most extensively studied VDR SNPs associated with cancer.
Although further research is necessary to fully comprehend the role of VDR polymorphisms in different cancers, genetic VDR variations are critical markers used for the selection of better treatment strategies. For instance, the results of a meta-analysis indicated that VDR polymorphisms could be associated with a greater risk of keratinocyte cancers [230]. The SNP in the Cdx2 (Caudal-type homeobox protein 2, which is an intestine-specific transcription factor with a polymorphic binding site in the VDR gene) has been associated with an overall increased risk of cancer [229]. The Taq1-targeting VDR polymorphism was linked to an increased risk of colorectal cancer [229]. However, no correlation was observed between prostate cancer and polymorphisms in the Apa1 and Cdx2 VDR genes [231]. However, to support this conclusion, additional evidence is needed [231].
Another study suggested that the Fok1 VDR polymorphism might be a viable target to predict the risk of prostate cancer [232]. According to an updated investigation, there is a possibility that the Fok1 VDR polymorphism raises the risk of prostate cancer in Caucasians. To verify these results, more population-based studies are warranted to confirm these data [233]. On the other hand, a research investigation of a subset of Caucasians revealed no correlation between the risk of breast cancer and the allele contrast for the Apa1, Fok1, Tag1, and Bsm1 VDR gene polymorphisms [234]. However, the Fok1 VDR polymorphism was associated with ovarian and breast cancers in another study [235]. Liu et al. (2017) showed a substantial correlation between Asian and African American men’s prostate cancer risk and the rs731236 VDR [236]. The development of breast tumors may be influenced by polymorphisms in the Apa1, Bsm1, Fok1, and Poly(A) VDR genes, according to a different systematic review and meta-analysis [237]. Further studies are warranted to confirm this conclusion.
Another study reported an absence of correlation between TaqI polymorphisms and susceptibility to CRC [238]. Among the Asian population (particularly Japanese people), the Taq1 polymorphism was linked to an increased incidence of prostate cancer [239]. Prostate cancer that progressed to an advanced stage was more common in those individuals carrying the T allele or TT genotype. Consequently, VDR-linked Taq1 polymorphisms might be regarded as a possible diagnostic biomarker for the susceptibility to prostate cancer [239]. Despite the need for more research, a recent study found that VDR polymorphism is likely to increase the risk of lung cancer [240].
The Fok1 polymorphism was correlated with the type and severity of CRCs [241]. It has been shown that Fok1 and Bsm1 are risk factors for CRC [242]. Moreover, a meta-analysis study indicated that the polymorphism of Fok1 may be linked to ovarian cancer in Caucasian populations [243]. However, the hypothesis was not confirmed in another study, which reported that the Bsm1 polymorphism is likely the best risk indicator of ovarian cancer in Caucasian patients [244]. The risk of breast and ovarian cancers was associated with Fok1 but not with Apa1 (rs7975232), Cdx2 (rs11568820), and Taq1 (rs731236) VDR polymorphisms, as indicated by a meta-analysis of published data [245]. Interestingly, the Bsm1 polymorphism was linked with a lower incidence of these malignancies [245]. Asian populations are more susceptible to renal cell cancer (RCC) due to the VDR gene polymorphisms Apa1 and Fok1 (FF genotype) [246]. According to a meta-analysis, polymorphisms in Bsm1, Cdx2, and Taq1 increase the risk of developing lung cancer [247]. Accordingly, the VDR Taq1 polymorphism was linked to an increased risk of malignancies related to tobacco usage and smoking [248].
A study suggested that Apa1, Bsm1, and Fok1 polymorphisms might influence the development of melanoma [249]. However, the presence of specific Cdx2 and Bsm1 variants was associated with a lower risk of lung cancer, while Taq1 was associated with an increased risk of this cancer. The presence of AA genotypes of Bsm1 and Apa1 variants was considered to be protective against lung cancers, whereas Taq1 and Fok1 polymorphisms were estimated as risk factors in Asian populations for this type of tumor [250].
VDR polymorphism (Apa1, Cdx2, and Taq1) is associated with an increased risk of developing various cancer types, including CRC [229], BCC, SCC, HCC, head and neck cancers, kidney cancers, and thyroid cancer. VDR gene RFLPs were also associated with an increased risk of 19 different types of cancers [251]. Therefore, strong associations were shown for breast (Apa1, Bsm1, Cdx2, Fok1, and Taq1), colorectal (Apa1, Bsm1, Fok1, and Taq1), prostate (Apa1, Bsm1, Cdx2, Fok1, and Taq1), and skin (Bsm1, Fok1, and Taq1) cancers [251]. Future research should focus on combining VDR-specific genetic polymorphisms with the measurement of vitamin D levels, with ethnicity as a stratum for the study.
Consequently, a comprehensive meta-analysis assessed the possibility of using VDR polymorphisms in diagnostics. After analyzing the data collected from 192 independent studies (98,209 cancer-free controls and 78,628 cases), [252] it was shown that Fok1, Bsm1, Cdx2, Apa1, and Taq1 are good markers for CRC, lung, ovarian, skin, multiple myeloma, and brain tumors [252]. When compared to other ethnic groups, Caucasians showed the strongest correlation. To determine the important correlations and causal relationships between VDR polymorphisms and particular cancer types, more research with larger cohorts is warranted. Studies depicting the role of VDR polymorphisms in various cancers are shown in Table 3.

5. Conclusions and Future Directions

Vitamin D has been recognized as a powerful anticancer agent because of its numerous inhibitory actions on tumor cells. Although several studies have demonstrated the anticancer effects of vitamin D and its analogs in lab-based and animal models, further studies are still required. Several controversies and concerns must be resolved before vitamin D can be considered for cancer treatment. The question “Is vitamin D a good cancer prevention or treatment agent?” is yet to be answered. Consequently, apart from deciphering the processes via which vitamin D and its analogs impede the proliferation of cancer cells, future studies should focus on why vitamin D by itself and in conjunction with other vitamins, such as vitamin E, has failed in reducing cancer incidence. The development of optimized versions of vitamin D and vitamin D carriers with cancer-cell-targeting capabilities should be the main emphasis of future research.

Author Contributions

Conceptualization, supervision, review, editing, and proofreading: S.M.N., R.R., P.V.S., E.T. and S.V.M. Review of literature, data collection, and writing original draft: S.D., S.V.T., V.G.B., M.K., C.A.U., P.G.A. and O.A.S. All authors have read and agreed to the published version of the manuscript.

Funding

The study was supported by Department of Biotechnology (DBT), Government of India, as a part of extramural research grant (BT/PR29598/PFN/20/1392/2018) awarded to SubbaRao V. Madhunapantula. The article processing charges (APC) were in part supported by JSS Academy of Higher Education & Research (JSS AHER), Mysuru, Karnataka, India.

Acknowledgments

S.D. would like to thank DBT, Government of India, for providing SRF as a part of grant support (BT/PR29598/PFN/20/1392/2018 to SubbaRao V. Madhunapantula). S.D. would like to thank the ICMR, Government of India, for the award of SRF (3/2/2/84/2022-NCD-III). S.V.T. would like to thank ICMR, Government of India, for providing SRF as a part of an extramural research grant (54/06/2019-HUM/BMS dated 22 November 2019) sanctioned to Suma M. Nataraj. M.K. would like to acknowledge CII-SERB for awarding the Prime Minister’s Fellowship (SERB/PM Fellow/CII-FICCI/Meeting/October/2021 dated 8 December 2021) for Doctoral Research and Novick Biosciences Pvt. Ltd. (Hyderabad, Telangana, India) for being a partner of the Prime Ministers Fellowship. P.G.A. would like to thank DST for providing the Senior Project Associateship as a part of the PURSE grant awarded to JSS AHER (SR/PURSE/2021/81 (c), dated 25 March 2022). S.V.M. would like to thank SIG-CBCSC, JSS Medical College, JSS AHER (Mysore, Karnataka, India). S.V.T., S.D., C.A.U., M.K., V.G.B., P.G.A., S.M.N., and S.V.M. would like to acknowledge the DST-FIST grant to the CEMR Laboratory (CR-FST-LS-1/2018/178) and to the Department of Biochemistry (SR/FST/LS-1-539/2012), DBT-BUILDER program (BT/INF/22/SP43045/2021), DST-PURSE (SR/PURSE/2021/81 (c)). P.V.S. and V.G.B. would like to thank SERB for providing the EMR grant (EMR/2017/001754). C.A.U. would like to acknowledge DST-CSRI for providing SRF as a part of grant support (SR/CSRI/44/2016(G)) to Rajalakshmi R.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

AF2Activation function 2
AktProtein kinase B
AMLAcute myeloid leukemia
AOMAzoxymethane
ARAndrogen receptor
BaxBcl-2 Associated X-protein
BCCBasal cell carcinoma
Bcl-2B-cell lymphoma 2
C/EBPCCAAT/enhancer binding protein
CAMChick chorioallantoic membrane
CD44Cluster of differentiation 44
CDKsCyclin-dependent kinases
CKIICasein kinase II
COX-2Cyclooxygenase-2
CRCColorectal cancer
EMTEpithelial-mesenchymal transition
EP2Prostaglandin E2 receptor
ERαEstrogen receptor alpha
ERKExtracellular signal-regulated kinase
ESRRαEstrogen-related receptor alpha
ESRRγEstrogen-related receptor gamma
FAKFocal adhesion kinase
GM-CSFGranulocyte-macrophage colony-stimulating factor
GnRHGonadotropin-releasing hormone
GPCRG protein-coupled receptor
HCCHepatocellular carcinoma
HERHuman epidermal growth factor receptor
hTERTHuman telomerase reverse transcriptase
HUVECsHuman umbilical vein endothelial cells
IGFBP3Insulin-like growth factor binding protein 3
ILInterleukin
IUInternational units
JNKc-Jun N-terminal kinase
KPSKarnofsky Performance Scale
MAPKMitogen-activated protein kinase
MARRSMembrane-associated, rapid response steroid-binding
MEKMitogen-activated protein kinase kinase
MMPMatrix metalloproteinase
MycMyelocytomatosis
NCoA62–SKIPNuclear coactivator-62 kDa/Ski-interacting protein
NF-κBNuclear factor kappa B
NMUN-methyl-N-nitrosourea
PARGPoly(ADP-ribose) glycohydrolase
PGDH15-Hydroxyprostaglandin dehydrogenase
PI3KPhosphoinositide 3-kinase
PKCProtein kinase C
PSAProstate-specific antigen
PTENPhosphatase and tensin homolog deleted on chromosome 10
PTHParathyroid hormone
RCCRenal cell carcinoma
RCTsRandomized controlled trials
RDARecommended dietary allowance
ROSReactive oxygen species
RXRRetinoic acid receptor
SCCSquamous cell carcinoma
SMRTSilencing mediator for retinoid and thyroid hormone receptors
SNAI1Snail family transcriptional repressor-1
SNPSingle nucleotide polymorphism
SRCsSteroid receptor coactivators
SVEC-vGPCRSarcoma-associated herpesvirus GPCR-transformed endothelial cells
Tcf-1T-cell factor 1
TDECsTumor-derived endothelial cells
TGFβTransforming growth factor-β
TNBCTriple-negative breast cancer
VDBPVitamin D-binding protein
VDRVitamin D receptor
VDREVitamin D-response elements
VEGFVascular endothelial growth factor
ZEB1Zinc finger E-box binding homeobox-1

References

  1. Stamm, E.; Acchini, A.; Da Costa, A.; Besse, S.; Christou, F.; Launay, C.; Balmer, P.; Hamart, M.; Nguyun, S.; Major, K.; et al. Year in review: Geriatrics. Rev. Medicale Suisse 2019, 15, 50–52. [Google Scholar] [CrossRef]
  2. Tang, H.; Li, D.; Li, Y.; Zhang, X.; Song, Y.; Li, X. Effects of vitamin D supplementation on glucose and insulin homeostasis and incident diabetes among nondiabetic adults: A meta-analysis of randomized controlled trials. Int. J. Endocrinol. 2018, 2018, 7908764. [Google Scholar] [CrossRef]
  3. Fink, C.; Peters, R.L.; Koplin, J.J.; Brown, J.; Allen, K.J. Factors affecting vitamin D status in infants. Children 2019, 6, 7. [Google Scholar] [CrossRef]
  4. Zhang, R.; Naughton, D.P. Vitamin D in health and disease: Current perspectives. Nutr. J. 2010, 9, 65. [Google Scholar] [CrossRef]
  5. Gupta, D.; Vashi, P.G.; Trukova, K.; Lis, C.G.; Lammersfeld, C.A. Prevalence of serum vitamin D deficiency and insufficiency in cancer: Review of the epidemiological literature. Exp. Ther. Med. 2011, 2, 181–193. [Google Scholar] [CrossRef]
  6. Keum, N.; Giovannucci, E. Vitamin D supplements and cancer incidence and mortality: A meta-analysis. Br. J. Cancer 2014, 111, 976–980. [Google Scholar] [CrossRef]
  7. Zhang, R.; Zhang, Y.; Liu, Z.; Pei, Y.; Xu, P.; Chong, W.; Hai, Y.; He, L.; He, Y.; Yu, J.; et al. Association between Vitamin D Supplementation and Cancer Mortality: A Systematic Review and Meta-Analysis. Cancers 2022, 14, 3717. [Google Scholar] [CrossRef]
  8. Amiri, M. Vitamin D and cancer; a contradictory problem. Immunopathol. Persa 2018, 4, e13. [Google Scholar] [CrossRef]
  9. Bouillon, R.; Carmeliet, G.; Verlinden, L.; Van Etten, E.; Verstuyf, A.; Luderer, H.F.; Lieben, L.; Mathieu, C.; Demay, M. Vitamin D and human health: Lessons from vitamin D receptor null mice. Endocr. Rev. 2008, 29, 726–776. [Google Scholar] [CrossRef]
  10. Chen, J.; Wang, J.; Kim, T.-K.; Tieu, E.W.; Tang, E.K.; Lin, Z.; Kovacic, D.; Miller, D.D.; Postlethwaite, A.; Tuckey, C.R.; et al. Novel vitamin D analogues as potential therapeutics: Metabolism, toxicity profiling, and antiproliferative activity. Anticancer Res. 2014, 34, 2153–2163. [Google Scholar]
  11. Haussler, M.R.; Whitfield, G.K.; Haussler, C.A.; Hsieh, J.C.; Thompson, P.; Selznick, S.H.; Dominguez, C.E.; Jurutka, P.W. The nuclear vitamin D receptor: Biological and molecular regulatory properties revealed. J. Bone Miner. Res. 1998, 13, 325–349. [Google Scholar] [CrossRef]
  12. Hsieh, J.C.; Shimizu, Y.; Minoshima, S.; Shimizu, N.; Haussler, C.A.; Jurutka, P.W.; Haussler, M.R. Novel nuclear localization signal between the two DNA-binding zinc fingers in the human vitamin D receptor. J. Cell. Biochem. 1998, 70, 94–109. [Google Scholar] [CrossRef]
  13. Michigami, T.; Suga, A.; Yamazaki, M.; Shimizu, C.; Cai, G.; Okada, S.; Ozono, K. Identification of amino acid sequence in the hinge region of human vitamin D receptor that transfers a cytosolic protein to the nucleus. J. Biol. Chem. 1999, 274, 33531–33538. [Google Scholar] [CrossRef]
  14. Yasmin, R.; Williams, R.M.; Xu, M.; Noy, N. Nuclear import of the retinoid X receptor, the vitamin D receptor, and their mutual heterodimer. J. Biol. Chem. 2005, 280, 40152–40160. [Google Scholar] [CrossRef]
  15. Uthaiah, C.A.; Beeraka, N.M.; Rajalakshmi, R.; Ramya, C.M.; Madhunapantula, S.V. Role of Neural Stem Cells and Vitamin D Receptor (VDR)-Mediated Cellular Signaling in the Mitigation of Neurological Diseases. Mol. Neurobiol. 2022, 59, 4065–4105. [Google Scholar] [CrossRef]
  16. Hsieh, J.-C.; Jurutka, P.W.; Galligan, M.A.; Terpening, C.M.; Haussler, C.A.; Samuels, D.S.; Shimizu, Y.; Haussler, M.R. Human vitamin D receptor is selectively phosphorylated by protein kinase C on serine 51, a residue crucial to its trans-activation function. Proc. Natl. Acad. Sci. USA 1991, 88, 9315–9319. [Google Scholar] [CrossRef]
  17. Jurutka, P.; Hsieh, J.C.; MacDonald, P.N.; Terpening, C.M.; Haussler, C.; Haussler, M.R.; Whitefield, G.K. Phosphorylation of serine 208 in the human vitamin D receptor. The predominant amino acid phosphorylated by casein kinase II, in vitro, and identification as a significant phosphorylation site in intact cells. J. Biol. Chem. 1993, 268, 6791–6799. [Google Scholar] [CrossRef]
  18. Christakos, S.; Dhawan, P.; Verstuyf, A.; Verlinden, L.; Carmeliet, G. Vitamin D: Metabolism, molecular mechanism of action, and pleiotropic effects. Physiol. Rev. 2016, 96, 365–408. [Google Scholar] [CrossRef]
  19. Heaney, R.P. Vitamin D in health and disease. Clin. J. Am. Soc. Nephrol. 2008, 3, 1535–1541. [Google Scholar] [CrossRef]
  20. Jäpelt, R.B.; Jakobsen, J. Vitamin D in plants: A review of occurrence, analysis, and biosynthesis. Front. Plant Sci. 2013, 4, 136. [Google Scholar] [CrossRef] [PubMed]
  21. Jones, G.; Prosser, D.E.; Kaufmann, M. Cytochrome P450-mediated metabolism of vitamin D. J. Lipid Res. 2014, 55, 13–31. [Google Scholar] [CrossRef] [PubMed]
  22. Schuster, I. Cytochromes P450 are essential players in the vitamin D signaling system. Biochim. Biophys. Acta (BBA)-Proteins Proteom. 2011, 1814, 186–199. [Google Scholar] [CrossRef]
  23. Takeyama, K.-I.; Kitanaka, S.; Sato, T.; Kobori, M.; Yanagisawa, J.; Kato, S. 25-Hydroxyvitamin D3 1α-hydroxylase and vitamin D synthesis. Science 1997, 277, 1827–1830. [Google Scholar] [CrossRef]
  24. Brenza, H.L.; DeLuca, H.F. Regulation of 25-hydroxyvitamin D3 1α-hydroxylase gene expression by parathyroid hormone and 1, 25-dihydroxyvitamin D3. Archives of Biochemistry and Biophysics. Arch. Biochem. Biophys. 2000, 381, 143–152. [Google Scholar] [CrossRef]
  25. Hewison, M.; Zehnder, D.; Bland, R.; Stewart, P. 1alpha-Hydroxylase the action of vitamin, D. J. Mol. Endocrinol. 2000, 25, 141–148. [Google Scholar] [CrossRef]
  26. Murayama, A.; Takeyama, K.-I.; Kitanaka, S.; Kodera, Y.; Hosoya, T.; Kato, S. The promoter of the human 25-hydroxyvitamin D31α-hydroxylase gene confers positive and negative responsiveness to PTH, calcitonin, and 1α, 25 (OH) 2D3. Biochemical and biophysical research communications. Biochem. Biophys. Res. Commun. 1998, 249, 11–16. [Google Scholar] [CrossRef]
  27. Deeb, K.K.; Trump, D.L.; Johnson, C.S. Vitamin D signalling pathways in cancer: Potential for anticancer therapeutics. Nat. Rev. Cancer 2007, 7, 684–700. [Google Scholar] [CrossRef]
  28. Haussler, M.R.; Jurutka, P.W.; Mizwicki, M.; Norman, A.W. Vitamin D receptor (VDR)-mediated actions of 1α, 25 (OH) 2vitamin D3: Genomic and non-genomic mechanisms. Best Pract. Res. Clin. Endocrinol. Metab. 2011, 25, 543–559. [Google Scholar] [CrossRef]
  29. Haussler, M.R.; Whitfield, G.K.; Kaneko, I.; Haussler, C.A.; Hsieh, D.; Hsieh, J.-C.; Jurukta, P.W. Molecular mechanisms of vitamin D action. Calcif. Tissue Int. 2013, 92, 77–98. [Google Scholar] [CrossRef] [PubMed]
  30. Sannappa Gowda, N.G.; Shiragannavar, V.D.; Puttahanumantharayappa, L.D.; Shivakumar, A.T.; Dallavalasa, S.; Basavaraju, C.G.; Bhat, S.S.; Prasad, S.K.; Vamadevaiah, R.M.; Madhunapantula, S.V.; et al. Quercetin activates vitamin D receptor and ameliorates breast cancer induced hepatic inflammation and fibrosis. Front. Nutr. 2023, 10, 1158633. [Google Scholar] [CrossRef] [PubMed]
  31. Pike, J.W.; Meyer, M.B. Fundamentals of vitamin D hormone-regulated gene expression. J. Steroid Biochem. Mol. Biol. 2014, 144, 5–11. [Google Scholar] [CrossRef] [PubMed]
  32. Dusso, A.S.; Thadhani, R.; Slatopolsky, E. (Eds.) Vitamin D receptor and analogues. In Seminars in Nephrology; Elsevier: Amsterdam, The Netherlands, 2004. [Google Scholar]
  33. Nemere, I.; Safford, S.E.; Rohe, B.; DeSouza, M.M.; Farach-Carson, M.C. Identification and characterization of 1, 25D3-membrane-associated rapid response, steroid (1, 25D3-MARRS) binding protein. J. Steroid Biochem. Mol. Biol. 2004, 89, 281–285. [Google Scholar] [CrossRef] [PubMed]
  34. Giammanco, M.; Di Majo, D.; La Guardia, M.; Aiello, S.; Crescimannno, M.; Flandina, C.; Tumminello, F.M.; Leto, G. Vitamin D in cancer chemoprevention. Pharm. Biol. 2015, 53, 1399–1434. [Google Scholar] [CrossRef]
  35. Zeeb, H. Vitamin D and cancer prevention. Curr. Nutr. Rep. 2012, 1, 24–29. [Google Scholar] [CrossRef]
  36. Muñoz, A.; Grant, W.B. Vitamin D and Cancer: An Historical Overview of the Epidemiology and Mechanisms. Nutrients 2022, 14, 1448. [Google Scholar] [CrossRef]
  37. Arayici, M.E.; Basbinar, Y.; Ellidokuz, H. Vitamin D Intake, Serum 25-Hydroxyvitamin-D (25(OH)D) Levels, and Cancer Risk: A Comprehensive Meta-Meta-Analysis Including Meta-Analyses of Randomized Controlled Trials and Observational Epidemiological Studies. Nutrients 2023, 15, 2722. [Google Scholar] [CrossRef]
  38. Sha, S.; Chen, L.J.; Brenner, H.; Schottker, B. Associations of 25-hydroxyvitamin D status and vitamin D supplementation use with mortality due to 18 frequent cancer types in the UK Biobank cohort. Eur. J. Cancer 2023, 191, 113241. [Google Scholar] [CrossRef]
  39. Li, J.; Zhang, H.; Zhu, H.; Dai, Z. 25-hydroxyvitamin D concentration is positively associated with overall survival in advanced pancreatic cancer: A systematic review and meta-analysis. Nutr. Res. 2023, 117, 73–82. [Google Scholar] [CrossRef]
  40. Kim, Y.; Chang, Y.; Cho, Y.; Chang, J.; Kim, K.; Park, D.I.; Park, S.K.; Joh, H.K.; Kim, M.K.; Kim, C.; et al. Serum 25-Hydroxyvitamin D Levels and Risk of Colorectal Cancer: An Age-Stratified Analysis. Gastroenterology 2023, 165, 920–931. [Google Scholar] [CrossRef]
  41. Wang, Y.F.; Li, L.; Deng, X.Q.; Fang, Y.J.; Zhang, C.X. Association of DNA methylation of vitamin D metabolic pathway related genes with colorectal cancer risk. Clin. Epigenetics 2023, 15, 140. [Google Scholar] [CrossRef]
  42. Chen, X.; Song, S.; Shi, J.; Wang, Z.; Song, W.; Wang, J.; Wang, G.; Wang, X. Evaluating the effect of body mass index and 25-hydroxy-vitamin D level on basal cell carcinoma using Mendelian randomization. Sci. Rep. 2023, 13, 16552. [Google Scholar] [CrossRef]
  43. Garland, C.F.; Garland, F.C.; Gorham, E.D.; Lipkin, M.; Newmark, H.; Mohr, S.B.; Holick, M.F. The role of vitamin D in cancer prevention. Am. J. Public Health 2006, 96, 252–261. [Google Scholar] [CrossRef] [PubMed]
  44. Izurieta-Pacheco, A.C.; Sangros-Gimenez, A.; Martinez-Garcia, E.; Perez-Jaume, S.; Mora, J.; Gorostegui-Obanos, M. Vitamin D Status in Children with High-risk Neuroblastoma. J. Pediatr. Hematol. Oncol. 2023, 45, e953–e958. [Google Scholar] [CrossRef] [PubMed]
  45. Rahman, S.T.; Waterhouse, M.; Pham, H.; Duarte Romero, B.; Baxter, C.; McLeod, D.S.A.; English, D.R.; Ebeling, E.R.; Hartel, G.; Armstrong, B.K.; et al. Effects of Vitamin D Supplementation on Telomere Length: An Analysis of Data from the Randomised Controlled D-Health Trial. J. Nutr. Health Aging 2023, 27, 609–616. [Google Scholar] [CrossRef]
  46. Manocha, A.; Brockton, N.T.; Cook, L.; Kopciuk, K.A. Low Serum Vitamin D Associated with Increased Tumor Size and Higher Grade in Premenopausal Canadian Women with Breast Cancer. Clin. Breast Cancer 2023, 23, e368–e376. [Google Scholar] [CrossRef]
  47. Reber, E.; Gomes, F.; Vasiloglou, M.F.; Schuetz, P.; Stanga, Z. Nutritional risk screening and assessment. J. Clin. Med. 2019, 8, 1065. [Google Scholar] [CrossRef]
  48. Qadir, J.; Majid, S.; Khan, M.S.; Wani, M.D.; Naikoo, N.A. Vitamin D receptor gene variations and their haplotypic association: Possible impact on gastric cancer risk. J. Cancer Res. Ther. 2023, 19, 1115–1125. [Google Scholar] [CrossRef] [PubMed]
  49. Cheema, H.A.; Fatima, M.; Shahid, A.; Bouaddi, O.; Elgenidy, A.; Rehman, A.U.; Kacimi, S.E.O.; Hasan, M.M.; Lee, K.Y. Vitamin D supplementation for the prevention of total cancer incidence and mortality: An updated systematic review and meta-analysis. Heliyon 2022, 8, e11290. [Google Scholar] [CrossRef]
  50. Woloszynska-Read, A.; Johnson, C.S.; Trump, D.L.; Endocrinology, R.C. Vitamin D and cancer: Clinical aspects. Best Pract. Res. Clin. Endocrinol. Metab. 2011, 25, 605–615. [Google Scholar] [CrossRef]
  51. Li, M.; Chen, P.; Li, J.; Chu, R.; Xie, D.; Wang, H. The impacts of circulating 25-hydroxyvitamin D levels on cancer patient outcomes: A systematic review and meta-analysis. J. Clin. Endocrinol. Metab. 2014, 99, 2327–2336. [Google Scholar] [CrossRef]
  52. Tretli, S.; Schwartz, G.G.; Torjesen, P.A.; Robsahm, T.E. Serum levels of 25-hydroxyvitamin D and survival in Norwegian patients with cancer of breast, colon, lung, and lymphoma: A population-based study. Cancer Causes Control 2012, 23, 363–370. [Google Scholar] [CrossRef]
  53. Robsahm, T.E.; Tretli, S.; Torjesen, P.A.; Babigumira, R.; Schwartz, G.G. Serum 25-hydroxyvitamin D levels predict cancer survival: A prospective cohort with measurements prior to and at the time of cancer diagnosis. Clin. Epidemiol. 2019, 11, 695–705. [Google Scholar] [CrossRef]
  54. Hernandez-Alonso, P.; Boughanem, H.; Canudas, S.; Becerra-Tomas, N.; Fernandez de la Puente, M.; Babio, N.; Gonzalez, M.M.; Salvado, J.S. Circulating vitamin D levels and colorectal cancer risk: A meta-analysis and systematic review of case-control and prospective cohort studies. Crit. Rev. Food Sci. Nutr. 2023, 63, 1–17. [Google Scholar] [CrossRef]
  55. McCullough, M.L.; Zoltick, E.S.; Weinstein, S.J.; Fedirko, V.; Wang, M.; Cook, N.R.; Eliessan, A.H.; Jacquotte, A.Z.; Agnoli, C.; Albanes, D.; et al. Circulating Vitamin D and Colorectal Cancer Risk: An International Pooling Project of 17 Cohorts. J. Natl. Cancer Inst. 2019, 111, 158–169. [Google Scholar] [CrossRef]
  56. Henn, M.; Martin-Gorgojo, V.; Martin-Moreno, J.M. Vitamin D in cancer prevention: Gaps in current knowledge and room for hope. Nutrients 2022, 14, 4512. [Google Scholar] [CrossRef]
  57. Poole, E.M.; Shu, X.; Caan, B.J.; Flatt, S.W.; Holmes, M.D.; Lu, W.; Kwan, M.L.; Nechuta, S.J.; Picrce, J.P.; Chen, W.Y. Postdiagnosis supplement use and breast cancer prognosis in the After Breast Cancer Pooling Project. Breast Cancer Res. Treat. 2013, 139, 529–537. [Google Scholar] [CrossRef]
  58. Zeichner, S.B.; Koru-Sengul, T.; Shah, N.; Liu, Q.; Markward, N.J.; Montero, A.J.; Gluck, S.; Silva, O.; Ahn, E.R. Improved clinical outcomes associated with vitamin D supplementation during adjuvant chemotherapy in patients with HER2+ nonmetastatic breast cancer. Clin. Breast Cancer 2015, 15, e1–e11. [Google Scholar] [CrossRef]
  59. Yokosawa, E.B.; Arthur, A.E.; Rentschler, K.M.; Wolf, G.T.; Rozek, L.S.; Mondul, A.M.J.T.L. Vitamin D intake and survival and recurrence in head and neck cancer patients. Laryngoscope 2018, 128, E371–E376. [Google Scholar] [CrossRef]
  60. Mulpur, B.H.; Nabors, L.B.; Thompson, R.C.; Olson, J.J.; LaRocca, R.V.; Thompson, Z.; Egan, K.M. Complementary therapy and survival in glioblastoma. Neuro-Oncol. Pract. 2015, 2, 122–126. [Google Scholar] [CrossRef]
  61. Madden, J.M.; Murphy, L.; Zgaga, L.; Bennett, K. De novo vitamin D supplement use post-diagnosis is associated with breast cancer survival. Breast Cancer Res. Treat. 2018, 172, 179–190. [Google Scholar] [CrossRef] [PubMed]
  62. Kim, H.; Lipsyc-Sharf, M.; Zong, X.; Wang, X.; Hur, J.; Song, M.; Wang, M.; Smith-Warner, S.A.; Fuchs, C.; Ogino, S.; et al. Total vitamin D intake and risks of early-onset colorectal cancer and precursors. Gastroenterology 2021, 161, 1208–1217.e9. [Google Scholar] [CrossRef]
  63. Lappe, J.M.; Travers-Gustafson, D.; Davies, K.M.; Recker, R.R.; Heaney, R.P. Vitamin D and calcium supplementation reduces cancer risk: Results of a randomized trial. Am. J. Clin. Nutr. 2007, 85, 1586–1591. [Google Scholar] [CrossRef]
  64. Bolland, M.J.; Grey, A.; Gamble, G.D.; Reid, I.R. Calcium and vitamin D supplements and health outcomes: A reanalysis of the Women’s Health Initiative (WHI) limited-access data set. Am. J. Clin. Nutr. 2011, 94, 1144–1149. [Google Scholar] [CrossRef]
  65. Lappe, J.; Watson, P.; Travers-Gustafson, D.; Recker, R.; Garland, C.; Gorham, E.; Baggerly, K.; McDonnel, S.L. Effect of vitamin D and calcium supplementation on cancer incidence in older women: A randomized clinical trial. JAMA 2017, 317, 1234–1243. [Google Scholar] [CrossRef]
  66. Zgaga, L. Heterogeneity of the effect of vitamin D supplementation in randomized controlled trials on cancer prevention. JAMA Netw. Open 2020, 3, e2027176. [Google Scholar] [CrossRef]
  67. Thacher, T.D.; Clarke, B.L. (Eds.) Vitamin D insufficiency. In Mayo Clinic Proceedings; Elsevier: Amsterdam, The Netherlands, 2011. [Google Scholar]
  68. Manson, J.E.; Cook, N.R.; Lee, I.-M.; Christen, W.; Bassuk, S.S.; Mora, S.; Gibson, H.; Gordan, D.; Copeland, T.; Agistino, D.; et al. Vitamin D supplements and prevention of cancer and cardiovascular disease. N. Engl. J. Med. 2019, 380, 33–44. [Google Scholar] [CrossRef]
  69. Chlebowski, R.T.; Johnson, K.C.; Kooperberg, C.; Pettinger, M.; Wactawski-Wende, J.; Rohan, T.; Rossouw, J.; Lane, D.; O’Sullivan, M.J.; Yasmeen, S. Calcium plus vitamin D supplementation and the risk of breast cancer. JNCI J. Natl. Cancer Inst. 2008, 100, 1581–1591. [Google Scholar] [CrossRef]
  70. Keum, N.; Lee, D.; Greenwood, D.; Manson, J.; Giovannucci, E. Vitamin D supplementation and total cancer incidence and mortality: A meta-analysis of randomized controlled trials. Ann. Oncol. 2019, 30, 733–743. [Google Scholar] [CrossRef]
  71. Virtanen, J.K.; Nurmi, T.; Aro, A.; Bertone-Johnson, E.R.; Hyppönen, E.; Kröger, H.; Lamberg-Allardt, C.; Manson, J.E.; Mursu, J.; Mäntyselkä, P. Vitamin D supplementation and prevention of cardiovascular disease and cancer in the Finnish Vitamin D Trial: A randomized controlled trial. Am. J. Clin. Nutr. 2022, 115, 1300–1310. [Google Scholar] [CrossRef]
  72. Garland, C.F.; French, C.B.; Baggerly, L.L.; Heaney, R.P. Vitamin D supplement doses and serum 25-hydroxyvitamin D in the range associated with cancer prevention. Anticancer Res. 2011, 31, 607–611. [Google Scholar]
  73. Bouillon, R. Comparative analysis of nutritional guidelines for vitamin D. Nat. Rev. Endocrinol. 2017, 13, 466–479. [Google Scholar] [CrossRef]
  74. Mead, M.N. Benefits of sunlight: A bright spot for human health. Natl. Inst. Environ. Health Sci. 2008, 116, A160–A167. [Google Scholar] [CrossRef]
  75. Freedman, D.; Dosemeci, M.; McGlynn, K. Sunlight and mortality from breast, ovarian, colon, prostate, and non-melanoma skin cancer: A composite death certificate based case-control study. Occup. Environ. Med. 2002, 59, 257–262. [Google Scholar] [CrossRef]
  76. Jacobs, E.T.; Kohler, L.N.; Kunihiro, A.G.; Jurutka, P.W. Vitamin D and colorectal, breast, and prostate cancers: A review of the epidemiological evidence. J. Cancer 2016, 7, 232. [Google Scholar] [CrossRef]
  77. Colston, K.; Colston, M.J.; Feldman, D. 1,25-dihydroxyvitamin D3 and malignant melanoma: The presence of receptors and inhibition of cell growth in culture. Endocrinology 1981, 108, 1083–1086. [Google Scholar] [CrossRef]
  78. Ahn, J.; Peters, U.; Albanes, D.; Purdue, M.P.; Abnet, C.C.; Chatterjee, N.; Horst, R.L.; Hollies, B.W.; Huang, W.; Shikany, J.M.; et al. Serum vitamin D concentration and prostate cancer risk: A nested case–control study. J. Natl. Cancer Inst. 2008, 100, 796–804. [Google Scholar] [CrossRef]
  79. Banach-Petrosky, W.; Ouyang, X.; Gao, H.; Nader, K.; Ji, Y.; Suh, N.; Dipaola, R.S.; Shen, C.A. Vitamin D inhibits the formation of prostatic intraepithelial neoplasia in Nkx3. 1; Pten mutant mice. Clin. Cancer Res. 2006, 12, 5895–5901. [Google Scholar] [CrossRef]
  80. Chung, I.; Han, G.; Seshadri, M.; Gillard, B.M.; Yu, W.-D.; Foster, B.A.; Trump, D.L.; Johnson, C.S. Role of vitamin D receptor in the antiproliferative effects of calcitriol in tumor-derived endothelial cells and tumor angiogenesis in vivo. Cancer Res. 2009, 69, 967–975. [Google Scholar] [CrossRef]
  81. Colston, K.; James, S.; Ofori-Kuragu, E.; Binderup, L.; Grant, A. Vitamin D receptors and anti-proliferative effects of vitamin D derivatives in human pancreatic carcinoma cells in vivo and in vitro. Br. J. Cancer 1997, 76, 1017–1020. [Google Scholar] [CrossRef]
  82. Getzenberg, R.H.; Light, B.W.; Lapco, P.E.; Konety, B.R.; Nangia, A.K.; Acierno, J.S.; Dhir, R.; Shurin, Z.; Day, R.S.; Trump, D.L. Vitamin D inhibition of prostate adenocarcinoma growth and metastasis in the Dunning rat prostate model system. Urology 1997, 50, 999–1006. [Google Scholar] [CrossRef]
  83. Jiang, F.; Bao, J.; Li, P.; Nicosia, S.V.; Bai, W. Induction of ovarian cancer cell apoptosis by 1, 25-dihydroxyvitamin D3 through the down-regulation of telomerase. J. Biol. Chem. 2004, 279, 53213–53221. [Google Scholar] [CrossRef]
  84. Krishnan, A.V.; Peehl, D.M.; Feldman, D. Inhibition of prostate cancer growth by vitamin D: Regulation of target gene expression. J. Cell. Biochem. 2003, 88, 363–371. [Google Scholar] [CrossRef]
  85. Ma, Y.; Yu, W.D.; Trump, D.L.; Johnson, C.S. 1,25D3 enhances antitumor activity of gemcitabine and cisplatin in human bladder cancer models. Cancer 2010, 116, 3294–3303. [Google Scholar] [CrossRef]
  86. Nakagawa, K.; Kawaura, A.; Kato, S.; Takeda, E.; Okano, T. 1α, 25-Dihydroxyvitamin D 3 is a preventive factor in the metastasis of lung cancer. Carcinogenesis 2005, 26, 429–440. [Google Scholar] [CrossRef]
  87. Pálmer, H.G.; Sánchez-Carbayo, M.; Ordóñez-Morán, P.; Larriba MaJs Cordón-Cardó, C.; Muñoz, A. Genetic signatures of differentiation induced by 1α, 25-dihydroxyvitamin D3 in human colon cancer cells. Cancer Res. 2003, 63, 7799–7806. [Google Scholar]
  88. Zhang, X.; Jiang, F.; Li, P.; Li, C.; Ma, Q.; Nicosia, S.V.; Bai, W. Growth suppression of ovarian cancer xenografts in nude mice by vitamin D analogue EB1089. Clin. Cancer Res. 2005, 11, 323–328. [Google Scholar] [CrossRef]
  89. Krishnan, A.V.; Feldman, D. Mechanisms of the anti-cancer and anti-inflammatory actions of vitamin D. Annu. Rev. Pharmacol. Toxicol. 2011, 51, 311–336. [Google Scholar] [CrossRef]
  90. Reitsma, P.; Rothberg, P.; Astrin, S.; Trial, J.; Bar-Shavit, Z.; Hall, A.; Teitelbaum, S.L.; Khan, A.J. Regulation of myc gene expression in HL-60 leukaemia cells by a vitamin D metabolite. Nature 1983, 306, 492–494. [Google Scholar] [CrossRef]
  91. Fleet, J.C.; Desmet, M.; Johnson, R.; Li, Y. Vitamin D and cancer: A review of molecular mechanisms. Biochem. J. 2012, 441, 61–76. [Google Scholar] [CrossRef]
  92. Moreno, J.; Krishnan, A.V.; Swami, S.; Nonn, L.; Peehl, D.M.; Feldman, D. Regulation of prostaglandin metabolism by calcitriol attenuates growth stimulation in prostate cancer cells. Cancer Res. 2005, 65, 7917–7925. [Google Scholar] [CrossRef]
  93. Sherman, M.H.; Ruth, T.Y.; Engle, D.D.; Ding, N.; Atkins, A.R.; Tiriac, H.; Collisson, E.A.; Connor, F.; Dyke, T.V.; Kozlov, S.; et al. Vitamin D receptor-mediated stromal reprogramming suppresses pancreatitis and enhances pancreatic cancer therapy. Cell 2014, 159, 80–93. [Google Scholar] [CrossRef]
  94. Swami, S.; Krishnan, A.V.; Williams, J.; Aggarwal, A.; Albertelli, M.A.; Horst, R.L.; Feldman, B.J.; Feldman, D. Vitamin D mitigates the adverse effects of obesity on breast cancer in mice. Endocr.-Relat. Cancer 2016, 23, 251. [Google Scholar] [CrossRef]
  95. Garland, C.F.; Gorham, E.D.; Mohr, S.B.; Garland, F.C. Vitamin D for cancer prevention: Global perspective. Ann. Epidemiol. 2009, 19, 468–483. [Google Scholar] [CrossRef]
  96. James, S.Y.; Mackay, A.G.; Colston, K.W. Effects of 1, 25 dihydroxyvitamin D3 and its analogues on induction of apoptosis in breast cancer cells. J. Steroid Biochem. Mol. Biol. 1996, 58, 395–401. [Google Scholar] [CrossRef]
  97. Bao, B.-Y.; Yao, J.; Lee, Y.-F. 1α, 25-dihydroxyvitamin D 3 suppresses interleukin-8-mediated prostate cancer cell angiogenesis. Carcinogenesis 2006, 27, 1883–1893. [Google Scholar] [CrossRef]
  98. Li, H.; Stampfer, M.J.; Hollis, J.B.W.; Mucci, L.A.; Gaziano, J.M.; Hunter, D.; Giovannucci, E.L.; Ma, J. A prospective study of plasma vitamin D metabolites, vitamin D receptor polymorphisms, and prostate cancer. PLoS Med. 2007, 4, e103. [Google Scholar] [CrossRef]
  99. Hossein-nezhad, A.; Spira, A.; Holick, M.F. Influence of vitamin D status and vitamin D3 supplementation on genome wide expression of white blood cells: A randomized double-blind clinical trial. PLoS ONE 2013, 8, e58725. [Google Scholar] [CrossRef]
  100. Pereira, F.; Larriba, M.J.; Munoz, A. Vitamin D and colon cancer. Endocr. Relat. Cancer 2012, 19, R51–R71. [Google Scholar] [CrossRef]
  101. Martinez-Reza, I.; Diaz, L.; Barrera, D.; Segovia-Mendoza, M.; Pedraza-Sanchez, S.; Soca-Chafre, G.; Larrea, F.; Garcia-Becerra, R. Calcitriol Inhibits the Proliferation of Triple-Negative Breast Cancer Cells through a Mechanism Involving the Proinflammatory Cytokines IL-1beta and TNF-alpha. J. Immunol. Res. 2019, 2019, 6384278. [Google Scholar] [CrossRef]
  102. Krishnan, A.V.; Swami, S.; Feldman, D. Equivalent anticancer activities of dietary vitamin D and calcitriol in an animal model of breast cancer: Importance of mammary CYP27B1 for treatment and prevention. J. Steroid Biochem. Mol. Biol. 2013, 136, 289–295. [Google Scholar] [CrossRef]
  103. Feldman, D.; Krishnan, A.V.; Swami, S.; Giovannucci, E.; Feldman, B.J. The role of vitamin D in reducing cancer risk and progression. Nat. Rev. Cancer 2014, 14, 342–357. [Google Scholar] [CrossRef]
  104. Friedrich, M.; Diesing, D.; Cordes, T.; Fischer, D.; Becker, S.; Chen, T.C.; Flanagnan, J.N.; Tangpricha, V.; Gherson, I.; Holick, M.F.; et al. Analysis of 25-hydroxyvitamin D3-1alpha-hydroxylase in normal and malignant breast tissue. Anticancer Res. 2006, 26, 2615–2620. [Google Scholar]
  105. Hewison, M.; Burke, F.; Evans, K.N.; Lammas, D.A.; Sansom, D.M.; Liu, P.; Modlin, R.L.; Adams, J.S. Extra-renal 25-hydroxyvitamin D3-1alpha-hydroxylase in human health and disease. J. Steroid Biochem. Mol. Biol. 2007, 103, 316–321. [Google Scholar] [CrossRef]
  106. Tangpricha, V.; Flanagan, J.N.; Whitlatch, L.W.; Tseng, C.C.; Chen, T.C.; Holt, P.R.; Lipkin, M.S.; Holick, M.F. 25-hydroxyvitamin D-1alpha-hydroxylase in normal and malignant colon tissue. Lancet 2001, 357, 1673–1674. [Google Scholar] [CrossRef]
  107. Hewison, M.; Kantorovich, V.; Liker, H.R.; Van Herle, A.J.; Cohan, P.; Zehnder, D.; Adams, J.S. Vitamin D-mediated hypercalcemia in lymphoma: Evidence for hormone production by tumor-adjacent macrophages. J. Bone Min. Res. 2003, 18, 579–582. [Google Scholar] [CrossRef]
  108. Maestro, M.A.; Molnar, F.; Carlberg, C. Vitamin D and its synthetic analogues. J. Med. Chem. 2019, 62, 6854–6875. [Google Scholar] [CrossRef]
  109. Leyssens, C.; Verlinden, L. Verstuyf AJFip. Future Vitam. D Analog. 2014, 5, 122. [Google Scholar]
  110. Duffy, M.J.; Murray, A.; Synnott, N.C.; O’Donovan, N.; Crown, J.J.C.R.i.O.H. Vitamin D analogues: Potential use in cancer treatment. Crit. Rev. Oncol. Hematol. 2017, 112, 190–197. [Google Scholar] [CrossRef]
  111. Hansen, C.M.; Hamberg, K.; Binderup, E.; Binderup, L. Seocalcitol (EB 1089) A vitamin D analogue of anti-cancer potential. Background, design, synthesis, pre-clinical and clinical evaluation. Curr. Pharm. Des. 2000, 6, 803–828. [Google Scholar] [CrossRef]
  112. James, S.Y.; Mercer, E.; Brady, M.; Binderup, L.; Colston, K.W. EB1089, a synthetic analogue of vitamin D, induces apoptosis in breast cancer cells in vivo and in vitro. Br. J. Pharmacol. 1998, 125, 953–962. [Google Scholar] [CrossRef]
  113. Flanagan, L.; Packman, K.; Juba, B.; O’Neill, S.; Tenniswood, M.; Welsh, J. Efficacy of Vitamin D compounds to modulate estrogen receptor negative breast cancer growth and invasion. J. Steroid Biochem. Mol. Biol. 2003, 84, 181–192. [Google Scholar] [CrossRef]
  114. Prudencio, J.; Akutsu, N.; Benlimame, N.; Wang, T.; Bastien, Y.; Lin, R.; Black, M.J.; Jamali, M.A.A.; White, J.H. Action of low calcemic 1α, 25-dihydroxyvitamin D3 analogue EB1089 in head and neck squamous cell carcinoma. J. Natl. Cancer Inst. 2001, 93, 745–753. [Google Scholar] [CrossRef]
  115. Kasiappan, R.; Sun, Y.; Lungchukiet, P.; Quarni, W.; Zhang, X.; Bai, W. Vitamin D suppresses leptin stimulation of cancer growth through microRNA. Cancer Res. 2014, 74, 6194–6204. [Google Scholar] [CrossRef]
  116. Li, Z.; Jia, Z.; Gao, Y.; Xie, D.; Wei, D.; Cui, J.; Mishra, L.; Haung, S.; Zhang, Y.; Xie, K. Activation of Vitamin D Receptor Signaling Downregulates the Expression of Nuclear FOXM1 Protein and Suppresses Pancreatic Cancer Cell StemnessVDR–FOXM1 Signaling in Pancreatic Cancer. Clin. Cancer Res. 2015, 21, 844–853. [Google Scholar] [CrossRef]
  117. Sundaram, S.; Sea, A.; Feldman, S.; Strawbridge, R.; Hoopes, P.J.; Demidenko, E.; Binderup, L.; Gewirtz, D.A. The combination of a potent vitamin D3 analog, EB 1089, with ionizing radiation reduces tumor growth and induces apoptosis of MCF-7 breast tumor xenografts in nude mice. Clin. Cancer Res. 2003, 9, 2350–2356. [Google Scholar]
  118. Ghous, Z.; Akhter, J.; Pourgholami, M.H.; Morris, D.L. Inhibition of hepatocellular cancer by EB1089: In vitro and in vivo study. Anticancer Res. 2008, 28, 3757–3761. [Google Scholar]
  119. Yoon, J.S.; Kim, J.Y.; Park, H.K.; Kim, E.S.; Ahn, K.S.; Yoon, S.S.; Cho, C.G.; Kim, B.K.; Lee, Y.Y. Antileukemic effect of a synthetic vitamin D3 analog, HY-11, with low potential to cause hypercalcemia. Int. J. Oncol. 2008, 32, 387–396. [Google Scholar] [CrossRef]
  120. Milczarek, M.; Rossowska, J.; Klopotowska, D.; Stachowicz, M.; Kutner, A.; Wietrzyk, J. Tacalcitol increases the sensitivity of colorectal cancer cells to 5-fluorouracil by downregulating the thymidylate synthase. J. Steroid Biochem. Mol. Biol. 2019, 190, 139–151. [Google Scholar] [CrossRef]
  121. Verlinden, L.; Verstuyf, A.; Van Camp, M.; Marcelis, S.; Sabbe, K.; Zhao, X.-Y.; Clercq, P.D.; Vandewalle, M.; Bouillon, R. Two novel 14-Epi-analogues of 1, 25-dihydroxyvitamin D3 inhibit the growth of human breast cancer cells in vitro and in vivo. Cancer Res. 2000, 60, 2673–2679. [Google Scholar]
  122. Okamoto, R.; Delansorne, R.; Wakimoto, N.; Doan, N.B.; Akagi, T.; Shen, M.; Ho, Q.H.; Said, J.W.; Koeffler, H.P. Inecalcitol, an analog of 1α, 25 (OH) 2D3, induces growth arrest of androgen-dependent prostate cancer cells. Int. J. Cancer 2012, 130, 2464–2473. [Google Scholar] [CrossRef]
  123. Ma, Y.; Yu, W.-D.; Hidalgo, A.A.; Luo, W.; Delansorne, R.; Johnson, C.S.; Trump, D.L. Inecalcitol, an analog of 1, 25D3, displays enhanced antitumor activity through the induction of apoptosis in a squamous cell carcinoma model system. Cell Cycle 2013, 12, 743–752. [Google Scholar] [CrossRef]
  124. Eelen, G.; Verlinden, L.; Rochel, N.; Claessens, F.; De Clercq, P.; Vandewalle, M.; Valentini, G.T.; Moras, D.; Bouillon, R.; Verstuyf, A. Superagonistic action of 14-epi-analogues of 1, 25-dihydroxyvitamin D explained by vitamin D receptor-coactivator interaction. Mol. Pharmacol. 2005, 67, 1566–1573. [Google Scholar] [CrossRef]
  125. González-Pardo, V.; Verstuyf, A.; Boland, R.; de Boland, A.R. Vitamin D analogue TX 527 down-regulates the NF-κB pathway and controls the proliferation of endothelial cells transformed by K aposi sarcoma herpesvirus. Br. J. Pharmacol. 2013, 169, 1635–1645. [Google Scholar] [CrossRef]
  126. Molnár, I.; Kute, T.; Willingham, M.C.; Schwartz, G.G.J.; Biology, M. 19-Nor-1α, 25-dihydroxyvitamin D2 (paricalcitol) exerts anticancer activity against HL-60 cells in vitro at clinically achievable concentrations. J. Steroid Biochem. Mol. Biol. 2004, 89, 539–543. [Google Scholar] [CrossRef]
  127. Bae, W.K.; Lee, J.H.; Park, M.S.; Ahn, J.S.; Hwang, J.E.; Shim, H.J.; Cho, S.H.; Chung, I.J. 19-nor-1α-25-Dihydroxyvitamin D2 (Paricalcitol) Induces Apoptosis in Gastric Cancer Cells. J. Cancer Prev. 2009, 14, 329–334. [Google Scholar]
  128. El-Shemi, A.G.; Refaat, B.; Kensara, O.A.; Mohamed, A.M.; Idris, S.; Ahmad, J.J.C.P.R. Paricalcitol enhances the chemopreventive efficacy of 5-fluorouracil on an intermediate-term model of azoxymethane-induced colorectal tumors in rats. Cancer Prev. Res. 2016, 9, 491–501. [Google Scholar] [CrossRef]
  129. Liu, Y.; Shin, D.-Y.; Oh, S.; Kim, S.; Koh, Y.; Kim, I. KML001 and doxercalciferol induce synergistic antileukemic effect in acute lymphoid leukemia cells. Oncol. Rep. 2017, 38, 481–487. [Google Scholar] [CrossRef]
  130. Kawa, S.; Yoshizawa, K.; Nikaido, T.; Kiyosawa, K.J.T.J.o.S.B.; Biology, M. Inhibitory effect of 22-oxa-1, 25-dihydroxyvitamin D3, maxacalcitol, on the proliferation of pancreatic cancer cell lines. J. Steroid Biochem. Mol. Biol. 2005, 97, 173–177. [Google Scholar] [CrossRef]
  131. Arensman, M.D.; Nguyen, P.; Kershaw, K.M.; Lay, A.R.; Ostertag-Hill, C.A.; Sherman, M.H.; Downes, M.; Liddle, C.; Evans, R.M.; Dawson, D.W. Calcipotriol Targets LRP6 to Inhibit Wnt Signaling in Pancreatic CancerCalcipotriol Targets LRP6. Mol. Cancer Res. 2015, 13, 1509–1519. [Google Scholar] [CrossRef] [PubMed]
  132. Sintov, A.C.; Berkovich, L.; Ben-Shabat, S. Inhibition of cancer growth and induction of apoptosis by BGP-13 and BGP-15, new calcipotriene-derived vitamin D 3 analogues, in-vitro and in-vivo studies. Investig. New Drugs 2013, 31, 247–255. [Google Scholar] [CrossRef] [PubMed]
  133. Filip-Psurska, B.; Psurski, M.; Anisiewicz, A.; Libako, P.; Zbrojewicz, E.; Maciejewska, M.; Chodyński, M.; Kutner, A.; Wietrzyk, J. Vitamin d compounds pri-2191 and pri-2205 enhance anastrozole activity in human breast cancer models. Int. J. Mol. Sci. 2021, 22, 2781. [Google Scholar] [CrossRef] [PubMed]
  134. Milczarek, M.; Chodyński, M.; Filip-Psurska, B.; Martowicz, A.; Krupa, M.; Krajewski, K.; Kutner, A.; Wietrzyk, J. Synthesis and biological activity of diastereomeric and geometric analogues of calcipotriol, PRI-2202 and PRI-2205, against human HL-60 leukemia and MCF-7 breast cancer cells. Cancers 2013, 5, 1355–1378. [Google Scholar] [CrossRef]
  135. Piatek, K.; Kutner, A.; Cacsire Castillo-Tong, D.; Manhardt, T.; Kupper, N.; Nowak, U.; Chodynski, M.; Marcinkowska, E.; Kallay, E.; Schepelmann, M. Vitamin D analogues regulate the vitamin d system and cell viability in ovarian cancer cells. Int. J. Mol. Sci. 2021, 23, 172. [Google Scholar] [CrossRef] [PubMed]
  136. Okamoto, R.; Gery, S.; Kuwayama, Y.; Borregaard, N.; Ho, Q.; Alvarez, R.; Akagi, T.; Liu, G.Y.; Uskokovic, M.R.; Koeffler, H.P. Novel Gemini vitamin D3 analogues: Large structure/function analysis and ability to induce antimicrobial peptide. Int. J. Cancer 2014, 134, 207–217. [Google Scholar] [CrossRef] [PubMed]
  137. So, J.Y.; Lee, H.J.; Smolarek, A.K.; Paul, S.; Wang, C.-X.; Maehr, H.; Uskovic, M.; Zheng, X.; Conney, A.H.; Cai, L.; et al. A novel Gemini vitamin D analog represses the expression of a stem cell marker CD44 in breast cancer. Mol. Pharmacol. 2011, 79, 360–367. [Google Scholar] [CrossRef]
  138. So, J.Y.; Wahler, J.E.; Yoon, T.; Smolarek, A.K.; Lin, Y.; Shih, W.J.; Maehr, H.; Uskokovic, M.; Liby, K.T.; Sporn, M.B.; et al. Oral Administration of a Gemini Vitamin D Analog, a Synthetic Triterpenoid and the Combination Prevents Mammary Tumorigenesis Driven by ErbB2 OverexpressionInhibition of Mammary Tumorigenesis by Vitamin D Analog and Synthetic Triterpenoid. Cancer Prev. Res. 2013, 6, 959–970. [Google Scholar] [CrossRef]
  139. So, J.Y.; Smolarek, A.K.; Salerno, D.M.; Maehr, H.; Uskokovic, M.; Liu, F.; Suh, N. Targeting CD44-STAT3 signaling by Gemini vitamin D analog leads to inhibition of invasion in basal-like breast cancer. PLoS ONE 2013, 8, e54020. [Google Scholar] [CrossRef]
  140. Wahler, J.; So, J.Y.; Kim, Y.C.; Liu, F.; Maehr, H.; Uskokovic, M.; Suh, N. Inhibition of the Transition of Ductal Carcinoma In Situ to Invasive Ductal Carcinoma by a Gemini Vitamin D AnalogInhibition of Breast Cancer Progression with a Vitamin D Analog. Cancer Prev. Res. 2014, 7, 617–626. [Google Scholar] [CrossRef]
  141. Huet, T.; Maehr, H.; Lee, H.J.; Uskokovic, M.R.; Suh, N.; Moras, D.; Rochel, N. Structure–function study of gemini derivatives with two different side chains at C-20, Gemini-0072 and Gemini-0097. MedChemComm 2011, 2, 424–429. [Google Scholar] [CrossRef]
  142. Lee, H.J.; Paul, S.; Atalla, N.; Thomas, P.E.; Lin, X.; Yang, I.; Buckley, B.; Lu, G.; Zheng, X.; Lou, Y.R.; et al. Gemini vitamin D analogues inhibit estrogen receptor–positive and estrogen receptor–negative mammary tumorigenesis without hypercalcemic toxicity. Cancer Prev. Res. 2008, 1, 476–484. [Google Scholar] [CrossRef]
  143. Chiang, K.-C.; Yeh, T.-S.; Chen, S.-C.; Pang, J.-H.S.; Yeh, C.-N.; Hsu, J.-T.; Chen, L.W.; Kuo, S.F.; Takano, M.; Kittaka, A.; et al. The vitamin D analog, MART-10, attenuates triple negative breast cancer cells metastatic potential. Int. J. Mol. Sci. 2016, 17, 606. [Google Scholar] [CrossRef] [PubMed]
  144. Lambert, J.R.; Eddy, V.J.; Young, C.D.; Persons, K.S.; Sarkar, S.; Kelly, J.A.; Genova, E.; Lucia, M.S.; Faller, D.V.; Ray, R. A vitamin D receptor-alkylating derivative of 1α, 25-dihydroxyvitamin D3 inhibits growth of human kidney cancer cells and suppresses tumor growth. Cancer Prev. Res. 2010, 3, 1596–1607. [Google Scholar] [CrossRef] [PubMed]
  145. Fichera, A.; Little, N.; Dougherty, U.; Mustafi, R.; Cerda, S.; Li, Y.C.; Delgado, J.; Arora, A.; Campbell, L.K.; Joseph, L.; et al. A vitamin D analogue inhibits colonic carcinogenesis in the AOM/DSS model. J. Surg. Res. 2007, 142, 239–245. [Google Scholar] [CrossRef] [PubMed]
  146. Uskokovic, M.R.; Norman, A.W.; Manchand, P.S.; Studzinski, G.P.; Campbell, M.J.; Koeffler, H.P.; Takeuchi, A.; Siu-Caldera, M.L.; Rao, D.S.; Reddy, G.S. Highly active analogues of 1α, 25-dihydroxyvitamin D3 that resist metabolism through C-24 oxidation and C-3 epimerization pathways. Steroids 2001, 66, 463–471. [Google Scholar] [CrossRef]
  147. Salomón, D.G.; Grioli, S.M.; Buschiazzo, M.; Mascaró, E.; Vitale, C.; Radivoy, G.; Perez, M.; Fall, Y.; Mesri, E.A.; Curino, A.C.; et al. Novel alkynylphosphonate analogue of calcitriol with potent antiproliferative effects in cancer cells and lack of calcemic activity. ACS Med. Chem. Lett. 2011, 2, 503–508. [Google Scholar] [CrossRef]
  148. Ferronato, M.J.; Obiol, D.J.; Fermento, M.E.; Gandini, N.A.; Alonso, E.N.; Salomón, D.G.; Vitale, C.; Mascaro, E.; Fall, Y.; Raimondi, R.; et al. The alkynylphosphonate analogue of calcitriol EM1 has potent anti-metastatic effects in breast cancer. J. Steroid Biochem. Mol. Biol. 2015, 154, 285–293. [Google Scholar] [CrossRef]
  149. Elmore, S. Apoptosis: A review of programmed cell death. Toxicol. Pathol. 2007, 35, 495–516. [Google Scholar] [CrossRef]
  150. Tosetti, F.; Ferrari, N.; De Flora, S.; Albini, A. ‘Angioprevention’: Angiogenesis is a common and key target for cancer chemopreventive agents. FASEB J. 2002, 16, 2–14. [Google Scholar] [CrossRef] [PubMed]
  151. Guzey, M.; Kitada, S.; Reed, J.C. Apoptosis induction by 1α, 25-dihydroxyvitamin D3 in prostate cancer. Mol. Cancer Ther. 2002, 1, 667–677. [Google Scholar]
  152. Negri, M.; Gentile, A.; de Angelis, C.; Montò, T.; Patalano, R.; Colao, A.; Pivonello, R.; Pivonello, C. Vitamin D-induced molecular mechanisms to potentiate cancer therapy and to reverse drug-resistance in cancer cells. Nutrients 2020, 12, 1798. [Google Scholar] [CrossRef]
  153. Narvaez, C.J.; Welsh, J. Role of mitochondria and caspases in vitamin D-mediated apoptosis of MCF-7 breast cancer cells. J. Biol. Chem. 2001, 276, 9101–9107. [Google Scholar] [CrossRef] [PubMed]
  154. Veeresh, P.K.M.; Basavaraju, C.G.; Dallavalasa, S.; Anantharaju, P.G.; Natraj, S.M.; Sukocheva, O.A.; Madhunapantula, S.V. Vitamin D3 Inhibits the Viability of Breast Cancer Cells In Vitro and Ehrlich Ascites Carcinomas in Mice by Promoting Apoptosis and Cell Cycle Arrest and by Impeding Tumor Angiogenesis. Cancers 2023, 15, 4833. [Google Scholar] [CrossRef]
  155. Mathiasen, I.S.; Sergeev, I.N.; Bastholm, L.; Elling, F.; Norman, A.W.; Jäättelä, M. Calcium and calpain as key mediators of apoptosis-like death induced by vitamin D compounds in breast cancer cells. J. Biol. Chem. 2002, 277, 30738–30745. [Google Scholar] [CrossRef]
  156. Høyer-Hansen, M.; Bastholm, L.; Mathiasen, I.; Elling, F.; Jäättelä, M. Vitamin D analog EB1089 triggers dramatic lysosomal changes and Beclin 1-mediated autophagic cell death. Cell Death Differ. 2005, 12, 1297–1309. [Google Scholar] [CrossRef]
  157. Darío Díaz, G.; Paraskeva, C.; Thomas, M.; Binderup, L.; Hague, A. Apoptosis is induced by the active metabolite of vitamin D3 and its analogue EB1089 in colorectal adenoma and carcinoma cells: Possible implications for prevention and therapy. Cancer Res. 2000, 60, 2304–2312. [Google Scholar]
  158. McGuire, T.F.; Trump, D.L.; Johnson, C.S. Vitamin D3-induced apoptosis of murine squamous cell carcinoma cells: Selective induction of caspase-dependent MEK cleavage and up-regulation of MEKK-1. J. Biol. Chem. 2001, 276, 26365–26373. [Google Scholar] [CrossRef] [PubMed]
  159. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [PubMed]
  160. Moffatt, K.A.; Johannes, W.U.; Hedlund, T.E.; Miller, G.J. Growth inhibitory effects of 1α, 25-dihydroxyvitamin D3 are mediated by increased levels of p21 in the prostatic carcinoma cell line ALVA-31. Cancer Res. 2001, 61, 7122–7129. [Google Scholar]
  161. Liu, M.; Lee, M.-H.; Cohen, M.; Bommakanti, M.; Freedman, L.P. Transcriptional activation of the Cdk inhibitor p21 by vitamin D3 leads to the induced differentiation of the myelomonocytic cell line U937. Genes Dev. 1996, 10, 142–153. [Google Scholar] [CrossRef]
  162. Saramäki, A.; Banwell, C.M.; Campbell, M.J.; Carlberg, C. Regulation of the human p21 (waf1/cip1) gene promoter via multiple binding sites for p53 and the vitamin D 3 receptor. Nucleic Acids Res. 2006, 34, 543–554. [Google Scholar] [CrossRef]
  163. Kawa, S.; Nikaido, T.; Aoki, Y.; Zhai, Y.; Kumagai, T.; Furihata, K.; Fujii, S.; Kiyosawa, K. Vitamin D analogues up-regulate p21 and p27 during growth inhibition of pancreatic cancer cell lines. Br. J. Cancer 1997, 76, 884–889. [Google Scholar] [CrossRef] [PubMed]
  164. Hershberger, P.A.; Modzelewski, R.A.; Shurin, Z.R.; Rueger, R.M.; Trump, D.L.; Johnson, C.S. 1,25-Dihydroxycholecalciferol (1, 25-D3) inhibits the growth of squamous cell carcinoma and down-modulates p21Waf1/Cip1 in vitro and in vivo. Cancer Res. 1999, 59, 2644–2649. [Google Scholar]
  165. Jiang, F.; Li, P.; Fornace, A.J.; Nicosia, S.V.; Bai, W. G2/M arrest by 1,25-dihydroxyvitamin D3 in ovarian cancer cells mediated through the induction of GADD45 via an exonic enhancer. J. Biol. Chem. 2003, 278, 48030–48040. [Google Scholar] [CrossRef]
  166. Bernardi, R.J.; Trump, D.L.; Yu, W.D.; McGuire, T.F.; Hershberger, P.A.; Johnson, C.S. Combination of 1alpha,25-dihydroxyvitamin D(3) with dexamethasone enhances cell cycle arrest and apoptosis: Role of nuclear receptor cross-talk and Erk/Akt signaling. Clin. Cancer Res. 2001, 7, 4164–4173. [Google Scholar] [PubMed]
  167. Liu, W.; Asa, S.L.; Fantus, I.G.; Walfish, P.G.; Ezzat, S. Vitamin D arrests thyroid carcinoma cell growth and induces p27 dephosphorylation and accumulation through PTEN/akt-dependent and-independent pathways. Am. J. Pathol. 2002, 160, 511–519. [Google Scholar] [CrossRef] [PubMed]
  168. Kemmochi, S.; Fujimoto, H.; Woo, G.-H.; Hirose, M.; Nishikawa, A.; Mitsumori, K.; Shibutani, M. Preventive effects of calcitriol on the development of capsular invasive carcinomas in a rat two-stage thyroid carcinogenesis model. J. Vet. Med. Sci. 2011, 73, 655–664. [Google Scholar] [CrossRef]
  169. Gonzalez-Pardo, V.; D’Elia, N.; Verstuyf, A.; Boland, R.; de Boland, A.R. NFκB pathway is down-regulated by 1α, 25 (OH) 2-vitamin D3 in endothelial cells transformed by Kaposi sarcoma-associated herpes virus G protein coupled receptor. Steroids 2012, 77, 1025–1032. [Google Scholar] [CrossRef]
  170. Capiati, D.A.; Rossi, A.M.; Picotto, G.; Benassati, S.; Boland, R.L. Inhibition of serum-stimulated mitogen activated protein kinase by 1α, 25 (OH) 2-vitamin D3 in MCF-7 breast cancer cells. J. Cell. Biochem. 2004, 93, 384–397. [Google Scholar] [CrossRef]
  171. Colston, K.; Perks, C.; Xie, S.; Holly, J. Growth inhibition of both MCF-7 and Hs578T human breast cancer cell lines by vitamin D analogues is associated with increased expression of insulin-like growth factor binding protein-3. J. Mol. Endocrinol. 1998, 20, 157–162. [Google Scholar] [CrossRef]
  172. Koli, K.; Keski-Oja, J. 1,25-Dihydroxyvitamin D3 enhances the expression of transforming growth factor β1 and its latent form binding protein in cultured breast carcinoma cells. Cancer Res. 1995, 55, 1540–1546. [Google Scholar]
  173. Alvarez-Díaz, S.; Valle, N.; Ferrer-Mayorga, G.; Lombardía, L.; Herrera, M.; Domínguez, O.; Segura, M.F.; Bonilla, F.; Hernando, E.; Munoz, A. MicroRNA-22 is induced by vitamin D and contributes to its antiproliferative, antimigratory and gene regulatory effects in colon cancer cells. Hum. Mol. Genet. 2012, 21, 2157–2165. [Google Scholar] [CrossRef] [PubMed]
  174. Guan, H.; Liu, C.; Chen, Z.; Wang, L.; Li, C.; Zhao, J.; Yu, Y.; Zhang, P.; Chen, W.; Jiang, A. 1, 25-Dihydroxyvitamin D3 up-regulates expression of hsa-let-7a-2 through the interaction of VDR/VDRE in human lung cancer A549 cells. Gene 2013, 522, 142–146. [Google Scholar] [CrossRef] [PubMed]
  175. Martin, T.A.; Ye, L.; Sanders, A.J.; Lane, J.; Jiang, W.G. Cancer invasion and metastasis: Molecular and cellular perspective. In Madame Curie Bioscience Database [Internet]; Landes Bioscience: Austin, TX, USA, 2013. [Google Scholar]
  176. Sporn, M.B. The war on cancer. Lancet 1996, 347, 1377–1381. [Google Scholar] [CrossRef] [PubMed]
  177. Liu, W.; Asa, S.L.; Ezzat, S. 1α, 25-Dihydroxyvitamin D3 targets PTEN-dependent fibronectin expression to restore thyroid cancer cell adhesiveness. Mol. Endocrinol. 2005, 19, 2349–2357. [Google Scholar] [CrossRef]
  178. Ma, Y.; Yu, W.D.; Su, B.; Seshadri, M.; Luo, W.; Trump, D.L.; Johnson, C.S. Regulation of motility, invasion, and metastatic potential of squamous cell carcinoma by 1α, 25-dihydroxycholecalciferol. Cancer 2013, 119, 563–574. [Google Scholar] [CrossRef]
  179. Bao, B.-Y.; Yeh, S.-D.; Lee, Y.-F. 1α, 25-dihydroxyvitamin D 3 inhibits prostate cancer cell invasion via modulation of selective proteases. Carcinogenesis 2005, 27, 32–42. [Google Scholar] [CrossRef]
  180. Hsu, J.-W.; Yasmin-Karim, S.; King, M.R.; Wojciechowski, J.C.; Mickelsen, D.; Blair, M.L.; Ting, H.J.; Ma, W.L.; Lee, Y.F. Suppression of prostate cancer cell rolling and adhesion to endothelium by 1α, 25-dihydroxyvitamin D3. Am. J. Pathol. 2011, 178, 872–880. [Google Scholar] [CrossRef]
  181. Pálmer, H.G.; González-Sancho, J.M.; Espada, J.; Berciano, M.T.; Puig, I.; Baulida, J.; Quintanilla, M.; Cano, A.; Herreros, A.G.; Lafarga, M.; et al. Vitamin D3 promotes the differentiation of colon carcinoma cells by the induction of E-cadherin and the inhibition of β-catenin signaling. J. Cell Biol. 2001, 154, 369–388. [Google Scholar] [CrossRef] [PubMed]
  182. Young, M.; Halpin, J.; Hussain, R.; Lozano, Y.; Djordjevic, A.; Devata, S.; Matthews, J.P.; Wright, M.A. Inhibition of tumor production of granulocyte-macrophage colony-stimulating factor by 1 alpha, 25-dihydroxyvitamin D3 reduces tumor motility and metastasis. Invasion Metastasis 1993, 13, 169–177. [Google Scholar]
  183. Young, M.R.I.; Lozano, Y. Inhibition of tumor invasiveness by 1-alpha-25-dihydroxy-vitamin D coupled to a decline in protein kinase A activity and an increase in cytoskeletal organization. Clin. Exp. Metastasis 1997, 15, 102–110. [Google Scholar] [CrossRef]
  184. Pendás-Franco, N.; García, J.M.; Peña, C.; Valle, N.; Pálmer, H.G.; Heinäniemi, M.; Carlberg, C.; Jimenez, B.; Bonilla, F.; Munoz, A.; et al. DICKKOPF-4 is induced by TCF/β-catenin and upregulated in human colon cancer, promotes tumour cell invasion and angiogenesis and is repressed by 1α, 25-dihydroxyvitamin D3. Oncogene 2008, 27, 4467–4477. [Google Scholar] [CrossRef] [PubMed]
  185. Chiang, K.-C.; Yeh, C.-N.; Hsu, J.-T.; Jan, Y.-Y.; Chen, L.-W.; Kuo, S.-F.; Takano, M.; Kittaka, A.; Chen, T.C.; Chen, W.T.; et al. The vitamin D analog, MART-10, represses metastasis potential via downregulation of epithelial–mesenchymal transition in pancreatic cancer cells. Cancer Lett. 2014, 354, 235–244. [Google Scholar] [CrossRef] [PubMed]
  186. Rebsamen, M.C.; Sun, J.; Norman, A.W.; Liao, J.K. 1α, 25-Dihydroxyvitamin D3 induces vascular smooth muscle cell migration via activation of phosphatidylinositol 3-kinase. Circ. Res. 2002, 91, 17–24. [Google Scholar] [CrossRef] [PubMed]
  187. Yudoh, K.; Matsuno, H.; Kimura, T. 1α, 25-Dihydroxyvitamin D3 inhibits in vitro invasiveness through the extracellular matrix and in vivo pulmonary metastasis of B16 mouse melanoma. J. Lab. Clin. Med. 1999, 133, 120–128. [Google Scholar] [CrossRef]
  188. Nakagawa, K.; Sasaki, Y.; Kato, S.; Kubodera, N.; Okano, T. 22-Oxa-1α, 25-dihydroxyvitamin D 3 inhibits metastasis and angiogenesis in lung cancer. Carcinogenesis 2005, 26, 1044–1054. [Google Scholar] [CrossRef]
  189. Bhatia, V.; Saini, M.K.; Shen, X.; Bi, L.X.; Qiu, S.; Weigel, N.L.; Falzon, M. EB1089 inhibits the parathyroid hormone–related protein–enhanced bone metastasis and xenograft growth of human prostate cancer cells. Mol. Cancer Ther. 2009, 8, 1787–1798. [Google Scholar] [CrossRef]
  190. Blackadar, C.B. Historical review of the causes of cancer. World J. Clin. Oncol. 2016, 7, 54. [Google Scholar] [CrossRef] [PubMed]
  191. Brackman, D.; Lund-Johansen, F.; Aarskog, D. Expression of cell surface antigens during the differentiation of HL-60 cells induced by 1, 25-ihydroxyvitamin D3, retinoic acid and DMSO. Leuk. Res. 1995, 19, 57–64. [Google Scholar] [CrossRef]
  192. Zakaria Hmama, D.N.; Sly, L.; Knutson, K.L.; Herrera-velit, P.; Reiner, N.E. 1,25-Dihydroxyvitamin D 3–induced Myeloid Cell Differentiation Is Regulated by a Vitamin D Receptor–Phosphatidylinositol 3-Kinase Signaling Complex. J. Exp. Med. 1999, 190, 1583–1594. [Google Scholar] [CrossRef]
  193. Wang, X.; Studzinski, G.P. Activation of extracellular signal-regulated kinases (ERKs) defines the first phase of 1, 25-dihydroxyvitamin D3-induced differentiation of HL60 cells. J. Cell. Biochem. 2001, 80, 471–482. [Google Scholar] [CrossRef]
  194. Wang, Q.; Wang, X.; Studzinski, G.P. Jun N-terminal kinase pathway enhances signaling of monocytic differentiation of human leukemia cells induced by 1, 25-dihydroxyvitamin D3. J. Cell. Biochem. 2003, 89, 1087–1101. [Google Scholar] [CrossRef] [PubMed]
  195. Ji, Y.; Studzinski, G.P. Retinoblastoma protein and CCAAT/enhancer-binding protein β are required for 1, 25-dihydroxyvitamin D3-induced monocytic differentiation of HL60 cells. Cancer Res. 2004, 64, 370–377. [Google Scholar] [CrossRef] [PubMed]
  196. Marcinkowska, E.; Garay, E.; Gocek, E.; Chrobak, A.; Wang, X.; Studzinski, G.P. Regulation of C/EBPβ isoforms by MAPK pathways in HL60 cells induced to differentiate by 1, 25-dihydroxyvitamin D3. Exp. Cell Res. 2006, 312, 2054–2065. [Google Scholar] [CrossRef]
  197. Chen, A.; Davis, B.H.; Bissonnette, M.; Scaglione-Sewell, B.; Brasitus, T.A. 1,25-Dihydroxyvitamin D3 stimulates activator protein-1-dependent Caco-2 cell differentiation. J. Biol. Chem. 1999, 274, 35505–35513. [Google Scholar] [CrossRef]
  198. Larriba, M.J.; Garcia de Herreros, A.; Munoz, A. Vitamin D and the Epithelial to Mesenchymal Transition. Stem Cells Int. 2016, 2016, 6213872. [Google Scholar] [CrossRef] [PubMed]
  199. Pendas-Franco, N.; Gonzalez-Sancho, J.M.; Suarez, Y.; Aguilera, O.; Steinmeyer, A.; Gamallo, C.; Berciano, M.T.; Lafarga, M.; Monoz, A. Vitamin D regulates the phenotype of human breast cancer cells. Differentiation 2007, 75, 193–207. [Google Scholar] [CrossRef]
  200. Lazzaro, G.; Agadir, A.; Qing, W.; Poria, M.; Mehta, R.; Moriarty, R.; Gupta, T.K.D.; Zhang, X.K.; Mehta, R.G. Induction of differentiation by 1α-hydroxyvitamin D5 in T47D human breast cancer cells and its interaction with vitamin D receptors. Eur. J. Cancer 2000, 36, 780–786. [Google Scholar] [CrossRef]
  201. Wang, Q.; Lee, D.; Sysounthone, V.; Chandraratna, R.A.; Christakos, S.; Korah, R.; Wieder, R. 1, 25-dihydroxyvitamin D3 and retonic acid analogues induce differentiation in breast cancer cells with function-and cell-specific additive effects. Breast Cancer Res. Treat. 2001, 67, 157–168. [Google Scholar] [CrossRef]
  202. Haselberger, M.; Springwald, A.; Konwisorz, A.; Lattrich, C.; Goerse, R.; Ortmann, O.; Treeck, O. Silencing of the icb-1 gene inhibits the induction of differentiation-associated genes by vitamin D3 and all-trans retinoic acid in gynecological cancer cells. Int. J. Mol. Med. 2011, 28, 121–127. [Google Scholar]
  203. Beer, T.M.; Garzotto, M.; Park, B.; Mori, M.; Myrthue, A.; Janeba, N.; Sauer, D.; Eilers, K. Effect of calcitriol on prostate-specific antigen in vitro and in humans. Clin. Cancer Res. 2006, 12, 2812–2816. [Google Scholar] [CrossRef]
  204. Folkman, J. Tumor angiogenesis: Therapeutic implications. N. Engl. J. Med. 1971, 285, 1182–1186. [Google Scholar] [PubMed]
  205. Folkman, J. (Ed.) Role of angiogenesis in tumor growth and metastasis. In Seminars in Oncology; Elsevier: Amsterdam, The Netherlands, 2002. [Google Scholar]
  206. Bernardi, R.J.; Johnson, C.S.; Modzelewski, R.A.; Trump, D.L. Antiproliferative effects of 1α, 25-dihydroxyvitamin D3 and vitamin D analogues on tumor-derived endothelial cells. Endocrinology 2002, 143, 2508–2514. [Google Scholar] [CrossRef] [PubMed]
  207. Merke, J.; Milde, P.; Lewicka, S.; Hügel, U.; Klaus, G.; Mangelsdorf, D.; Haussler, M.R.; Rauterburg, E.W.; Ritz, E. Identification and regulation of 1, 25-dihydroxyvitamin D3 receptor activity and biosynthesis of 1, 25-dihydroxyvitamin D3. Studies in cultured bovine aortic endothelial cells and human dermal capillaries. J. Clin. Investig. 1989, 83, 1903–1915. [Google Scholar] [CrossRef] [PubMed]
  208. Chung, I.; Wong, M.K.; Flynn, G.; Yu, W.-D.; Johnson, C.S.; Trump, D.L. Differential antiproliferative effects of calcitriol on tumor-derived and matrigel-derived endothelial cells. Cancer Res. 2006, 66, 8565–8573. [Google Scholar] [CrossRef]
  209. Gonzalez-Pardo, V.; Martin, D.; Gutkind, J.S.; Verstuyf, A.; Bouillon, R.; de Boland, A.R.; Boland, R.L. 1α, 25-dihydroxyvitamin D3 and its TX527 analog inhibit the growth of endothelial cells transformed by Kaposi sarcoma-associated herpes virus G protein-coupled receptor in vitro and in vivo. Endocrinology 2010, 151, 23–31. [Google Scholar] [CrossRef]
  210. Mantell, D.; Owens, P.; Bundred, N.; Mawer, E.; Canfield, A. 1α, 25-dihydroxyvitamin D3 inhibits angiogenesis in vitro and in vivo. Circ. Res. 2000, 87, 214–220. [Google Scholar] [CrossRef]
  211. Oikawa, T.; Hirotani, K.; Ogasawara, H.; Katayama, T.; Nakamura, O.; Iwaguchi, T.; Hiragun, A. Inhibition of angiogenesis by vitamin D3 analogues. Eur. J. Pharmacol. 1990, 178, 247–250. [Google Scholar] [CrossRef]
  212. Suzuki, T.; Sano, Y.; Kinoshita, S. Effects of 1α, 25-dihydroxyvitamin D3 on Langerhans cell migration and corneal neovascularization in mice. Investig. Ophthalmol. Vis. Sci. 2000, 41, 154–158. [Google Scholar]
  213. Albert, D.M.; Scheef, E.A.; Wang, S.; Mehraein, F.; Darjatmoko, S.R.; Sorenson, C.M.; Sheibani, N. Calcitriol is a potent inhibitor of retinal neovascularization. Investig. Ophthalmol. Vis. Sci. 2007, 48, 2327–2334. [Google Scholar] [CrossRef]
  214. Zehnder, D.; Bland, R.; Chana, R.S.; Wheeler, D.C.; Howie, A.J.; Williams, M.C.; Stewart, P.M.; Hewison, M. Synthesis of 1, 25-dihydroxyvitamin D3 by human endothelial cells is regulated by inflammatory cytokines: A novel autocrine determinant of vascular cell adhesion. J. Am. Soc. Nephrol. 2002, 13, 621–629. [Google Scholar] [CrossRef]
  215. Chen, Y.; Liu, W.; Sun, T.; Huang, Y.; Wang, Y.; Deb, D.K.; Yoon, D.; Kong, J.; Thadhani, R.; Li, Y.C. 1, 25-Dihydroxyvitamin D promotes negative feedback regulation of TLR signaling via targeting MicroRNA-155–SOCS1 in macrophages. J. Immunol. 2013, 190, 3687–3695. [Google Scholar] [CrossRef] [PubMed]
  216. Eisman, J.A.; Barkla, D.H.; Tutton, P.J. Suppression of in vivo growth of human cancer solid tumor xenografts by 1, 25-dihydroxyvitamin D3. Cancer Res. 1987, 47, 21–25. [Google Scholar] [PubMed]
  217. Evans, S.R.; Houghton, A.M.; Schumaker, L.; Brenner, R.V.; Buras, R.R.; Davoodi, F.; Nauta, R.J.; Shabahang, M. Vitamin D Receptor and Growth Inhibition by 1, 25-Dihydroxyvitamin D3in Human Malignant Melanoma Cell Lines. J. Surg. Res. 1996, 61, 127–133. [Google Scholar] [CrossRef]
  218. Fu, B.; Wang, H.; Wang, J.; Barouhas, I.; Liu, W.; Shuboy, A.; Bushinsky, D.A.; Zhou, D.; Favus, M.J. Epigenetic regulation of BMP2 by 1, 25-dihydroxyvitamin D3 through DNA methylation and histone modification. PLoS ONE 2013, 8, e61423. [Google Scholar] [CrossRef]
  219. Hansen, C.M.; Madsen, M.W.; Arensbak, B.; Skak-Nielsen, T.; Latini, S.; Binderup, L. Down-regulation of Laminin-Binding Integrins by 1α, 25= Dihydroxyvitamin D3 in Human Melanoma Cells in Vitro. Cell Adhes. Commun. 1998, 5, 109–120. [Google Scholar] [CrossRef]
  220. Newton-Bishop, J.A.; Beswick, S.; Randerson-Moor, J.; Chang, Y.-M.; Affleck, P.; Elliott, F.; Chan, M.; Leake, S.; Karpavicius, B.; Haynes, S.; et al. Serum 25-hydroxyvitamin D3 levels are associated with breslow thickness at presentation and survival from melanoma. J. Clin. Oncol. 2009, 27, 5439. [Google Scholar] [CrossRef] [PubMed]
  221. Osborne, J.; Hutchinson, P. Vitamin D and systemic cancer: Is this relevant to malignant melanoma? Br. J. Dermatol. 2002, 147, 197–213. [Google Scholar] [CrossRef]
  222. Elzehery, R.R.; Baiomy, A.A.; Hegazy, M.A.-F.; Fares, R.; El-Gilany, A.-H.; Hegazi, R. Vitamin D status, receptor gene BsmI (A/G) polymorphism and breast cancer in a group of Egyptian females. Egypt. J. Med. Hum. Genet. 2017, 18, 269–273. [Google Scholar] [CrossRef]
  223. Zhang, K.; Song, L. Association between vitamin D receptor gene polymorphisms and breast cancer risk: A meta-analysis of 39 studies. PLoS ONE 2014, 9, e96125. [Google Scholar] [CrossRef]
  224. Fuhrman, B.J.; Freedman, D.M.; Bhatti, P.; Doody, M.M.; Fu, Y.-P.; Chang, S.-C.; Linet, M.S.; Sigurdson, A.J. Sunlight, polymorphisms of vitamin D-related genes and risk of breast cancer. Anticancer Res. 2013, 33, 543–551. [Google Scholar]
  225. Rai, V.; Abdo, J.; Agrawal, S.; Agrawal, D.K. Vitamin D receptor polymorphism and cancer: An update. Anticancer Res. 2017, 37, 3991–4003. [Google Scholar]
  226. Tagliabue, E.; Raimondi, S.; Gandini, S. Vitamin D, cancer risk, and mortality. Adv. Food Nutr. Res. 2015, 75, 1–52. [Google Scholar] [PubMed]
  227. Raimondi, S.; Johansson, H.; Maisonneuve, P.; Gandini, S. Review and meta-analysis on vitamin D receptor polymorphisms and cancer risk. Carcinogenesis 2009, 30, 1170–1180. [Google Scholar] [CrossRef] [PubMed]
  228. Vidigal, V.M.; Silva, T.D.; de Oliveira, J.; Pimenta, C.A.M.; Felipe, A.V.; Forones, N.M. Genetic polymorphisms of vitamin D receptor (VDR), CYP27B1 and CYP24A1 genes and the risk of colorectal cancer. Int. J. Biol. Markers 2017, 32, 224–230. [Google Scholar] [CrossRef] [PubMed]
  229. Serrano, D.; Gnagnarella, P.; Raimondi, S.; Gandini, S. Meta-analysis on vitamin D receptor and cancer risk: Focus on the role of TaqI, ApaI, and Cdx2 polymorphisms. Eur. J. Cancer Prev. 2016, 25, 85. [Google Scholar] [CrossRef]
  230. Von Schuckmann, L.A.; Law, M.H.; Montgomery, G.W.; Green, A.C.; van der Pols, J.C. Vitamin D pathway gene polymorphisms and keratinocyte cancers: A nested case-control study and meta-analysis. Anticancer. Res. 2016, 36, 2145–2152. [Google Scholar]
  231. Wang, K.; Wu, G.; Li, J.; Song, W. Role of vitamin D receptor gene Cdx2 and Apa1 polymorphisms in prostate cancer susceptibility: A meta-analysis. BMC Cancer 2016, 16, 674. [Google Scholar] [CrossRef] [PubMed]
  232. Kang, S.; Zhao, Y.; Liu, J.; Wang, L.; Zhao, G.; Chen, X.; Yao, A.; Zhang, L.; Zhang, X.; Li, X. Association of Vitamin D receptor Fok I polymorphism with the risk of prostate cancer: A meta-analysis. Oncotarget 2016, 7, 77878. [Google Scholar] [CrossRef]
  233. Mi, Y.-Y.; Chen, Y.-Z.; Chen, J.; Zhang, L.-F.; Zuo, L.; Zou, J.-G. Updated analysis of vitamin D receptor gene FokI polymorphism and prostate cancer susceptibility. Arch. Med. Sci. 2017, 13, 1449–1458. [Google Scholar] [CrossRef]
  234. Lu, D.; Jing, L.; Zhang, S. Vitamin D receptor polymorphism and breast cancer risk: A meta-analysis. Medicine 2016, 95, e3535. [Google Scholar] [CrossRef]
  235. Laczmanski, L.; Lwow, F.; Osina, A.; Kepska, M.; Laczmanska, I.; Witkiewicz, W. Association of the vitamin D receptor FokI gene polymorphism with sex-and non-sex-associated cancers: A meta-analysis. Tumor Biol. 2017, 39, 1010428317727164. [Google Scholar] [CrossRef] [PubMed]
  236. Liu, S.; Cai, H.; Cheng, W.; Zhang, H.; Pan, Z.; Wang, D. Association of VDR polymorphisms (Taq I and Bsm I) with prostate cancer: A new meta-analysis. J. Int. Med. Res. 2017, 45, 3–10. [Google Scholar] [CrossRef] [PubMed]
  237. Iqbal, M.u.N.; Khan, T.A. Association between vitamin D receptor (Cdx2, Fok1, Bsm1, Apa1, Bgl1, Taq1, and Poly (A)) gene polymorphism and breast cancer: A systematic review and meta-analysis. Tumor Biol. 2017, 39, 1010428317731280. [Google Scholar] [CrossRef]
  238. Sheng, S.; Chen, Y.; Shen, Z. Correlation between polymorphism of vitamin D receptor TaqI and susceptibility to colorectal cancer: A meta-analysis. Medicine 2017, 96, e7242. [Google Scholar] [CrossRef] [PubMed]
  239. Chen, L.; Wei, J.; Zhang, S.; Lou, Z.; Wang, X.; Ren, Y.; Qi, H.; Xie, Z.; Chen, Y.; Chen, F.; et al. Association of VDR gene TaqI polymorphism with the susceptibility to prostate cancer in Asian population evaluated by an updated systematic meta-analysis. OncoTargets Ther. 2018, 11, 3267. [Google Scholar] [CrossRef]
  240. Yu, Z.-H.; Chen, M.; Zhang, Q.-Q.; Hu, X. The association of vitamin d receptor gene polymorphism with lung cancer risk: An update meta-analysis. Comb. Chem. High Throughput Screen. 2018, 21, 704–710. [Google Scholar] [CrossRef]
  241. Cho, Y.; Lee, J.; Oh, J.H.; Chang, H.J.; Sohn, D.K.; Shin, A.; Kim, J. Vitamin D receptor FokI polymorphism and the risks of colorectal cancer, inflammatory bowel disease, and colorectal adenoma. Sci. Rep. 2018, 8, 12899. [Google Scholar] [CrossRef]
  242. Pan, Z.; Chen, M.; Hu, X.; Wang, H.; Yang, J.; Zhang, C.; Pan, F.; Sun, G. Associations between VDR gene polymorphisms and colorectal cancer susceptibility: An updated meta-analysis based on 39 case-control studies. Oncotarget 2018, 9, 13068. [Google Scholar] [CrossRef]
  243. Chen, H.; Zhu, J. Vitamin D receptor rs2228570 polymorphism and susceptibility to ovarian cancer: An updated meta-analysis. J. Obstet. Gynaecol. Res. 2018, 44, 556–565. [Google Scholar] [CrossRef]
  244. Li, J.; Li, B.; Jiang, Q.; Zhang, Y.; Liu, A.; Wang, H.; Zhang, J.; Qin, Q.; Hong, Z.; Li, B.A. Do genetic polymorphisms of the vitamin D receptor contribute to breast/ovarian cancer? A systematic review and network meta-analysis. Gene 2018, 677, 211–227. [Google Scholar] [CrossRef]
  245. Mun, M.-J.; Kim, T.-H.; Hwang, J.-Y.; Jang, W.-C. Vitamin D receptor gene polymorphisms and the risk for female reproductive cancers: A meta-analysis. Maturitas 2015, 81, 256–265. [Google Scholar] [CrossRef] [PubMed]
  246. Lin, Z.-J.; Zhang, X.-L.; Yang, Z.-S.; She, X.-Y.; Xie, Y.; Xie, W.-J. Relationship between Vitamin D receptor gene polymorphism and renal cell carcinoma susceptibility. J. Cancer Res. Ther. 2018, 14, 820. [Google Scholar] [CrossRef] [PubMed]
  247. Li, M.; Liu, X.; Liu, N.; Yang, T.; Shi, P.; He, R.; Chen, M. Association between polymorphisms of Vitamin D receptor and lung cancer susceptibility: Evidence from an updated Meta-analysis. J. Cancer 2019, 10, 3639. [Google Scholar] [CrossRef] [PubMed]
  248. Laczmanski, L.; Laczmanska, I.; Lwow, F. Association of select vitamin D receptor gene polymorphisms with the risk of tobacco-related cancers—A meta-analysis. Sci. Rep. 2019, 9, 16026. [Google Scholar] [CrossRef] [PubMed]
  249. Birke, M.; Schoepe, J.; Wagenpfeil, S.; Vogt, T.; Reichrath, J. Association of vitamin D receptor gene polymorphisms with melanoma risk: A meta-analysis and systematic review. Anticancer Res. 2020, 40, 583–595. [Google Scholar] [CrossRef]
  250. Duan, G.-Q.; Zheng, X.; Li, W.-K.; Zhang, W.; Li, Z.; Tan, W. The association between VDR and GC polymorphisms and lung cancer risk: A systematic review and meta-analysis. Genet. Test. Mol. Biomark. 2020, 24, 285–295. [Google Scholar] [CrossRef]
  251. Gnagnarella, P.; Raimondi, S.; Aristarco, V.; Johansson, H.A.; Bellerba, F.; Corso, F.; Gandini, S. Vitamin D receptor polymorphisms and cancer. In Sunlight, Vitamin D and Skin Cancer; Springer: Berlin/Heidelberg, Germany, 2020; pp. 53–114. [Google Scholar]
  252. Gnagnarella, P.; Raimondi, S.; Aristarco, V.; Johansson, H.; Bellerba, F.; Corso, F.; Angelis, S.P.D.; Belloni, P.; Caini, S.; Gandini, S. Ethnicity as modifier of risk for Vitamin D receptors polymorphisms: Comprehensive meta-analysis of all cancer sites. Crit. Rev. Oncol./Hematol. 2021, 158, 103202. [Google Scholar] [CrossRef]
Figure 1. Schematic depiction of the VDR gene and protein. The q13–q14 arm of chromosome 12 contains the 48 kDa VDR protein-encoding gene. The VDR gene consists of a promoter and regulatory regions 1a through 1f. The VDR gene has eight exons (2–9) that together encode the six domains (A–F) of the complete VDR protein. The VDR protein has many domains, including dimerization, transactivation, hormone ligand binding, nuclear localization, and DNA-binding domains.
Figure 1. Schematic depiction of the VDR gene and protein. The q13–q14 arm of chromosome 12 contains the 48 kDa VDR protein-encoding gene. The VDR gene consists of a promoter and regulatory regions 1a through 1f. The VDR gene has eight exons (2–9) that together encode the six domains (A–F) of the complete VDR protein. The VDR protein has many domains, including dimerization, transactivation, hormone ligand binding, nuclear localization, and DNA-binding domains.
Cancers 16 03211 g001
Figure 2. Vitamin D metabolism. In the liver, the dietary forms of vitamin D (vitamin D2 and vitamin D3) are metabolized into calcidiol by vitamin D 25-hydroxylase. Calcidiol is further metabolized by 25-hydroxyvitamin D3-1α-hydroxylase into calcitriol in the kidney. Calcitriol (active form) regulates the downstream targets by genomic and non-genomic pathways. In the genomic pathway, calcitriol binds to cytosolic VDRs, which promotes the phosphorylation of VDRs and heterodimerization with RXR. The complex then binds to VDREs in the nucleus and regulates the mRNA expression of the target genes. In the non-genomic pathway, calcitriol specifically binds to membrane-bound VDRs and regulates calcium and MAPK signaling cascades.
Figure 2. Vitamin D metabolism. In the liver, the dietary forms of vitamin D (vitamin D2 and vitamin D3) are metabolized into calcidiol by vitamin D 25-hydroxylase. Calcidiol is further metabolized by 25-hydroxyvitamin D3-1α-hydroxylase into calcitriol in the kidney. Calcitriol (active form) regulates the downstream targets by genomic and non-genomic pathways. In the genomic pathway, calcitriol binds to cytosolic VDRs, which promotes the phosphorylation of VDRs and heterodimerization with RXR. The complex then binds to VDREs in the nucleus and regulates the mRNA expression of the target genes. In the non-genomic pathway, calcitriol specifically binds to membrane-bound VDRs and regulates calcium and MAPK signaling cascades.
Cancers 16 03211 g002
Figure 3. Key signaling pathways involved in the vitamin D-induced effects in different cancers. Antiproliferative effects of vitamin D are mediated by VDR-dependent and VDR-independent mechanisms. Vitamin D was shown to signal via various cell-surface receptors and impact the activation of telomerase, which is involved in the regulation of cancer cell proliferation and apoptosis. Vitamin D/VDR/VDRE/RXR complexes can trigger the expression of several cell cycle regulators and apoptosis modulators. Vitamin D was also shown to impact β-catenin-Tcf4 complex formation, which subsequently regulates cell cycle/apoptosis, Myelocytomatosis (Myc), T-cell factor (Tcf1), cluster of differentiation 44 (CD44), and Poly(ADP-ribose) glycohydrolase (PARG).
Figure 3. Key signaling pathways involved in the vitamin D-induced effects in different cancers. Antiproliferative effects of vitamin D are mediated by VDR-dependent and VDR-independent mechanisms. Vitamin D was shown to signal via various cell-surface receptors and impact the activation of telomerase, which is involved in the regulation of cancer cell proliferation and apoptosis. Vitamin D/VDR/VDRE/RXR complexes can trigger the expression of several cell cycle regulators and apoptosis modulators. Vitamin D was also shown to impact β-catenin-Tcf4 complex formation, which subsequently regulates cell cycle/apoptosis, Myelocytomatosis (Myc), T-cell factor (Tcf1), cluster of differentiation 44 (CD44), and Poly(ADP-ribose) glycohydrolase (PARG).
Cancers 16 03211 g003
Table 1. Recent observational studies with vitamin D analogs.
Table 1. Recent observational studies with vitamin D analogs.
Sl. No.Study DesignYearSample SizeConclusionReferences
1Case–Control Study2023293 (143 gastric cancer patients and 150 controls)VDR Fok1 polymorphism is significantly associated with GC risk in the Kashmiri population[48]
2Ancillary Study20231519 participants (vitamin D: n = 744; placebo: n = 775)Vitamin D supplementation in older adults with vitamin D deficiency has no effect on the telomere length[45]
3Case–Control Study2023204
(cases—102; controls—102)
Methylation levels of significant CpG sites in VDRs, CYP24A1, and CYP2R1 are inversely associated with CRC risk[41]
4Cohort Study2023236,382 participantsStudy showed the beneficial association of Serum 25(OH)D with risk of developing CRC. [40]
5Prospective Cohort Study 2023476 women with incident stage I–III breast cancer (BC)Women with sufficient vitamin D had smaller and lower-grade tumors compared to the women with insufficient vitamin D[46]
Table 2. The anticancer effects of vitamin D analogs.
Table 2. The anticancer effects of vitamin D analogs.
S. No.NameStructure *Type of
Cancer
Animal
Type
Induction
Method
Results
Obtained
1EB-1089
(Seocalcitol)
Cancers 16 03211 i001Breast cancer

HCC
Mice (I)

Mice (I)
Subcutaneously

Subcutaneously
Tumor growth inhibition
2HY-11Cancers 16 03211 i002Mice were inoculated with leukemia cellsMice (I)Intraperitoneally
3Tacalcitol
(PRI-2191)
Cancers 16 03211 i003Colorectal cancerMice (I)SubcutaneouslyTumor growth inhibition
4InecalcitolCancers 16 03211 i004Squamous cell carcinomaMice (I)SubcutaneouslyInhibition of tumor growth, increased apoptosis, and decreased proliferation
5TX527Cancers 16 03211 i005Kaposi’s sarcomaMice (I)SubcutaneouslyTumor growth Inhibition
6ParicalcitolCancers 16 03211 i006Metastatic breast
cancer
Mice (I)SubcutaneouslyTumor inhibition was accompanied by in vivo
upregulation of p21 and p27 expression
7DoxercalciferolCancers 16 03211 i007Neuroblastoma Mice (I)FlanksTumor growth inhibition
8MaxacalcitolCancers 16 03211 i008Cholangial carcinoma
Mice (I)SubcutaneouslyInhibition of tumor growth and inhibition of proliferation
9CalcipotriolCancers 16 03211 i009Non-melanoma skin cancerMice (I)SubcutaneouslyTumor growth inhibition
10BGP-13Cancers 16 03211 i010Colorectal cancer (CRC)Mice (I)SubcutaneouslyInhibition of growth of HT-29 tumors in mice
11PRI-2205Cancers 16 03211 i011Breast cancerMice (I)SubcutaneouslyLowering the expression of estrogen receptors and aromatase activity
12PRI-1906Cancers 16 03211 i012Breast cancerMice (I)OrthotopicallyTumor growth and
metastases inhibition
13BXL-01-0126Cancers 16 03211 i013Acute myeloid leukemiaMice (I)Intrahepatic (IH) or facial (FV) veinActivation of apoptosis
14BXL0124Cancers 16 03211 i014Breast cancer Mice (I)Mammary fat padsProliferation, angiogenesis, invasion, and metastasis
15Gemini0097Cancers 16 03211 i015Breast cancer Mice (I)Mammary fat padsSuppressed tumor growth and inhibition of tumor burden
16MART-10Cancers 16 03211 i016Pancreatic cancer Mice (I)SubcutaneouslyInhibition of tumor growth
17(1,25(OH)2D3-3-BE)Cancers 16 03211 i017Kidney cancer Mice (I)Subcutaneously to the flanksInhibition of tumor growth and increase in apoptosis
18Ro26-2198Cancers 16 03211 i018Colorectal cancer (CRC)Mice (C)Administration of Dextran sulfate sodium (DSS)Inhibition of dysplasia
progression and inhibition of
proliferation and
pro-inflammatory signals
19EM1Cancers 16 03211 i019Breast cancer Mice (I)Subcutaneously
Reduced the formation of metastasis
* The structure formulas have been sourced from PubChem and Tocris Bioscience [109,110]. I: implanted/injected, C: chemically induced.
Table 3. Studies depicting the role of VDR polymorphisms in various cancers.
Table 3. Studies depicting the role of VDR polymorphisms in various cancers.
S. No.VDR PolymorphismType of StudyCancer TypesStudy Outcome or Key FindingReference
1Apa1 (rs7975232), Cdx2 (rs11568820), and Taq1 (rs731236)Meta-analysis23 cancer typesCdx2 showed increased risk of cancer. Taq1 was associated with increased risk of CRC. Apa1 was not associated with cancer risk.[229]
2Apa1 (rs7975232), Bsm1 (rs1544410), Bgl1 (rs739837), and Fok1 (rs2228570)Nested case–control study and meta-analysisKeratinocyte cancersVDR polymorphisms may be associated with the risk of keratinocyte cancers.[230]
3ApaI1 (rs7975232) and Cdx2 (rs11568820)Meta-analysisProstate cancerVDR Cdx2 and Apa1 polymorphisms were not associated with prostate cancer.[231]
4Fok1 (rs10735810)Meta-analysisProstate cancerVDR Fok1 polymorphism could be a promising target and might be capable of causing prostate cancer risk.[232]
5Fok1 (rs10735810)Meta-analysisProstate cancerVDR Fok1 polymorphism may contribute to the risk of developing prostate cancer in Caucasian and population-based studies.[233]
6Apa1 (rs7975232), Bsm1 (rs1544410), Fok1 (rs2228570), and Taq1 (rs731236)Meta-analysisBreast cancerVDR Fok1, Bsm1, Taq1, and Apa1 polymorphisms were not associated with the risk of breast cancer in the general as well as Caucasian population.[234]
7Fok1 (rs10735810)Meta-analysisSex- and non-sex-associated cancersFok1 polymorphism was associated with breast and ovarian cancers.[235]
8Bsm1 (rs1544410) and Taq1 (rs731236)Meta-analysisProstate cancerTaq1 was significantly associated with risk of prostate cancer in Asians and African Americans but not Bsm 1 polymorphism.[236]
9Apa1, Bsm1, BgI1, Cdx2, Fok1, Taq1, and Poly (A)Systematic review and meta-analysisBreast cancerVDR gene polymorphisms (Bsm1, Apa1, Fok1, and Poly (A)) may increase susceptibility to breast cancer development.[237]
10Taq1Meta-analysisCRCThere was no correlation between Taq1 polymorphisms and susceptibility to CRC.[238]
11Taq1Systematic meta-analysisProstate cancerThe VDR Taq1 polymorphism might be associated with risk of prostate cancer in Asian (especially Japanese) populations.[239]
12Apa1 (rs7975232), Bsm1 (rs1544410), Fok1 (rs10735810), and Taq1 (rs731236)Meta-analysisLung cancerVDR genetic polymorphism may be correlated with the risk of lung cancer.[240]
13Fok1 (rs2228570)Systematic meta-analysisCRCRole of VDR Fok1 polymorphism may differ based on the type and severity of colorectal disease.[241]
14Apa1, Fok1, Bsm1, Taq1, and Cdx2Meta-analysisCRCBsm1 polymorphism was associated with CRC risk, and Fok1 might be a risk factor for CRC.[242]
15Fok1 (rs2228570)Meta-analysisOvarian cancerFok1 polymorphism increased the risk of ovarian cancer in Caucasian populations in a dominant genetic model.[243]
16Bsm1 (rs1544410), Cdx2, and Fok1 (rs2228570)A systematic review and network meta-analysisBreast and ovarian cancersFok1 and Bsm1 polymorphism are likely the best genetic model for detecting the risk of breast and ovarian cancers, respectively, in Caucasian patients.[244]
17Apa1, Fok1, Bsm1, Taq1, and Cdx2Meta-analysisFemale reproductive cancersFok1 and Bsm1 VDR gene polymorphisms may be significantly associated with gynecological cancers.[245]
18Apa1 (rs7975232), Fok1 (rs2228570), Bsm1 (rs1544410), and Taq1 (rs731236)Meta-analysisRCCApaI gene polymorphism and Fok1 FF genotype were associated with RCC susceptibility in Asians.[246]
19Apa1 (rs7975232 C > A), Bsm1 (rs1544410 G > A), Cdx2 (rs11568820 T > C), and Taq1 (rs731236 T > C)Meta-analysisLung cancerBsm1, Taq1, and Cdx-2 polymorphisms may contribute to lung cancer susceptibility.[247]
20Apa1 (rs7975232), Bsm1 (rs1544410), Fok1 (rs10735810), and Taq1 (rs731236)Meta-analysisTobacco-related cancersTaq1 polymorphism and the risk of tobacco-related cancers were correlated with each other.[248]
21A-1012G (rs4516035), Apa1 (rs7975232), Bsm1 (rs1544410), BgI1 (rs739837), Cdx2 (rs11568820), Fok1 (rs2228570), and Taq1 (rs731236)Systematic review and meta-analysisMelanomaApa1, Bsm1, and Fok1 polymorphisms may influence the development of melanoma.[249]
22Apa1 (rs7975232), Bsm1 (rs1544410, A/G), Cdx2 (rs11568820, C/T), Fok1 (rs2228570, T/C), and Taq1 (rs731236, T/C)Systematic review and meta-analysisLung cancerBsm1 and Cdx2 polymorphisms decreased lung cancer risk, while Taq1 increased it.[250]
23Apa1, Bsm1, Cdx2, Fok1, and Taq1 Systematic review and meta-analysis18 cancer typesSignificant associations with VDR polymorphisms have been reported for prostate (Fok1, Bsm1, Taq1, Apa1, and Cdx2), breast (Fok1, Bsm1, Taq1, Apa1, and Cdx2), colorectal (Fok1, Bsm1, Taq1, and Apa1), and skin cancer (Fok1, Bsm1, and Taq1).[251]
24Apa1 (rs7975232), Bsm1 (rs1544410), Cdx2 (rs11568820), Fok1 (rs10735810), and Taq1 (rs731236)Comprehensive meta-analysis22 cancer typesVDR polymorphisms were linked to cancer susceptibility. Ethnicity may be a modifier of cancer risk, in particular for hormone-dependent cancers.[252]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Dallavalasa, S.; Tulimilli, S.V.; Bettada, V.G.; Karnik, M.; Uthaiah, C.A.; Anantharaju, P.G.; Nataraj, S.M.; Ramashetty, R.; Sukocheva, O.A.; Tse, E.; et al. Vitamin D in Cancer Prevention and Treatment: A Review of Epidemiological, Preclinical, and Cellular Studies. Cancers 2024, 16, 3211. https://doi.org/10.3390/cancers16183211

AMA Style

Dallavalasa S, Tulimilli SV, Bettada VG, Karnik M, Uthaiah CA, Anantharaju PG, Nataraj SM, Ramashetty R, Sukocheva OA, Tse E, et al. Vitamin D in Cancer Prevention and Treatment: A Review of Epidemiological, Preclinical, and Cellular Studies. Cancers. 2024; 16(18):3211. https://doi.org/10.3390/cancers16183211

Chicago/Turabian Style

Dallavalasa, Siva, SubbaRao V. Tulimilli, Vidya G. Bettada, Medha Karnik, Chinnappa A. Uthaiah, Preethi G. Anantharaju, Suma M. Nataraj, Rajalakshmi Ramashetty, Olga A. Sukocheva, Edmund Tse, and et al. 2024. "Vitamin D in Cancer Prevention and Treatment: A Review of Epidemiological, Preclinical, and Cellular Studies" Cancers 16, no. 18: 3211. https://doi.org/10.3390/cancers16183211

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

Article Metrics

Back to TopTop