You are currently viewing a new version of our website. To view the old version click .
NeuroSci
  • Systematic Review
  • Open Access

16 December 2025

The Role of Vitamin D in Parkinson’s Disease: Evidence from Serum Concentrations, Supplementation, and VDR Gene Polymorphisms

and
Neurology Department, Cooper University Hospital, Camden, NJ 08103, USA
*
Author to whom correspondence should be addressed.
This article belongs to the Special Issue Parkinson's Disease Research: Current Insights and Future Directions

Abstract

Background/aim: Vitamin D (VitD) has been implicated in neuroprotection, yet its role in Parkinson’s disease (PD) remains unclear. This systematic review and meta-analysis aimed to evaluate the association between VitD status, supplementation, and vitamin D receptor (VDR) gene polymorphisms with PD risk and outcomes. Methodology: Following PRISMA guidelines, we searched PubMed, Scopus, and Google Scholar through August 2025 for observational studies, clinical trials, and genetic association studies. Primary outcomes included serum VitD levels in PD versus healthy controls (HCs), prevalence of VitD insufficiency/deficiency, and effects of VitD supplementation on motor symptoms. Secondary outcomes assessed associations between VDR polymorphisms and PD susceptibility. Data were synthesized using random- and fixed-effects models, with heterogeneity and publication bias evaluated. PROSPERO (CRD420251133875). Results: Sixty-three studies (n ≈ 10,700 participants) met inclusion criteria. PD patients exhibited significantly lower VitD levels (SMD = −0.46; 95% CI: −0.51 to −0.41) and higher odds of insufficiency (OR = 1.52) and deficiency (OR = 2.20) compared to HC. Cohort data suggested sufficient VitD may reduce PD risk (HR = 0.83). Supplementation yielded modest, non-significant improvements in motor outcomes. Among 20 genetic studies, FokI (rs2228570) was most consistently associated with PD, while other VDR SNPs showed variable or null associations. Conclusions: VitD deficiency is common in PD and may influence disease risk and motor function. Current evidence indicates limited benefit of supplementation for motor outcomes, and genetic associations remain inconsistent.

1. Introduction

PD is a progressive neurodegenerative disorder primarily characterized by the loss of dopaminergic neurons in the substantia nigra pars compacta, leading to striatal dopamine depletion and disruption of basal ganglia circuitry. It is the second most prevalent neurodegenerative disorder worldwide, with its incidence steadily increasing due to aging populations and extended life expectancy [1]. The neurochemical imbalance manifests clinically as cardinal motor symptoms, including bradykinesia, resting tremor, rigidity, and postural instability. Beyond motor dysfunction, PD encompasses a wide spectrum of non-motor features such as cognitive impairment, mood disorders, autonomic dysfunction, and sleep disturbances, which significantly impact quality of life [2].
The etiology of PD is multifactorial, encompassing both genetic predispositions and environmental influences. Recent research has emphasized the importance of modifiable lifestyle and environmental factors in shaping disease risk and progression [3]. Among these factors, nutritional status and particularly vitamin intake has attracted increasing interest from the scientific community and general public. Vitamins are essential bioactive compounds with antioxidant properties that support normal physiological and neurological function. Of particular relevance is VitD, which has emerged as a potential neuroprotective agent (Figure 1) [4].
Figure 1. Vitamin D metabolism and receptor (VDR). Abbreviations: Ca, calcium; CUBN, cubilin; DBP, vitamin D binding protein; DHCR7, 7-dehydrocholesterol reductase; hv, light energy; L3, lumisterol 3; LRP2, low-density lipoprotein-related protein 2; MAPK, mitogen-activated protein kinase; NFAT1, nuclear factor of activated T cells; NF-κB, nuclear factor kappa-light-chain-enhancer of activated B cells; PDIA3 (ERp57), protein disulfide isomerase family A member 3; PKC, protein kinase C; PLAA, phospholipase A2 activating protein; PLA2, phospholipase A2; RXR, retinoid X receptor; SHH, sonic hedgehog; SRC, steroid receptor coactivator; STAT 1-3, signal transducers and activators of transcription 1 and 3; T3, tachysterol 3; VDR, vitamin D receptor; VDRE, vitamin D response element; WNT, wingless-type MMTV integration site family; △T, temperature change; ↑, increase/upregulate; ↓, decrease/downregulate.
One of the earliest reports suggesting a link between VitD deficiency and PD in humans was published by Derex et al. in 1997, who at the time considered the association to be rare [5]. In this context, emerging evidence suggests that VitD may play a neuroprotective role by mitigating dopaminergic neuron loss and enhancing both motor and cognitive outcomes [6]. In vitro studies revealed that VitD is related to enhanced mesencephalic cell viability [7] and survival [8] in 6-OHDA PD model. In vivo, its deficiency has been linked to increased risk of PD [9], and one of the hypotheses is related to the VitD’s role in immune modulation [10].
VitD exerts neuroprotective effects through multiple molecular pathways relevant to PD. It modulates calcium homeostasis, reduces oxidative stress by upregulating antioxidant enzymes, and attenuates neuroinflammation via inhibition of NF-κB signaling and suppression of pro-inflammatory cytokines such as TNF-α and IL-6 [11]. Additionally, vitamin D influences apoptotic pathways by promoting anti-apoptotic proteins and reducing caspase activation, thereby supporting neuronal survival. These mechanisms collectively help preserve dopaminergic neurons in the substantia nigra pars compacta, the primary site of degeneration in PD [12]. Importantly, vitamin D receptor (VDR) and activating enzymes such as 1α-hydroxylase are highly expressed in dopaminergic neurons of the substantia nigra and glial cells, suggesting a direct role in motor control and neuroimmune regulation [13]. Dysregulation of VDR signaling, whether through deficiency or genetic polymorphisms (e.g., FokI, BsmI), may impair dopamine synthesis and neuronal survival, contributing to both motor symptoms (bradykinesia, rigidity) and non-motor features (cognitive decline, mood disorders) [14].
Although Wang et al. found no causal relationship between genetically predicted VitD levels and PD using bidirectional Mendelian randomization across three large cohorts, their findings do not eliminate the relevance of VitD in PD research [15]. Observational studies consistently report lower serum VitD levels in PD patients, and VDR are present in dopaminergic neurons, suggesting potential neuroprotective roles [16]. Moreover, VitD may influence PD progression through immune modulation, antioxidant activity, and calcium homeostasis—mechanisms not fully captured by genetic analyses. Therefore, further longitudinal and mechanistic studies are warranted to explore VitD’s diagnostic, prognostic, and therapeutic potential in PD beyond genetic predisposition. This study aims to elucidate the role of VitD in PD by examining its involvement across three key domains: serum concentrations, therapeutic supplementation, and genetic variability.

2. Methodology

2.1. Primary and Secondary Outcomes

The primary outcome was to (1) compare serum VitD levels between patients with PD and HC. Secondary outcomes included (2) comparisons of the prevalence of VitD deficiency and insufficiency between these groups, as well as the evaluation of the efficacy of VitD supplementation versus placebo on motor symptoms in patients with PD. An additional outcome was the analysis of (3) genotype frequencies of VDR polymorphisms and their potential influence on the risk of developing PD.

2.2. Literature Search Strategy

This meta-analysis followed the PRISMA statement (Table S1—PRISMA Statement) [17]. Two reviewers (J.P.R. and A.L.F.C.) systematically searched the electronic PubMed/Medline, Scopus, and Google Scholar databases by the following search terms: vitamin D, calciferol, and Parkinson’s disease (Table S2—Search Strategy). The search included studies published from database inception through 29 August 2025, with no restrictions on language. The protocol of this study was registered on PROSPERO (CRD420251133875).

2.3. Inclusion and Exclusion Criteria

Studies were considered eligible if they met the following criteria. (1) Observational studies, including case–control, cohort, and longitudinal designs, that investigated serum VitD levels, deficiency, or insufficiency in individuals with PD and HC were included. (2) Clinical trials assessing the efficacy of VitD supplementation compared to placebo in patients with PD were also eligible. Additionally, (3) case–control studies examining the association between vitamin D receptor (VDR) gene polymorphisms and susceptibility to PD were considered. (4) Only studies in which PD was clinically diagnosed by a neurologist according to the United Kingdom Parkinson’s Disease Brain Bank criteria were included. (5) Studies were required to report sufficient data on sample size, serum VitD levels, and genotypic distribution of the VDR gene. (6) All participants had to be adults aged 18 years or older.
(1) Studies were excluded if they were reviews, case reports, meta-analyses, editorials, or commentaries. (2) Studies lacking detailed information on genotyping, VitD status, or population characteristics were excluded. (3) Research focusing on genes other than VDR or reporting additional VDR genotypes not relevant to the analysis was also excluded. (4) Animal studies and studies that did not assess patients clinically diagnosed with PD were not considered.
No language restriction was applied. In cases where the non-English literature or when the English abstract provided insufficient information, as observed in articles written in Turkish, Chinese, Russian, the Google Translate service was employed [18].

2.4. Data Extraction

Two investigators (J.P.R. and A.L.F.C.) independently reviewed the literature and extracted relevant data. For each study, information was collected on the first author, country of origin, and year of publication. For studies comparing PD patients and HC, data were extracted on sample sizes, serum VitD levels, and the number of individuals classified as having VitD insufficiency or deficiency. In studies reporting incidence, the investigators recorded sample size, number of incident PD cases, VitD status, and HR. For clinical trials assessing VitD supplementation, data were extracted on sample size and motor outcomes, including UPDRS-Total, UPDRS-III, and Timed Up and Go (TUG) scores, all measured during the “ON” medication period. Any discrepancies between reviewers were resolved through discussion.

2.5. Quality Assessment

The Newcastle-Ottawa Scale (NOS) was used to evaluate the process in terms of case selection, comparability of queues, and evaluation of results. A study with a score of at least six was considered as high-quality literature. Higher NOS scores showed higher literature quality. The NOS was used to rate the quality of observational cohort and case–control studies and an adapted version of the NOS for quality of cross-sectional studies [19].
The risk of bias in randomized controlled trials (RCTs) was assessed using the ROBVIS tool, a web-based application designed to visually summarize risk of bias judgments based on established frameworks such as the Cochrane Risk of Bias tool [20].

2.6. Statistical Analysis

Statistical analyses were conducted to compare outcomes between individuals with PD and HC. Data extraction and organization were performed using Microsoft Excel (version 2023 for macOS). Serum VitD levels were reported as means with SD, and SMD were calculated using Cohen’s d to assess differences in VitD levels between PD and HC groups. For studies lacking reported SDs, values were estimated assuming a large effect size (Cohen’s d ≥ 0.8). Ninety-five percent confidence intervals (CIs) were derived using the standard error formula for Cohen’s d and the normal approximation. To ensure consistency with clinical standards, VitD concentrations originally reported in nanomoles per liter (nmol/L) were converted to nanograms per milliliter (ng/mL) using the molecular weight of the analyte.
To evaluate the association between VitD status (insufficiency and deficiency) and PD occurrence, ORs were calculated. For studies assessing the effect of VitD supplementation, SMDs were computed to compare motor outcomes between treatment and placebo groups, using pooled SDs and group means with sample sizes. For continuous outcomes, pooled SMDs were calculated using the inverse variance method, with between-study variance (τ2) estimated via the DerSimonian–Laird approach [21]. Confidence intervals for τ2 and τ were computed using the Jackson method. Heterogeneity was assessed using the I2 statistic derived from Cochran’s Q, and prediction intervals were calculated using a t-distribution with 2 degrees of freedom to estimate the expected range of effects in future studies. For binary outcomes, ORs with 95% CIs were pooled using both fixed-effect and random-effects models. The Mantel–Haenszel method was applied for fixed-effect analysis, while the inverse variance method was used for random-effects modeling, incorporating τ2 estimated via DerSimonian–Laird. The Mantel–Haenszel estimator was also used in the calculation of Q and τ2. Prediction intervals for binary outcomes were calculated using a t-distribution with 15 degrees of freedom. Heterogeneity was again assessed using I2, with values exceeding 50% indicating substantial variability [22].
To run all the above-mentioned analyses and demonstrate results via forest plots, we carried out initial calculations using Meta-Mar (version 4.1.2, https://www.meta-mar.com/, (accessed on 17 September 2025)), a free online meta-analysis service developed by Beheshti et al. [23]. Forest plots for VDR SNPs were generated using GraphPad Prism (version 10.4.1 for macOS; GraphPad Software, Boston, MA, USA; www.graphpad.com). For calculations of ORs and additive genetic models the analyses were performed using SPSS software (version 22.0; IBM Corp., Armonk, NY, USA). A p-value of less than 0.05 was considered statistically significant.
Genotypic data were analyzed to assess HWE in control groups using Fisher’s exact test for allele frequencies and Pearson’s chi-square test (df = 1). All polymorphisms were evaluated using the forward DNA strand to ensure data standardization. To investigate the association between VDR SNPs and PD, multiple genetic models were applied, including allele, dominant, recessive, and additive models. Comparisons were made using ORs, 95% CIs, and p-values. The allele model compared A (dominant) vs. a (recessive); the dominant model compared AA vs. Aa + aa; the recessive model compared aa vs. AA + Aa; and the additive model treated genotypes as ordinal variables (AA = 0, Aa = 1, aa = 2), assuming a linear increase in risk per recessive allele. Logistic regression was used to estimate additive effects, yielding ORs per allele with corresponding CIs and p-values. For studies lacking isolated allele counts, allele frequencies were calculated using the following formula: A alleles = (2 × AA) + (1 × Aa); a alleles = (2 × aa) + (1 × Aa) [24].

2.7. Definitions

According to the 2011 Endocrine Society Clinical Practice Guideline, serum 25(OH)D3 levels were categorized as follows: deficiency was defined as <20 ng/mL (50 nmol/L), insufficiency as 21–29 ng/mL (52.5–72.5 nmol/L), and sufficiency as ≥30 ng/mL (≥75 nmol/L) [25]. However, the updated 2024 guideline no longer endorses these thresholds for generally healthy individuals, citing insufficient evidence from randomized controlled trials to support specific cutoffs for clinical benefit [26].
Genetic analysis focused on several SNPs within the VDR gene and related loci (Table S3–SNPs). The following polymorphisms were included: ApaI (rs7975232), A-1012G (rs4516035), BsmI (rs1544410), Cdx2 (rs11568820), FokI (rs2228570, also known as rs10735810), TaqI (rs731236), and Tru9I (rs757343). Additional SNPs analyzed included rs1989969, rs2853559, rs4334089, rs7299460, rs7968585, rs7976091, and rs10083198. Both the significant SNPs identified through GWAS and those incorporated into the optimal PRS model were annotated using the online tool SnpXplorer [27].
A variety of assays have been used to measure serum VitD levels in PD studies, each with distinct sensitivity and specificity profiles (Table S4–Assays) [28,29,30,31,32,33,34,35]. Traditional methods like the Competitive Protein Binding Assay (CBPA) and Radioimmunoassay offer foundational approaches but are less commonly used today due to limitations in accuracy. More advanced techniques such as Liquid Chromatography–Mass Spectrometry (LC-MS) and LC-MS/MS provide high precision and are considered gold standards for quantifying VitD metabolites. Immunoassay-based methods, including ELISA, Electrochemiluminescence Immunoassay (ECL), and Chemiluminescent Microparticle Immunoassay, are widely used in clinical settings due to their efficiency and scalability, though they may vary in cross-reactivity and calibration standards. The choice of assay could have influenced reported VitD levels and should be considered when comparing results across studies.

3. Results

3.1. Study Selection

An initial screening of PubMed/Medline, Scopus, and Google Scholar yielded 2575 articles, of which 63 met the predefined inclusion criteria. A comprehensive summary of the study selection process is illustrated in the following Figure 2. There were 35 articles about VitD status [28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62], 8 clinical trials [63,64,65,66,67,68,69,70], and 20 assessing VDR SNPs [40,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89]. A total of 192 analyses were performed across the studies.
Figure 2. PRISMA flowchart for the identification of included studies.

3.2. Quality Assessment

The studies assessing VitD status had NOS scores ranging from 4 to 9, with a mean score of 7.26 and a median of 7.5 (Table S5–NOS Levels) [28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62]. Of the 34 studies included, 26 were classified as high quality (NOS ≥ 7), 7 as moderate quality (NOS 5–6), and 1 as low quality (NOS < 5). Risk of bias was further evaluated using the ROBVIS tool, which indicated that 2 studies were rated as low risk, 5 had some concerns, and 1 was rated as high risk (Table S6–Robvis) [63,64,65,66,67,68,69,70].
For the genetic polymorphism studies, NOS scores ranged from 5 to 8, with a mean of 6.40 and a median of 6.0. Among the 20 studies assessed, 6 were classified as high quality, 14 as moderate quality, and none were considered low quality (Table S7–NOS Polymorphism) [40,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89].

3.3. Meta-Analysis

3.3.1. VitD Serum in PD

A meta-analysis of 33 studies involving 3920 individuals with PD and 6792 HC demonstrated a consistent pattern of lower serum VitD levels in PD patients (Figure 3) (Table S8–Serum 25(OH)D) [28,29,30,31,32,33,34,35,36,37,38,40,41,42,43,44,45,46,47,48,50,51,52,53,54,55,56,57,58,60,61,62,86]. Employing the DerSimonian–Laird estimator and inverse variance weighting, the fixed effect model yielded an SMD of −0.46 (95% CI: −0.51 to −0.41), while the random effects model produced a slightly larger effect size of −0.59 (95% CI: −0.71 to −0.46). I2 value was 82%, Cochran’s Q test was significant (Q = 177.69, p < 2.2 × 10−16), suggesting notable variability among studies. Publication bias was assessed using multiple methods, all of which indicated significant small-study effects: Egger’s test (p = 0.00086), Begg’s test (p = 0.00157), and the Thompson-Sharp test (p = 1.76 × 10−5). The Trim-and-Fill method estimated 10 potentially missing studies, adjusting the effect size to −0.37 (95% CI: −0.51 to −0.24), representing a 36% reduction in magnitude.
Figure 3. Forest plots of the vitamin D levels in patients with Parkinson’s disease versus controls.

3.3.2. VitD Insufficiency in PD

A meta-analysis was conducted on 12 studies reporting binary outcomes, utilizing the metabin model and Odds Ratio (OR) as the summary measure (Figure 4) (Table S9–25(OH)D Insufficiency) [29,30,31,33,34,35,42,44,48,49,53,58]. The analysis applied the DerSimonian–Laird estimator for between-study variance and employed the Mantel–Haenszel method for pooling. The fixed effect model produced a statistically significant OR of 1.52 (95% CI: 1.33 to 1.74), while the random effects model yielded an OR of 1.43 (95% CI: 1.05 to 1.96), accompanied by a wide prediction interval (0.51 to 4.05), indicating uncertainty in the estimate. I2 statistic was 75%, Cochran’s Q test was significant (Q = 43.65, p = 8.36 × 10−6), indicating notable variability among the studies. Assessment of publication bias through Egger’s test, Begg’s test, the Thompson-Sharp test, and the Trim-and-Fill method revealed no significant evidence of bias, with all p-values exceeding the 0.05 threshold and no missing studies identified. These findings indicate a statistically significant effect size with some heterogeneity, but no apparent publication bias, providing a reliable synthesis of the included studies.
Figure 4. Forest plots of the vitamin D insufficiency in patients with Parkinson’s disease versus controls.

3.3.3. VitD Deficiency in PD

A meta-analysis included 17 studies investigating the association between VitD deficiency and PD compared to HC (Figure 5) (Table S10–25(OH)D Deficiency) [29,31,32,34,35,39,41,42,43,44,47,48,49,51,53,58,59]. The fixed effect model yielded a statistically significant OR of 2.20 (95% CI: 1.91 to 2.53), while the random effects model produced a slightly higher OR of 2.38 (95% CI: 1.79 to 3.17), with a wide prediction interval (0.85 to 6.66), indicating variability across studies. Heterogeneity assessment revealed a τ2 of 0.21 and an I2 of 69%, Cochran’s Q statistic (Q = 51.35, p = 1.395 × 10−5) indicated significant variability. Publication bias analysis showed mixed results: Egger’s test was non-significant (p = 0.17), while Begg’s test and the Thompson-Sharp test indicated significant bias (p = 0.03 and 0.02, respectively). The Trim-and-Fill method estimated five potentially missing studies, adjusting the effect size from 0.86 to 0.63, reflecting a 27.2% reduction and suggesting the presence of publication bias.
Figure 5. Forest plots of the vitamin D deficiency in patients with Parkinson’s disease versus controls.

3.3.4. VitD Serum and PD Incidence

The comparative analysis of two large cohort studies investigated the relationship between VitD status and the risk of developing PD (Table S11–PD Incidence) [90,91]. Gröninger et al. reported a HR of 1.08 (95% CI: 0.80–1.45) for individuals with VitD deficiency, suggesting a slightly increased but statistically non-significant risk [90]. In contrast, Veronese et al. found a significantly reduced risk of PD among individuals with sufficient VitD levels, with an HR of 0.83 (95% CI: 0.74–0.93) [91]. These findings imply that maintaining adequate VitD levels may offer a protective effect against PD, while deficiency does not show a strong or conclusive association with increased risk. The standardized incidence rates further support this trend, with higher PD occurrence observed in the sufficient group (662.31 vs. 328.56 per 100,000 people), likely reflecting differences in sample size and follow-up duration rather than contradicting the protective association.

3.3.5. VitD Supplementation in PD

The effects of VitD supplementation versus placebo on motor outcomes in patients with PD were assessed across eight studies, comprising a total of 284 individuals with PD and 282 HC (Figure 6) (Table S12–VitD Supplementation) [63,64,65,66,67,68,69,70]. For UPDRS Total scores during “ON” periods, the fixed effect model yielded an SMD of −0.08 (95% CI: −0.30 to 0.14), while the random effects model produced a slightly larger effect size of −0.39 (95% CI: −1.10 to 0.33). The I2 value was 89%, and Cochran’s Q test was significant (Q = 27.52, p < 4.57 × 10−6). For UPDRS-III scores, the fixed effect model yielded an SMD of −0.10 (95% CI: −0.27 to 0.07), while the random effects model produced a slightly larger effect size of −0.18 (95% CI: −0.49 to 0.12). The I2 value was 63%, and Cochran’s Q test was significant (Q = 16.41, p < 1.17 × 10−2). Regarding TUG performance, the fixed effect model yielded an SMD of −0.18 (95% CI: −0.43 to 0.06), while the random effects model produced a slightly larger effect size of −0.17 (95% CI: −0.44 to 0.10). The I2 value was 8%, and Cochran’s Q test was significant (Q = 3.27, p < 3.51 × 10−1). These findings suggest that VitD supplementation may have modest and variable effects on motor function in PD, with the most consistent trend observed in TUG performance.
Figure 6. Forest plot of the effect of vitamin D on motor symptoms in patients with Parkinson’s disease compared to healthy controls. TUG, Timed Up and Go test; UPDRS, Unified Parkinson Disease Rating Scale; UPDRS—III, Part III (Motor) of UPDRS.

3.3.6. VitD Polymorphism

Several SNPs showed varying degrees of association with PD (Figure 7) (Table S13–Polymorphism) [40,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89]. ApaI (rs7975232) was widely studied, with some significant findings in allele and additive models, though overall results were inconsistent. BsmI (rs1544410) also showed sporadic significance, particularly in the additive and recessive models in studies by Kim et al. [71] and Gatto et al. [92]. In contrast, SNPs like A-1012G (rs4516035), BglI (rs739837), and Cdx2 (rs11568820) consistently showed no significant associations across all genetic models.
Figure 7. Forest plots showing odds ratios and 95% confidence intervals for vitamin d-related SNPs in Parkinson’s disease compared with healthy controls under allelic, dominant, recessive, and additive genetic models. Bolded results indicate statistically significant associations with p < 0.05.
The most robust and consistent associations were observed with FokI (rs2228570). Multiple studies, including Han et al. [72], Török et al. [75], Hu et al. [84], and Agliardi et al. [85], reported significant results in allele and additive models, with overall meta-analytic significance (e.g., OR = 1.15, p = 0.03 in recessive model). TaqI (rs731236) showed mixed results, with Gatto et al. reporting significant associations in allele and additive models [92], while other studies did not replicate these findings. rs4334089 also showed significance in specific studies, particularly in allele and additive models reported by Fazeli et al [78].
Other SNPs such as Tru9I (rs757343), rs1989969, rs2853559, rs7299460, and rs7968585 generally lacked significant associations. However, rs7976091 showed a notable additive effect in Kang et al. (OR = 2.24, p < 0.01) [79], and rs10083198 demonstrated a strong additive association (OR = 2.05, p < 0.01) in Lin et al.’s study [76].
The HWE results show that several SNPs, including ApaI (rs7975232), FokI (rs2228570), and TaqI (rs731236), exhibited significant deviations from HWE in multiple studies, particularly in case groups, suggesting possible genetic disequilibrium or sampling biases. Conversely, SNPs such as A-1012G (rs4516035), BglI (rs739837), and Cdx2 (rs11568820) generally maintained equilibrium across studies, with p-values well above 0.05.

4. Discussion

4.1. Summary of Results

The current systematic review and meta-analysis of 63 studies investigated the relationship between VitD and PD, including 35 on VitD status, 8 clinical trials, and 20 on VDR SNPs. Quality assessment revealed that most studies were of moderate to high quality, with VitD studies averaging a NOS score of 7.26 and genetic studies averaging 6.40. Meta-analyses showed that PD patients had significantly lower serum VitD levels (SMD = −0.46), higher odds of VitD insufficiency (OR = 1.52), and deficiency (OR = 2.20), with high heterogeneity across studies. Cohort data suggested that sufficient VitD levels may reduce PD risk (HR = 0.83), while deficiency showed no strong association. VitD supplementation had modest effects on motor outcomes, with the most consistent improvement seen in TUG performance. Genetic analyses identified FokI (rs2228570) as the most consistently associated SNP with PD, while ApaI and BsmI showed variable significance. Other SNPs like A-1012G, BglI, and Cdx2 showed no consistent associations. HWE analysis revealed deviations in SNPs such as ApaI, FokI, and TaqI, particularly in case groups, suggesting potential genetic imbalance or sampling bias.

4.2. Motor and Non-Motor PD Symptoms

There have been increasing reports linking both motor and non-motor symptoms of PD with serum VitD levels (Table 1) [45,48,49,67,91,93,94,95,96,97,98,99,100,101,102,103,104]. In this chapter, we will review selected studies that explore these associations, highlighting the potential role of VitD in the clinical presentation and progression of PD.
Table 1. Summary of associations between VitD and PD risk, symptoms, and genetic factors.
Iranian patients with PD showed significantly lower serum 25(OH)D3 levels compared to normal ranges. This deficiency was notably linked to increased postural instability, abnormal posture, and freezing episodes [95]. Similarly, an Egyptian study found that VitD levels were negatively correlated with age and age at onset; they were not significantly associated with disease duration or severity (UPDRS and H&Y) [49]. Zhang et al. reported that lower VitD levels were significantly associated with increased falls, insomnia, higher scores for depression and anxiety, and poor sleep quality, but not with UPDRS scores or disease duration [48]. Similarly, Meamar et al. found that serum VitD levels were negatively correlated with UPDRS scores (r = −0.34, p < 0.05) [102]. Peterson et al. showed that VitD levels were correlated with motor symptom severity and gait features, such as automatic postural response strength and stance weight asymmetry during backward perturbation tasks [96].
In the oxford discovery cohort of early PD patients, serum VitD levels averaged 49.1 nmol/L (SD 25.0), with lower levels significantly associated with worse baseline activities of daily living (UPDRS-II intercept: −0.75, p = 0.005, q = 0.007). However, VitD did not predict motor or cognitive progression over time (UPDRS-III slope: p = 0.91, q = 0.91) [103]. Similarly, Chitsaz et al. found a high prevalence of VitD insufficiency (72.8%) and deficiency (38.4%) among Iranian PD patients, but no significant association between serum 25(OH)D3 levels and disease severity as measured by H&Y or UPDRS-III scores [105].
Kim et al. showed that VitD levels were independently associated with olfactory dysfunction and NMSS, with stronger associations observed in the akinetic-rigid subtype [101]. Kwon et al. revealed that low serum VitD levels were independently correlated with delayed gastric emptying in drug-naive, de novo PD patients [100]. Also, Canlı et al. found a strong inverse correlation between fatigue severity and both VitD concentration (r = −0.851, p < 0.001) and physical activity (r = −0.757, p < 0.001), while disease stage (r = 0.879, p < 0.001) and anxiety levels (r = 0.797, p < 0.001) were positively associated with fatigue intensity [98].
Jang et al. found that PD patients with OH had significantly reduced levels of both 25(OH)D3 and calcitriol compared to those without OH, and these levels were negatively correlated with blood pressure changes and symptom severity [99]. However, Arici Duz et al. showed no significant differences in VitD levels among dipper, non-dipper, and reverse dipper (p = 0.192), suggesting that VitD status was not associated with autonomic dysfunction [106].
Peterson et al. demonstrated that higher serum VitD concentrations were significantly associated with better verbal fluency and memory (e.g., vegetable fluency: β = 0.264, t = 4.31, p < 0.001; HVLT immediate recall: β = 0.196, t = 3.04, p = 0.0083) and lower depression scores (GDS: β = −0.205, t = −3.08, p = 0.0083) in non-demented PD patients [97]. However, Lien et al. found no direct correlation between serum VitD levels and cognitive or motor severity in PD. Additionally, PD patients at risk of malnutrition showed worse UPDRS scores (p = 0.046) and poorer memory and calculation performance (p = 0.002 and p = 0.010) [107]. Expanding on this, Barichella et al. demonstrated that lower serum 25(OH)D3 levels (mean 17.4 ng/mL) were independently associated with more severe motor symptoms (UPDRS Part III: β = −2.10, p = 0.006), higher H&Y stage (β = −0.08, p = 0.035), and poorer global cognition (MMSE: β = 0.47, p = 0.041), despite similar VitD intake (2.5 μg/day) and sunlight exposure across groups [51].
In a longitudinal study of 60 untreated, de novo PD patients, 93.3% had VitD insufficiency at baseline, with a mean serum 25(OH)D3 level of 17.6 ± 7.3 ng/mL. After 48 months, patients who developed mild cognitive impairment (PD-MCI) had significantly lower baseline VitD levels (12.7 ± 5.4 ng/mL) compared to those who remained cognitively normal (19.7 ± 7.8 ng/mL, p = 0.005), and logistic regression confirmed low 25(OH)D3 as an independent predictor of PD-MCI (OR = 0.854, 95% CI: 0.746–0.977, p = 0.022) [108].
Resting-state fMRI revealed that patients with lower VitD levels exhibited altered spontaneous neuronal activity in default-mode and visual networks, suggesting a dose-dependent impact of VitD on brain function [104]. In this context, Russel et al. observed a progressive decline in serum 25(OH)D3 levels and regional cerebral blood flow in the basal ganglia as clinical stages advanced. Notably, patients in stages 4 and 5 exhibited significantly lower HMPAO uptake ratios and VitD concentrations compared to those in earlier stages [109].
Ogura et al. demonstrated that VitD biomarkers can effectively differentiate patient groups, with 25(OH)D3 showing strong accuracy in identifying both PD and MSA from HCs, while 1,25(OH)2D was more effective in distinguishing MSA from PD [50]. Supporting this, Guo et al. [93] found that serum levels of 25(OH)D3, along with Klotho and homocysteine, were significantly altered in MSA patients compared to healthy controls, and that these biomarkers correlated with disease severity, including motor and cognitive impairments. Although PD patients also exhibited changes in these biomarkers, the combination of Klotho, 25(OH)D3, and homocysteine provided superior diagnostic accuracy for distinguishing MSA from PD, particularly in male patients [93]. In patients with PD, CSF levels of VitD binding protein (VDBP) were significantly elevated compared to HCs, suggesting a potential involvement of VitD transport or regulation in PD pathology. This increase in VDBP was among the top contributors to distinguishing PD cases in a multianalyte biomarker profile, highlighting its relevance in disease classification [110].
Thaler et al. found that serum VitD levels were not significantly different across PD subtypes (idiopathic, GBA-related, and LRRK2-related). However, among non-manifesting LRRK2 mutation carriers, lower VitD levels were significantly associated with higher prodromal PD probability scores (β = −0.615, 95% CI: −1.179 to −0.051, p < 0.001) [94]. Tassorelli et al. reported that PD patients with a history of fall-related fractures had significantly lower VitD levels (8.5 ± 5.3 ng/mL) compared to those without fractures (16.9 ± 15.7 ng/mL, p = 0.01) [111]. Deng et al. found no statistically significant differences in serum VitD levels among the three PD subtypes. The mean concentrations were 22.8 ± 7.5 ng/mL in the severe cluster (A), 23.2 ± 7.3 ng/mL in the intermediate cluster (B), and 21.7 ± 6.7 ng/mL in the mild cluster (C), with a p-value 0.52 [112].
In patients younger than 60 years, serum VitD levels were significantly lower in the PD group (20.24 ± 12.62 ng/mL) compared to HCs (32.46 ± 19.14 ng/mL; p = 0.010). This deficiency was associated with increased disease severity and duration, suggesting that lower VitD levels may contribute to earlier and more aggressive progression of PD in younger individuals [45]. Similarly, in patients aged 60 and older, serum VitD levels were significantly lower in the PD group (18.40 ± 5.91 ng/mL) compared to controls (28.13 ± 6.26 ng/mL; p = 0.001). Among older patients, those with higher disease severity (scores above 3) had the lowest VitD levels, particularly males (11.80 ± 2.12 ng/mL). Furthermore, VitD levels declined with longer disease duration, especially in females, where those with over 10 years of PD had levels of 23.40 ± 4.67 ng/mL, showing a statistically significant trend (p = 0.049) [61].
A three-month regimen combining butyrate triglyceride, Crocus sativus L., and VitD3 led to a notable 7.7% improvement in motor function and increased stool frequency among PD patients. This formulation, acting through the gut–brain axis, demonstrated potential as a complementary therapy targeting both motor and non-motor symptoms, particularly constipation [113].

4.3. Outdoors Sunlight Exposure

Huang et al. reported that outdoor time during summer is associated with a decreased risk of PD, and this was statistically more significant in individuals with genetic mutations characterized by high penetrance [114]. Noteworthy, some aspects like physical activity, sleep patterns, and VitD may likely be confounding factors for these findings. Supporting this, Hu et al. found that spending more than 3.5 h outdoors daily was associated with a 15% reduction in PD risk (HR = 0.85, 95% CI: 0.75–0.96) and that increased outdoor time was positively correlated with serum VitD levels, which themselves were linked to a lower PD risk [115]. Complementing these findings, Kenborg et al. conducted a large population-based case–control study in Denmark involving 3819 male PD patients and 19,282 controls, revealing that maximal outdoor work was associated with a 28% lower risk of PD (adjusted OR = 0.72, 95% CI: 0.63–0.82) [116]. Extending this evidence, Kravietz et al. performed a nationwide ecological study in France using ultraviolet B (UV-B) radiation as a proxy for VitD exposure. Analyzing over 69,000 incident PD cases, they found a quadratic, age-dependent association between UV-B exposure and PD incidence: in individuals under 70, higher UV-B exposure was linked to reduced PD risk (e.g., RR = 0.85, 95% CI: 0.77–0.94 for ages 45–49), whereas in older adults, the trend reversed [117].
Zhu et al. found that individuals engaging in more than 6 h of outdoor activity per week had a 56% lower risk of PD (adjusted OR = 0.437, 95% CI: 0.241–0.795), and participants in the highest quartile of total VitD intake (>12 μg/day) had a 46% reduced risk (adjusted OR = 0.538, 95% CI: 0.301–0.960) [118]. Similarly, Kwon et al. conducted a population-based case–control study in western Washington State and observed a modest inverse association between outdoor occupational exposure and PD risk. Specifically, individuals whose entire job history involved outdoor work had a lower risk of PD compared to those with exclusively indoor work (OR = 0.74, 95% CI: 0.44–1.25), although the trend was not statistically significant [119].

4.4. In Vitro and In Vivo Studies with VitD and PD

In vitro and in vivo studies have increasingly explored the role of VitD in the pathophysiology of PD, revealing its potential neuroprotective and neuromodulatory effects (Figure 8) (Table S14–In Vitro) [7,8,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136] (Table S15–In Vivo) [8,130,132,134,137,138,139,140,141,142,143,144,145,146,147,148,149,150,151,152,153,154,155,156,157,158,159,160,161,162] (Table 2) [133,144,163,164,165,166,167,168].
Figure 8. Vitamin D in Parkinson’s disease. Abbreviations: AIF, apoptosis-inducing factor; AMPK, adenosine monophosphate-activated protein kinase; BDNF, brain-derived neurotrophic factor; Ca, calcium; Calbindin-D28k, calcium-binding protein that buffers intracellular Ca2+; DA, dopamine; GDNF, glial cell derived neurotrophic factor; GFR⍺1, GDNF family receptor alpha-1; GSH, glutathione; IL4, interleukin 4; IL10, interleukin 10; iNOS, inducible nitric oxide synthase; LC3, microtubule-associated protein 1A/1B-light chain 3; LIF, leukemia inhibitory factor; LPS, lipopolysaccharide; MAO, monoamine oxidase; MAP2, microtubule-associated protein 2; M-CSF, macrophage colony-stimulating factor; NEFH, neurofilament heavy chain; NGF, nerve growth factor; NLRP3, nod-like receptor family, pyrin domain-containing 3; NO2, nitrogen dioxide; nSMase, neutral sphingomyelinase; Nur77, nuclear receptor 77; PARP1, poly(ADP-ribose) polymerase 1; PD, Parkinson’s disease; PI3K, phosphoinositide 3-kinase; ROS, reactive oxygen species; RXRβ/γ, retinoid X receptor isoforms β and γ; SM, sphingomyelin; TGF-β, transforming growth factor-beta; TH, tyrosine hydroxylase; TLR4, Toll-like receptor 4; TNF-α, tumor necrosis factor alpha; VDR, vitamin D receptor; VEGF, vascular endothelial growth factor; VitD, vitamin D; VMAT2, vesicular monoamine transporter 2; αSyn, α-synuclein; γ-GT, gamma-glutamyltransferase; ↑, increase; ↓, decrease; ↔, no effect, no dependent [7,8,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136].
Table 2. Genes roles in dopaminergic neuron function and response to VitD.

4.4.1. In Vitro

Gul et al. investigated the neuroprotective effects of 1,25(OH)2D3 (calcitriol) and a multi-strain probiotic formulation (SLAB51) against rotenone-induced neurotoxicity in SH-SY5Y human neuroblastoma cells, a model for PD [169]. Rotenone (150 nM) significantly reduced cell viability, mimicking PD-like oxidative stress. Pretreatment and post-treatment with calcitriol (1.25–5 µM) and probiotics (0.01–0.1 mg/mL), alone or in combination, significantly improved cell viability and modulated antioxidant enzyme levels. Specifically, treatments increased SOD, GSH, and CAT activities while reducing GSR levels. Calcitriol at 2.5 µM and probiotics at 0.05–0.1 mg/mL were particularly effective. The combination therapy showed synergistic effects, with the highest cell viability increase (66.4%) observed in post-treatment groups.
Rotenone, a mitochondrial complex I inhibitor, induces excessive ROS production, leading to oxidative stress and dopaminergic neurodegeneration in Parkinson’s disease models. Antioxidant enzymes such as SOD, catalase, and glutathione peroxidase play a pivotal role in neutralizing ROS and maintaining redox homeostasis. Experimental studies in rotenone-induced PD models have shown that enhancing these enzyme activities significantly reduces oxidative markers like thiobarbituric acid reactive substances (TBARS), restores GSH levels, and improves neuronal survival and motor function [170]. Compounds such as quercetin and selenium have been reported to upregulate antioxidant enzyme expression, mitigating rotenone-induced neurotoxicity and apoptosis, highlighting their therapeutic potential in PD management [171].
VDR plays a crucial neuroprotective role in PD by reducing oxidative stress in dopaminergic neurons and suppressing inflammation in microglia via inhibition of the NLRP3/caspase-1 pathway [13]. In both cellular and Caenorhabditis elegans PD models, VDR activation—either through vitamin D3 supplementation or genetic upregulation—restores mitochondrial function, reduces αSyn toxicity, and improves motor behavior. Importantly, Zheng et al. identified DUB3 as a novel deubiquitinase that stabilizes VDR protein levels by interacting with its ligand-binding domain, thereby enhancing VDR’s protective effects and offering a promising therapeutic target for PD intervention.
Pro-inflammatory and pro-apoptotic pathways are central to PD-related neurodegeneration, as chronic neuroinflammation amplifies oxidative stress and triggers caspase-mediated neuronal apoptosis. VitD’s immunomodulatory properties are particularly relevant here, as it downregulates pro-inflammatory cytokines (e.g., TNF-α, IL-6) and inhibits NF-κB signaling, thereby reducing microglial activation and subsequent neuronal damage [172]. This mechanism underscores its potential neuroprotective role in mitigating both inflammatory and apoptotic cascades that accelerate dopaminergic neuron loss. Furthermore, previous studies have demonstrated that VitD, its receptor (VDR), and activating enzymes (such as 1α-hydroxylase and CYP27B1) are expressed in the substantia nigra [173]. Low VitD levels may therefore impair VDR-mediated signaling, contributing to dysfunction or death of dopaminergic neurons. Such deficits could exacerbate motor symptoms like bradykinesia and rigidity, as well as non-motor features including cognitive decline and mood disturbances [174].
NSC-34 motor neuron-like cells express key enzymes involved in VitD metabolism, including CYP24A1, and actively convert 25(OH)D3 into 24,25OHD3. The expression of CYP24A1 is upregulated by calcitriol and synthetic analogs, indicating a functional regulatory mechanism. These findings suggest NSC-34 cells are a promising in vitro model for investigating neuronal VitD metabolism [175]. Compared to in vivo models like 6-OHDA-lesioned rats or MPTP-treated mice, NSC-34 cells offer a controlled environment to study molecular mechanisms without systemic variables. Unlike whole-animal models, they allow direct manipulation of VitD pathways at the cellular level.
VitD significantly downregulated LAMP3 expression in human monocyte-derived dendritic cells, reducing mRNA levels by 1.62-fold at 24 h and 2.2-fold at 168 h during differentiation (p < 0.001). This suppression, mediated via NF-κB pathway inhibition, mirrors the tolerogenic phenotype induced by IL-10 and dexamethasone, suggesting a potential immunomodulatory role relevant to PD, where LAMP3 is a known genetic risk factor [176].

4.4.2. In Vivo

In a rotenone-induced PD model using Wistar rats, VitD was administered orally at 500 IU/kg/day, either as monotherapy or combined with canagliflozin (10 mg/kg/day) [177]. VitD alone improved striatal dopamine content, reduced oxidative stress markers such as malondialdehyde (MDA), and increased antioxidant defenses including GSH and catalase (CAT). It also attenuated inflammatory cytokines (TNF-α, IL-6, IL-10) and partially normalized NF-κB and Nrf2 expression. Histological analysis revealed preservation of striatal architecture and reduced neuronal damage, while immunohistochemistry showed improved tyrosine hydroxylase (TH) expression and decreased α-synuclein accumulation. Although the combination therapy produced more pronounced effects across all parameters, the study design does not allow clear attribution of these benefits to VitD alone. Therefore, only monotherapy findings should be considered for this review to avoid confounding effects from canagliflozin.
Experimental models, such as MPTP- and 6-OHDA-induced PD in rodents, have shown that VitD supplementation significantly reduces oxidative stress and inflammation, preserves dopaminergic neurons in the substantia nigra, and improves motor coordination [152,154]. These effects are mediated through the upregulation of anti-inflammatory cytokines (e.g., IL-10, TGF-β) and downregulation of pro-inflammatory markers (e.g., TNF-α, IL-1β) [178], as well as through enhanced expression of neurotrophins like BDNF [179] and GDNF [133]. Additionally, VitD modulates neurotransmitter systems, particularly the cholinergic and dopaminergic pathways, by restoring acetylcholine levels and reducing acetylcholinesterase activity [180].
Emerging evidence also suggests that VitD may regulate iron homeostasis and inhibit ferroptosis by modulating the expression of iron transport and detoxification proteins [181]. Furthermore, VitD influences synaptic plasticity and ion channel function, including L-type calcium channels [182] and NMDA receptors [183]. Epigenetically, VitD acts through VDR to modulate gene expression via histone modification and DNA methylation [184].
Calcitriol significantly improved motor performance and protected dopaminergic neurons in 6-OHDA-induced hemiparkinsonian mice [185]; it promoted Treg expansion, which suppressed T-cell infiltration and microglial M1 polarization, reducing pro-inflammatory cytokines like TNF-α, IL-1β, and IL-6; depletion of Treg using anti-CD25 antibody abolished calcitriol’s anti-inflammatory effects.
VitD significantly attenuated L-DOPA-induced abnormal involuntary movements (AIMs) in a mouse model of PD, showing a maximal reduction of 32.7% on day 14 of treatment without altering dopamine metabolic enzymes such as TH and MAO-B. Additionally, VD3 reduced oxidative stress markers like p47phox and inflammatory cytokines such as IL-1β [186].
In a haloperidol-induced mouse model of Parkinsonism, supplementation with VitD3 (800 IU) and vitamin A (1000 IU) for 21 days significantly improved motor function, as evidenced by increased bilateral climbing attempts and reduced unsuccessful climbing attempts compared to the haloperidol-only group. Additionally, the combination of VitD and A markedly reduced MDA levels, indicating decreased lipid peroxidation, and restored SOD activity to near-control levels, suggesting enhanced antioxidant defense in the striatum and primary motor cortex [187].
Vitamin D3 supplementation (800 IU/day) in dopamine-deficient mice (-D2) + VD group) partially reduced neurodegenerative changes in the visual cortex (p < 0.05) and improved retinal structure compared to the untreated -VitD2 group, although LDH levels remained elevated, indicating ongoing cytotoxic stress. Notably, when VitD was combined with vitamin A (-D2 + VD + VA group), oxidative stress markers (MDA) and cytotoxicity (LDH) were significantly lower, and histological analysis showed complete preservation of retinal architecture and absence of cortical degeneration, highlighting a synergistic protective effect [188].
In a mouse model selectively bred for high voluntary wheel-running, increased expression of the VDR was observed in both the cerebellum and striatum, suggesting a role in locomotor regulation and reward-dependent behavior. Zhang et al. showed a log2 fold change of −3.37 in cerebellum and −0.39 in striatum for VDR expression between control and high-activity lines (FDR-adjusted p = 4.4 × 10−4) [189].

4.4.3. Computational Modeling

Computational screening identified three NURR1 agonists—Lifechemical_16901310, Maybridge_2815310, and NPACT_392450—with strong binding affinities (−59.06 to −46.77 kcal/mol) and stable molecular dynamics profiles, suggesting that VitD-related pathways may be leveraged for neuroprotection. These compounds exhibited favorable ADME properties and high reactivity (HOMO-LUMO gaps of −0.14 to −0.18 eV) [190].

4.4.4. Human Studies

A 12-week BMI-adjusted VitD supplementation in PD patients with deep brain stimulation resulted in measurable biochemical changes [191]. The intervention group showed a reduction in TNF-α levels (from 7.07 ± 2.30 to 5.98 ± 2.09 pg/mL), while the placebo group experienced an increase (from 6.30 ± 2.00 to 7.23 ± 1.94 pg/mL). Additionally, 3-hydroxykynurenine (3-HK), a neurotoxic metabolite, remained stable in the VitD group but increased in the placebo group, with post-treatment levels of 6.23 ± 1.73 ng/mL versus 8.78 ± 4.93 ng/mL. Picolinic acid, known for its neuroprotective role, decreased in the placebo group but was preserved in those receiving VitD. These findings suggest that VitD may help regulate inflammation and neurotoxicity in PD.
Patients with PD showed reduced serum levels of VitD metabolites, with CYP27A1 significantly lower in PD and CYP2R1 and CYP24A1 levels remained unchanged [52]. Mazzetti et al. found that CYP27B1, the VitD–activating enzyme, was upregulated in a subpopulation of astrocytes in PD brains, particularly in regions affected by pathology, and these astrocytes were associated with αSyn oligomer clearance and neuroprotection [173]. Alaylıoğlu et al. demonstrated that although DHCR7/NADSYN1 and CYP2R1 gene variants were not associated with serum 25(OH)D3 levels or PD risk, specific SNPs in these loci correlated with milder motor symptoms, shorter disease duration, and reduced incidence of freezing of gait and falls [192].
Patients with PD and small intestinal bacterial overgrowth showed significantly lower serum triglyceride (72.8 ± 12.7 mg/dL vs. 100.2 ± 28.6 mg/dL, p = 0.024) and total bilirubin levels (0.86 ± 0.26 mg/dL vs. 0.98 ± 0.21 mg/dL, p = 0.019), suggesting impaired bile acid metabolism [193]. Hasuike et al. hypothesize that reduced bile acid from dysbiosis may hinder VitD absorption, contributing to osteoporosis and fractures commonly observed in PD.

4.5. VitD Polymorphisms and PD

VDR signaling plays a pivotal role in regulating gene expression involved in neurogenesis. Upon activation by 1,25(OH)2D3 (calcitriol), VDR forms a heterodimer with the RXR, which binds to VitD response elements in target gene promoters. This complex recruits co-activators such as DRIP and SRC, facilitating transcriptional activation and chromatin remodeling [194]. VDR signaling intersects with key developmental pathways like Wnt/β-catenin and SHH [195], which are essential for NSC proliferation and differentiation.
There are over 45 unique mutations in the VDR gene, including SNPs such as FokI (rs2228570), BsmI (rs1544410), ApaI (rs7975232), and TaqI (rs731236), which have been linked to altered susceptibility to PD. Mutations in the DNA-binding domain (e.g., Arg30) and ligand-binding domain (e.g., Arg274, Arg391) impair VDR’s ability to bind DNA, interact with co-activators, or heterodimerize with RXR, thereby disrupting downstream signaling [196]. For example, the Arg274Leu mutation reduces calcitriol binding affinity, while Glu420Lys affects co-activator recruitment without altering DNA or ligand binding [197].
Gholipour et al. showed that VDR was significantly overexpressed in PD patients, especially among females, while SNHG6 was notably underexpressed, particularly in female patients. No significant differences were found for SNHG16 and LINC00346. VDR and SNHG6 expression levels demonstrated diagnostic potential in PD with AUC values of 0.86 and 0.66, respectively [198].
Gatto et al. identified a significant association between the FokI recessive allele of the VDR gene and cognitive decline in PD, with each additional copy of the allele linked to a 0.115-point annual decrease in MMSE scores (β = −0.115, SE(β) = 0.05, p = 0.03). This effect was comparable in magnitude to the impact of disease duration and education level on cognitive performance [92].
VPA modulates the VDR pathway by altering the expression of nine key proteins involved in chromatin remodeling and transcriptional regulation, including HDAC1, SMARCA4, SMARCC1, SMARCD1, SMARCE1, BAZ1B, NCOA2, SUPT16H, and TOP2B. At a concentration of 2 mM, VPA downregulated these proteins in SH-SY5Y neuroblastoma cells, suggesting that VPA may influence VitD-mediated gene expression through epigenetic mechanisms such as histone acetylation and chromatin accessibility [199].
Genetic variants in the CACNA1C gene significantly influence PD risk, but only in individuals with VitD deficiency, suggesting a gene-environment interaction. Specifically, the G allele of SNP rs34621387 was associated with PD in VitD–deficient individuals (OR = 2.0–2.1, p = 0.002), while no association was found in non-deficient individuals (p > 0.8); additionally, VitD deficiency itself was linked to increased PD risk in both data sets (UDALL OR = 2.64, NGRC OR = 1.56) [200].
Butler et al. conducted a comprehensive genetic analysis of the VDR gene to assess its role in PD, focusing on both disease risk and age-at-onset (AAO). Using a two-stage design, they first genotyped 30 tagSNPs in 770 non-Hispanic Caucasian PD families and found that several SNPs in the 5′ region of VDR—particularly rs4334089—were significantly associated with earlier AAO (p = 0.0008). These SNPs were located in intronic regions between alternatively spliced exons, suggesting a possible mechanism involving transcript variation. In the second stage, they validated these findings using GWAS data from the NINDS cohort, where SNP rs7968585 in the 3′ region showed a significant association with AAO (p = 0.003), though not with disease risk [201].

4.6. Dietary Intake of VitD and PD

NHANES data (2007–2016) suggests that dietary intake of VitD does not show a statistically significant association with PD. Although PD patients consumed slightly less VitD (mean intake: 4.27 µg) compared to non-PD individuals (4.80 µg), the difference was modest and the adjusted OR (OR = 0.98) did not reach statistical significance (p = 0.17) [202]. Also, an MR analysis using 104 SNPs also showed no causal relationship between genetically predicted 25(OH)D3 levels and PD risk (IVW OR = 1.08; 95% CI: 0.90–1.29; p = 0.39), confirming the observational findings [62]. Yong et al. demonstrated a U-shaped association between serum 25(OH)D3 levels and all-cause mortality in PD patients, with the lowest risk observed at 78.68 nmol/L, and significantly increased mortality at both <50 nmol/L (HR 3.52, 95% CI 1.58–7.86) and >100 nmol/L (HR 2.92, 95% CI 1.06–8.05) [203].
Turcu-Stiolica et al. investigated the effects of daily supplementation with VitD3 (800 IU), folic acid (1000 µg), and vitamin B12 (15 µg) over six months in 24 levodopa-treated PD patients. The intervention significantly reduced serum homocysteine levels (from 19.98 ± 4.99 to 16.18 ± 3.73 µmol/L, p < 0.0001) and increased VitD levels (from 18.96 ± 7.28 to 22.68 ± 7.65 ng/mL, p = 0.025), while vitamin B12 levels remained statistically unchanged. Importantly, patients reported a significant improvement in health-related QoL, particularly in dimensions such as speech, mental function, discomfort and symptoms, and depression, with the overall 15D score increasing from 0.75 ± 0.11 (p = 0.02) [204].
In a cohort of 157 patients with early PD from the DATATOP study, 69.4% had VitD insufficiency (25[OH]D < 30 ng/mL) and 26.1% had deficiency (<20 ng/mL) at baseline. Over an average follow-up of 18.9 months, mean serum 25[OH]D levels increased from 26.3 ng/mL to 31.3 ng/mL, suggesting that VitD status does not decline during early disease progression [205].
In a cross-sectional study of Belgian PD patients, 55.9% had an inadequate intake of VitD, with mean daily intake at 10.1 ± 3.9 µg, below the recommended 10 µg threshold for many participants. Despite 64.4% of patients being aware of the interaction between dietary proteins and levodopa, only 18.2% consistently took their medication outside of meals, highlighting a gap in practical application of nutritional knowledge [206].
In a cross-sectional survey of 205 individuals with PD, 83.4% reported using at least one dietary supplement in the past 30 days, with VitD (74.3%), vitamin B12 (56.1%), and multivitamins (52.6%) being the most common. Notably, 43% of users took six or more supplements daily, and 29.2% had not discussed their supplement use with a healthcare provider [207].

4.7. Amount of VITD and Recommendations

The 2024 Endocrine Society guideline recommends VitD supplementation tailored to age, deficiency status, and comorbidities. For general adults up to age 74, routine supplementation is not broadly advised, though doses of 600–800 IU/day are suggested depending on age. Deficient individuals (<29 ng/mL) should receive 1000–2000 IU/day or 50,000 IU/week for 6–8 weeks, aiming for serum levels of 30–40 ng/mL. Obese and deficient individuals require higher doses (up to 100,000 IU/week). Those with diabetes or pre-diabetes should take 3500 IU/day. Adults aged 75+ may benefit from empiric supplementation (~900 IU/day), with similar deficiency protocols. Vitamin D3 is preferred, routine testing is discouraged, and calcium co-supplementation is advised for older adults (Table 3) [208].
Table 3. Vitamin D supplementation summary according to the last guideline by the Endocrine Society [208].
Older adults and those with PD should maintain sufficient VitD levels to support bone and brain health. VitD levels should be monitored through blood tests, and supplementation should be considered if levels are low, under the guidance of a healthcare provider. Safe sun exposure—5–15 min, 2–3 times per week—is also advised to naturally boost VitD. Excessive supplementation should be avoided due to risks of toxicity, and individuals taking medications should consult their providers about potential interactions [209].
While the body can produce VitD through sunlight exposure, dietary sources are essential—especially for individuals with limited sun exposure or increased needs. Fatty fish such as rainbow trout, salmon, and tuna are among the richest natural sources, offering substantial amounts per serving. Fortified foods like milk, plant-based beverages (soy, almond, rice), yogurt, and orange juice also contribute meaningfully to daily intake. Additionally, UV-exposed mushrooms provide a plant-based option with variable but potentially high VitD content (Figure 9) (Table S16–Daily VitD) [209].
Figure 9. VitD content across selected food sources. * Recommended dietary allowance (1000 IU per day).
VitD supplementation plays a crucial role in maintaining neurological health and may offer protective effects against PD. However, excessive intake can be harmful, as Chatterjee et al. reported a case of reversible parkinsonism induced by VitD toxicity, where a patient developed tremors, rigidity, and cognitive impairment after prolonged high-dose supplementation (60,000 IU/week for four years), resulting in hypercalcemia and suppressed parathyroid hormone levels [210]. Similarly, in klotho-deficient mice, elevated VitD activity triggered age-related loss of dopaminergic neurons in key brain regions, a process that was fully reversed by dietary VitD restriction [211]. This underscores the importance of balanced supplementation and medical oversight, as over-the-counter misuse of VitD can mimic or exacerbate neurodegenerative symptoms.

4.8. Conflicting Results

Kuhn et al. conducted a case–control study with 60 PD patients and matched controls, reporting mean 25(OH)D3 levels of 17.59 ± 9.28 µg/L in PD patients versus 15.69 ± 8.43 µg/L in controls, with no significant difference or correlation to disease severity or cognitive scores [212]. Fullard et al., using the PARS cohort of 198 asymptomatic individuals, found that those at high risk for PD had slightly higher vitamin D levels (27.8 ng/mL vs. 24.7 ng/mL, p = 0.09), and no association with dopamine transporter binding, suggesting that VitD insufficiency is not a preclinical marker [213]. Shrestha et al. analyzed data from 12,762 participants over 17 years and found no association between mid-life serum vitamin D levels and PD risk (HR = 1.05 for 20–30 ng/mL and HR = 1.14 for ≥30 ng/mL compared to <20 ng/mL) [214].
Luthra et al. evaluated the National Institutes of Health Exploratory Trials in Parkinson’s Disease (NET-PD) and found that only 12% of participants took more than 400 IU/day of VitD supplementation, and 34% used multivitamins estimated to contain 400 IU/day. Despite this, no significant differences in PD progression over three years were observed across groups taking VitD, multivitamins, both, or no supplements [215].
Ding et al. showed that the andrographolide derivative ZAV-12 (10–20 µM) acts as a VDR antagonist, significantly increasing LC3-II/LC3-I ratio and autophagy flux without affecting mTOR signaling, while also reducing A53T αSyn aggregation and lactate dehydrogenase release in SH-SY5Y cells, suggesting neuroprotective effects via SREBP2 activation [156].
Miyake et al. revealed that VitD intake showed no significant association with PD risk. The adjusted OR for the highest quartile of VitD intake (≥12.82 µg/day) compared to the lowest (<7.13 µg/day) was 0.82 (95% CI: 0.46–1.47; p for trend = 0.69), indicating no measurable protective or harmful effect [216].
The study by Wang et al. using bidirectional MR across three large PD cohorts does provide strong evidence that genetically predicted VitD levels are not causally associated with PD risk. By leveraging 131 high-quality SNPs related to VitD and excluding confounders like C-reactive protein and lipid levels, the authors found no significant associations using multiple MR methods, including IVW, MR-Egger, and weighted median approaches [15].
Larsson et al. suggests no causal link between VitD status and PD, as MR analysis using four SNPs associated with 25(OH)D3 showed no significant association with PD risk (OR per 10% lower 25OHD = 0.98; 95% CI: 0.93–1.04; p = 0.56). Although one SNP (rs6013897 in CYP24A1) showed a modest association with PD (OR = 1.09; p = 0.008), the overall genetic instrument did not support a causal role, indicating that previously observed lower VitD levels in PD patients may reflect reverse causation or confounding [217].
Anderson et al. showed that higher intake of VitD-rich foods was associated with an increased risk of PD, although this relationship appeared to be confounded by concurrent animal fat consumption [218]. Supporting this, Chen et al. found a significant association between dairy consumption—including dairy-derived VitD—and increased PD risk in men, but not in women, suggesting a gender-specific dietary influence [219]. A key limitation of both studies is the potential for dietary misclassification, particularly due to reliance on food frequency questionnaires that may not accurately capture portion sizes or long-term intake patterns.
Despite 60.5% of participants reporting pain, statistical analysis revealed no significant association between VitD levels and pain presence, suggesting that pain in PD may be influenced by factors other than VitD status [220].
Despite prior hypotheses and experimental evidence suggesting a neuroprotective role for VitD, several clinical studies have failed to demonstrate a consistent association between VitD levels and PD risk, severity, or symptomatology. One important consideration is the metabolic link between VitD and cholesterol. VitD synthesis in the skin relies on 7-dehydrocholesterol, a cholesterol precursor, and both pathways share biosynthetic mechanisms [221]. Consequently, many VitD-rich foods are also high in fat or derived from dairy products, which introduces a potential confounding factor. Notably, multiple epidemiological studies have reported a modest but significant association between high dairy consumption and increased PD risk, particularly in men [222].
Another limitation in many of the studies is the lack of randomization in sample selection. This introduces selection bias and may obscure true associations. Moreover, critical lifestyle factors such as physical activity and sun exposure—both of which influence VitD metabolism—were often not accounted for. Sunlight exposure, in particular, has been shown to correlate strongly with reduced PD risk, independent of VitD supplementation [223]. Taken together, these findings highlight the complexity of studying the role of VitD in neurodegeneration. Confounding dietary sources, metabolic interdependencies, and unmeasured lifestyle variables all contribute to the challenge of understanding the true effect of VitD on PD.

5. Future Studies

Despite the promising molecular and preclinical evidence supporting VitD’s role in neuroprotection and neural stem cell regulation, several limitations hinder its translation into clinical practice. A major concern is the lack of robust clinical validation; many findings are based on animal models or in vitro systems, and few have been confirmed through large-scale, randomized human trials. Additionally, the functional consequences of numerous VDR mutations remain poorly characterized, especially in the context of neurological diseases, making it difficult to predict individual responses to VitD-based therapies [224]. Supplementation studies have also yielded inconsistent results—some report cognitive improvements, while others show no significant benefit—highlighting the need for standardized dosing protocols and better stratification of patient populations. Furthermore, the long-term safety and efficacy of high-dose VitD or its analogs in neurodegenerative conditions are still uncertain, raising concerns about potential adverse effects such as hypercalcemia.
Molecular docking revealed that calcipotriol and inecalcitol exhibited strong binding affinities with p38 MAPK, engaging critical residues such as Tyr36 and Leu168, which are essential for kinase activity. Additionally, most VitD analogues, including calcitriol, were predicted to be blood–brain barrier permeable and non-neurotoxic [172]. Moreover, VitD was found to interact with 24 PD-related targets, with ESR2 emerging as a key hub. Molecular docking revealed a strong binding affinity between VitD and ESR2 (−8.6 kcal/mol), and Mendelian randomization analysis showed a significant protective effect of ESR2 activation on PD risk (OR = 0.46, 95% CI: 0.28–0.76, p = 0.002) [225].

Clinical Trials

There are nine clinical trials registered in ClinicalTrials.gov assessing the effects of VitD on various PD outcomes (Figure 10) (Table S17–Clinical Trials). Primary endpoints included motor and non-motor symptoms, cardiac autonomic dysfunction, EEG changes, balance, bone density, and immune modulation. Several trials focused on motor function using UPDRS and TUG, while others explored inflammation, cognitive performance, and gut microbiome. Notably, NCT06697626 and NCT06539260 included comprehensive secondary outcomes such as inflammatory markers and cognitive scales.
Figure 10. Clinical trials assessing the effect of vitamin D on Parkinson’s disease (PD). EEG, electroencephalogram.
VitD therapy varied across studies, ranging from 400 IU BID to 50,000 IU weekly, with durations spanning 8 to 52 weeks. Some trials combined VitD with calcium or functional drinks, while others used BMI-adjusted dosing. High-dose regimens (e.g., 50 K IU/week) were common in short-term interventions, whereas lower daily doses were used in longer trials.
An interesting point to consider is that the true effect of VitD on neurodegeneration is extremely difficult to isolate and assess. Many patients included in the studies cited above were taking multivitamins containing VitD, and even those who are not currently supplementing may still be consuming VitD through fortified foods. As a result, determining the independent impact of VitD on neurodegenerative processes is inherently challenging.
Notably, the most recent guidelines from the Endocrine Society recommend VitD supplementation only for individuals aged 75 and older, suggesting that younger populations may not benefit from routine supplementation. This further complicates the design of studies aiming to evaluate VitD’s role in neuroprotection, as such research would likely require strict dietary control or even complete replacement of standard food sources to eliminate confounding variables.
As discussed earlier, numerous animal studies have demonstrated promising effects of VitD, including enhanced survival of dopaminergic neurons and slowed progression of neurodegeneration. However, these benefits have not been consistently replicated in clinical trials. This discrepancy may be due to the multifactorial nature of VitD’s effects, which are influenced not only by supplementation but also by lifestyle factors such as physical activity and sun exposure.
Therefore, indirect measures of VitD status may not fully capture its biological impact. While some studies have attempted to use sun exposure and UV-B radiation as proxies to better understand VitD’s role in neurodegenerative diseases, these approaches are also prone to variability and may yield unreliable results.

6. Conclusions

The current comprehensive review highlights a significant association between VitD deficiency and PD, with consistent evidence of lower serum levels and increased odds of insufficiency and deficiency among PD patients. VitD supplementation shows modest benefits in motor function, particularly balance and gait. Genetic analyses reveal that the FokI polymorphism in the VDR gene is most consistently linked to PD susceptibility, while other SNPs show variable or non-significant associations. Environmental factors such as outdoor exposure and dietary intake further influence PD risk, underscoring the multifactorial nature of VitD’s role in PD. Although experimental models support VitD’s neuroprotective mechanisms, clinical translation remains limited. Future studies should focus on longitudinal designs, standardized supplementation protocols, and mechanistic exploration to clarify VitD’s diagnostic, prognostic, and therapeutic potential in PD.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/neurosci6040130/s1, Table S1: PRISMA 2020—Checklist [17]; Table S2: FreeText and MeSH search terms in the US National Library of Medicine; Table S3: Chromosome 12 SNPs Associated with the Vitamin D Receptor; Table S4: VitD assay methods [28,29,30,31,32,33,34,35]; Table S5: Quality assessment of observational studies of vitamin D levels based on Newcastle-Ottawa Scale [28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62]; Table S6: Quality assessment of clinical trial studies based on robvis [63,64,65,66,67,68,69,70]; Table S7: Quality assessment of observational studies of genetic polymorphism based on Newcastle-Ottawa Scale [35,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84]; Table S8: Serum 25(OH)D levels in PD versus control [28,29,30,31,32,33,34,35,36,37,38,40,41,42,43,44,45,46,47,48,50,51,52,53,54,55,56,57,58,60,61,62,86]; Table S9: 25(OH)D insufficiency in PD versus control [29,30,31,33,34,35,42,44,48,49,53,58]; Table S10: 25(OH)D deficiency in PD versus control [29,31,32,34,35,39,41,42,43,44,47,48,49,51,53,58,59]; Table S11: VitD and PD incidence [90,91]; Table S12: VitD supplementation versus placebo in PD [63,64,65,66,67,68,69,70]; Table S13: Comparison between alleles and genotypes of VitD SNPs and Parkinson’s disease [40,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89]; Table S14: Effects of Vitamin D in In Vitro Models of Parkinson’s Disease [7,8,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136]; Table S15: Effects of Vitamin D in In Vivo Models of Parkinson’s Disease [8,130,132,134,137,138,139,140,141,142,143,144,145,146,147,148,149,150,151,152,153,154,155,156,157,158,159,160,161,162]; Table S16: Amounts of Vitamin D and Energy per Standard Portion; Table S17: Clinical trials assessing the effect of VitD on PD.

Author Contributions

Conceptualization, J.P.R. and A.L.F.C.; methodology, A.L.F.C.; software, A.L.F.C.; validation, A.L.F.C. and J.P.R.; formal analysis, A.L.F.C.; investigation, A.L.F.C.; resources, A.L.F.C.; data curation, J.P.R.; writing—original draft preparation, J.P.R.; writing—review and editing, J.P.R.; visualization, A.L.F.C.; supervision, A.L.F.C.; project administration, A.L.F.C.; funding acquisition, J.P.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The original contributions presented in this study are included in the article/Supplementary Material. Further inquiries can be directed to the corresponding author(s).

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
AUCArea under the curve
BAZ1BBromodomain adjacent to zinc finger domain 1B
BDNFBrain-derived neurotrophic factor
BST-1Bone marrow stromal cell antigen 1
CATCatalase
CSFCerebrospinal fluid
DADopamine
DATATOPDeprenyl and tocopherol antioxidative therapy of parkinsonism
fMRIFunctional MRI
GAKCyclin G associated kinase
GDNFGlial cell-line-derived neurotrophic factor
GSHGlutathione
GSRGlutathione reductase
HCHealthy control
HDAC1Histone Deacetylase 1
HLA-DRAMajor Histocompatibility Complex, Class II, DR Alpha
HRHazard ratio
H&YHoehn and Yahr scale
IVWInverse-variance weighted
LRRK2Leucine-Rich Repeat Kinase 2
MAO-BMonoamine oxidase-B
MCIMild cognitive impairment
MDAMalondialdehyde
MMSEMini-mental state examination
MRMendelian randomization
MRIMagnetic resonance imaging
MSAMultiple system atrophy
NCOA2Nuclear receptor coactivator 2
NMDAN-methyl-D-aspartate
NMSSNon-motor symptoms scale for PD
NOSNewcastle-Ottawa scale
NSCNeural stem cell
OHOrthostatic hypotension
OROdds ratio
PARSParkinson associated risk syndrome
PDParkinson’s disease
QoLQuality of life
ROTRotenone
RRRelative risk
RXRRetinoid X receptor
SDStandard deviation
SFXN2Sideroflexin 2
SHHSonic hedgehog
SMARCA4SWI/SNF related, matrix associated, actin dependent regulator of chromatin subfamily a member 4
SMARCC1SWI/SNF Related, Matrix Associated, Actin Dependent Regulator of Chromatin Subfamily C Member 1
SMARCD1SWI/SNF Related, Matrix Associated, Actin Dependent Regulator of Chromatin Subfamily D Member 1
SMARCE1SWI/SNF Related, Matrix Associated, Actin Dependent Regulator of Chromatin Subfamily E Member 1
SMDStandard mean deviation
SODSuperoxide dismutase
STBD1Starch binding domain 1
SUPT16HSuppressor of Ty 16 homolog
THTyrosine hydroxylase
TNF-αTumor necrosis factor-alpha
TOP2BDNA topoisomerase II beta
TregRegulatory T cell
UPDRSUnified PD rating scale
UPDRS-IIUPDRS Part II
UPDRS-IIIUPDRS Part III
UV-BUltraviolet B
VDRVitamin D receptor
VitDVitamin D
VPAValproic acid
αSynα-Synuclein
1,25(OH)2D31,25-dihydroxyvitamin D3
25(OH)D325-hydroxyvitamin D3

References

  1. Pringsheim, T.; Jette, N.; Frolkis, A.; Steeves, T.D.L. The Prevalence of Parkinson’s Disease: A Systematic Review and Meta-Analysis. Mov. Disord. 2014, 29, 1583–1590. [Google Scholar] [CrossRef]
  2. Rissardo, J.P.; Caprara, A.L. A Narrative Review on Biochemical Markers and Emerging Treatments in Prodromal Synucleinopathies. Clin. Pract. 2025, 15, 65. [Google Scholar] [CrossRef]
  3. Ben-Shlomo, Y.; Darweesh, S.; Llibre-Guerra, J.; Marras, C.; San Luciano, M.; Tanner, C. The Epidemiology of Parkinson’s Disease. Lancet 2024, 403, 283–292. [Google Scholar] [CrossRef]
  4. Zhou, Z.; Zhou, R.; Zhang, Z.; Li, K. The Association Between Vitamin D Status, Vitamin D Supplementation, Sunlight Exposure, and Parkinson’s Disease: A Systematic Review and Meta-Analysis. Med. Sci. Monit. 2019, 25, 666–674. [Google Scholar] [CrossRef] [PubMed]
  5. Derex, L.; Trouillas, P. Reversible Parkinsonism, Hypophosphoremia, and Hypocalcemia under Vitamin D Therapy. Mov. Disord. 1997, 12, 612–613. [Google Scholar] [CrossRef] [PubMed]
  6. Raggi, A.; Ferri, R. The Role of Vitamins in Neurodegeneration: A Brief Review of Mechanisms, Clinical Evidence, and Therapeutic Perspectives. Psychogeriatrics 2025, 25, e70071. [Google Scholar] [CrossRef]
  7. Ibi, M.; Sawada, H.; Nakanishi, M.; Kume, T.; Katsuki, H.; Kaneko, S.; Shimohama, S.; Akaike, A. Protective Effects of 1α,25-(OH)2D3 Against the Neurotoxicity of Glutamate and Reactive Oxygen Species in Mesencephalic Culture. Neuropharmacology 2001, 40, 761–771. [Google Scholar] [CrossRef]
  8. Wang, J.Y.; Wu, J.N.; Cherng, T.L.; Hoffer, B.J.; Chen, H.H.; Borlongan, C.V.; Wang, Y. Vitamin D3 Attenuates 6-Hydroxydopamine-Induced Neurotoxicity in Rats. Brain Res. 2001, 904, 67–75. [Google Scholar] [CrossRef]
  9. Marras, C.; Canning, C.G.; Goldman, S.M. Environment, Lifestyle, and Parkinson’s Disease: Implications for Prevention in the Next Decade. Mov. Disord. 2019, 34, 801–811. [Google Scholar] [CrossRef] [PubMed]
  10. Bellou, V.; Belbasis, L.; Tzoulaki, I.; Evangelou, E.; Ioannidis, J.P.A. Environmental Risk Factors and Parkinson’s Disease: An Umbrella Review of Meta-Analyses. Park. Relat. Disord. 2016, 23, 1–9. [Google Scholar] [CrossRef]
  11. Fullard, M.E.; Duda, J.E. A Review of the Relationship Between Vitamin D and Parkinson Disease Symptoms. Front. Neurol. 2020, 11, 454. [Google Scholar] [CrossRef]
  12. Al-Kuraishy, H.M.; Al-Gareeb, A.I.; Selim, H.M.; Alexiou, A.; Papadakis, M.; Negm, W.A.; Batiha, G.E.-S. Does Vitamin D Protect or Treat Parkinson’s Disease? A Narrative Review. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2024, 397, 33–40. [Google Scholar] [CrossRef]
  13. Zheng, Z.; Chen, M.; Feng, S.; Zhao, H.; Qu, T.; Zhao, X.; Ruan, Q.; Li, L.; Guo, J. VDR and Deubiquitination Control Neuronal Oxidative Stress and Microglial Inflammation in Parkinson’s Disease. Cell Death Discov. 2024, 10, 150. [Google Scholar] [CrossRef]
  14. Li, C.; Qi, H.; Wei, S.; Wang, L.; Fan, X.; Duan, S.; Bi, S. Vitamin D Receptor Gene Polymorphisms and the Risk of Parkinson’s Disease. Neurol. Sci. 2015, 36, 247–255. [Google Scholar] [CrossRef]
  15. Wang, Z.; Xia, H.; Ding, Y.; Lu, R.; Yang, X. No Association Between Genetically Predicted Vitamin D Levels and Parkinson’s Disease. PLoS ONE 2024, 19, e0313631. [Google Scholar] [CrossRef] [PubMed]
  16. Rahnemayan, S.; Ahari, S.G.; Rikhtegar, R.; Riyahifar, S.; Sanaie, S. An Umbrella Review of Systematic Reviews with Meta-Analysis on the Role of Vitamins in Parkinson’s Disease. Acta Neurol. Belg. 2023, 123, 69–83. [Google Scholar] [CrossRef] [PubMed]
  17. Page, M.J.; McKenzie, J.E.; Bossuyt, P.M.; Boutron, I.; Hoffmann, T.C.; Mulrow, C.D.; Shamseer, L.; Tetzlaff, J.M.; Akl, E.A.; Brennan, S.E.; et al. The PRISMA 2020 Statement: An Updated Guideline for Reporting Systematic Reviews. BMJ 2021, 372, n71. [Google Scholar] [CrossRef]
  18. De Vries, E.; Schoonvelde, M.; Schumacher, G. No Longer Lost in Translation: Evidence that Google Translate Works for Comparative Bag-of-Words Text Applications. Polit. Anal. 2018, 26, 417–430. [Google Scholar] [CrossRef]
  19. Wells, G.A.; Shea, B.; O’Connell, D.; Peterson, J.; Welch, V.; Losos, M.; Tugwell, P. The Newcastle-Ottawa Scale (NOS) for Assessing the Quality of Nonrandomised Studies in Meta-Analyses 2000. Available online: https://web.archive.org/web/20210716121605id_/www3.med.unipmn.it/dispense_ebm/2009-2010/Corso%20Perfezionamento%20EBM_Faggiano/NOS_oxford.pdf (accessed on 17 September 2025).
  20. McGuinness, L.A.; Higgins, J.P.T. Risk-of-bias VISualization (robvis): An R Package and Shiny Web App for Visualizing Risk-of-Bias Assessments. Res. Synth. Methods 2021, 12, 55–61. [Google Scholar] [CrossRef]
  21. DerSimonian, R.; Laird, N. Meta-Analysis in Clinical Trials. Control. Clin. Trials 1986, 7, 177–188. [Google Scholar] [CrossRef] [PubMed]
  22. Lee, J.; Kim, K.W.; Choi, S.H.; Huh, J.; Park, S.H. Systematic Review and Meta-Analysis of Studies Evaluating Diagnostic Test Accuracy: A Practical Review for Clinical Researchers-Part II. Statistical Methods of Meta-Analysis. Korean J. Radiol. 2015, 16, 1188–1196. [Google Scholar] [CrossRef]
  23. Beheshti, A.; Chavanon, M.-L.; Christiansen, H. Emotion Dysregulation in Adults with Attention Deficit Hyperactivity Disorder: A Meta-Analysis. BMC Psychiatry 2020, 20, 120. [Google Scholar] [CrossRef]
  24. Liu, H.-M.; Zheng, J.-P.; Yang, D.; Liu, Z.-F.; Li, Z.; Hu, Z.-Z.; Li, Z.-N. Recessive/Dominant Model: Alternative Choice in Case-Control-Based Genome-Wide Association Studies. PLoS ONE 2021, 16, e0254947. [Google Scholar] [CrossRef] [PubMed]
  25. Holick, M.F.; Binkley, N.C.; Bischoff-Ferrari, H.A.; Gordon, C.M.; Hanley, D.A.; Heaney, R.P.; Murad, M.H.; Weaver, C.M. Evaluation, Treatment, and Prevention of Vitamin D Deficiency: An Endocrine Society Clinical Practice Guideline. J. Clin. Endocrinol. Metab. 2011, 96, 1911–1930. [Google Scholar] [CrossRef]
  26. McCartney, C.R.; McDonnell, M.E.; Corrigan, M.D.; Lash, R.W. Vitamin D Insufficiency and Epistemic Humility: An Endocrine Society Guideline Communication. J. Clin. Endocrinol. Metab. 2024, 109, 1948–1954. [Google Scholar] [CrossRef] [PubMed]
  27. Tesi, N.; van der Lee, S.; Hulsman, M.; Holstege, H.; Reinders, M.J.T. snpXplorer: A Web Application to Explore Human SNP-Associations and Annotate SNP-Sets. Nucleic Acids Res. 2021, 49, W603–W612. [Google Scholar] [CrossRef] [PubMed]
  28. Abou-Raya, S.; Helmii, M.; Abou-Raya, A. Bone and Mineral Metabolism in Older Adults with Parkinson’s Disease. Age Ageing 2009, 38, 675–680. [Google Scholar] [CrossRef]
  29. Evatt, M.L.; Delong, M.R.; Khazai, N.; Rosen, A.; Triche, S.; Tangpricha, V. Prevalence of Vitamin D Insufficiency in Patients with Parkinson Disease and Alzheimer Disease. Arch. Neurol. 2008, 65, 1348–1352. [Google Scholar] [CrossRef]
  30. van den Bos, F.; Speelman, A.D.; van Nimwegen, M.; van der Schouw, Y.T.; Backx, F.J.G.; Bloem, B.R.; Munneke, M.; Verhaar, H.J.J. Bone Mineral Density and Vitamin D Status in Parkinson’s Disease Patients. J. Neurol. 2013, 260, 754–760. [Google Scholar] [CrossRef]
  31. Serdaroğlu Beyazal, M.; Kırbaş, S.; Tüfekçi, A.; Devrimsel, G.; Küçükali Türkyılmaz, A. The Relationship of Vitamin D with Bone Mineral Density in Parkinson’s Disease Patients. Eur. Geriatr. Med. 2016, 7, 18–22. [Google Scholar] [CrossRef]
  32. Yoon, J.H.; Park, D.K.; Yong, S.W.; Hong, J.M. Vitamin D Deficiency and its Relationship with Endothelial Dysfunction in Patients with Early Parkinson’s Disease. J. Neural Transm. 2015, 122, 1685–1691. [Google Scholar] [CrossRef]
  33. Wang, J.; Yang, D.; Yu, Y.; Shao, G.; Wang, Q. Vitamin D and Sunlight Exposure in Newly-Diagnosed Parkinson’s Disease. Nutrients 2016, 8, 142. [Google Scholar] [CrossRef]
  34. Ding, H.; Dhima, K.; Lockhart, K.C.; Locascio, J.J.; Hoesing, A.N.; Duong, K.; Trisini-Lipsanopoulos, A.; Hayes, M.T.; Sohur, U.S.; Wills, A.-M.; et al. Unrecognized Vitamin D3 Deficiency is Common in Parkinson Disease: Harvard Biomarker Study. Neurology 2013, 81, 1531–1537. [Google Scholar] [CrossRef] [PubMed]
  35. Wang, L.; Evatt, M.L.; Maldonado, L.G.; Perry, W.R.; Ritchie, J.C.; Beecham, G.W.; Martin, E.R.; Haines, J.L.; Pericak-Vance, M.A.; Vance, J.M.; et al. Vitamin D from Different Sources is Inversely Associated with Parkinson Disease. Mov. Disord. 2015, 30, 560–566. [Google Scholar] [CrossRef]
  36. Knekt, P.; Kilkkinen, A.; Rissanen, H.; Marniemi, J.; Sääksjärvi, K.; Heliövaara, M. Serum Vitamin D and the Risk of Parkinson Disease. Arch. Neurol. 2010, 67, 808–811. [Google Scholar] [CrossRef] [PubMed]
  37. Topal, K.; Paker, N.; Bugdayci, D.; Ozer, F.; Tekdos, D. Bone Mineral Density and Vitamin D Status with Idiopathic Parkinson’s Disease; Springer: London, UK, 2010; Volume 21, pp. 141–142. [Google Scholar]
  38. Senel, K.; Alp, F.; Baykal, T.; Melikoglu, M.; Erdal, A.; Ugur, M. Preliminary Study: Is There a Role of Vitamin D in Parkinson Disease? Springer: London, UK, 2011; Volume 22, p. 167. [Google Scholar]
  39. Meamar, R.; Maracy, M.; Chitsaz, A.; Ghazvini, M.R.A.; Izadi, M.; Tanhaei, A.P. Association Between Serum Biochemical Levels, Related to Bone Metabolism and Parkinson’s Disease. J. Res. Med. Sci. 2013, 18, S39–S42. [Google Scholar]
  40. Petersen, M.S.; Bech, S.; Christiansen, D.H.; Schmedes, A.V.; Halling, J. The role of vitamin D levels and vitamin D receptor polymorphism on Parkinson’s disease in the Faroe Islands. Neurosci. Lett. 2014, 561, 74–79. [Google Scholar] [CrossRef]
  41. Liu, Y.; Zhang, B.-S. Serum 25-Hydroxyvitamin D Predicts Severity in Parkinson’s Disease Patients. Neurol. Sci. 2014, 35, 67–71. [Google Scholar] [CrossRef]
  42. Ozturk, E.A.; Gundogdu, I.; Tonuk, B.; Kocer, B.G.; Tombak, Y.; Comoglu, S.; Cakci, A. Bone Mass and Vitamin D Levels in Parkinson’s Disease: Is There Any Difference Between Genders? J. Phys. Ther. Sci. 2016, 28, 2204–2209. [Google Scholar] [CrossRef]
  43. Hatem, A.K. The State of Vitamin D in Iraqi Patients with Parkinson Disease. Al Kindy Coll. Med. J. 2019, 13, 137–141. [Google Scholar] [CrossRef]
  44. Sleeman, I.; Aspray, T.; Lawson, R.; Coleman, S.; Duncan, G.; Khoo, T.K.; Schoenmakers, I.; Rochester, L.; Burn, D.; Yarnall, A. The Role of Vitamin D in Disease Progression in Early Parkinson’s Disease. J. Park. Dis. 2017, 7, 669–675. [Google Scholar] [CrossRef]
  45. Ahangar, A.A.; Saadat, P.; Hajian, K.; Kiapasha, G. The Association Between Low Levels of Serum Vitamin D and the Duration and Severity of Parkinson’s Disease. Arch. Neurosci. 2018, 5, e61085. [Google Scholar] [CrossRef]
  46. Mollenhauer, B.; Zimmermann, J.; Sixel-Döring, F.; Focke, N.K.; Wicke, T.; Ebentheuer, J.; Schaumburg, M.; Lang, E.; Friede, T.; Trenkwalder, C. Baseline Predictors for Progression 4 Years After Parkinson’s Disease Diagnosis in the De Novo Parkinson Cohort (DeNoPa). Mov. Disord. 2019, 34, 67–77. [Google Scholar] [CrossRef]
  47. Soliman, R.H.; Oraby, M.I.; Hussein, M.; Abd El-Shafy, S.; Mostafa, S. Could Vitamin D Deficiency Have an Impact on Motor and Cognitive Function in Parkinson’s Disease? Egypt. J. Neurol. Psychiatry Neurosurg. 2019, 55, 34. [Google Scholar] [CrossRef]
  48. Zhang, H.-J.; Zhang, J.-R.; Mao, C.-J.; Li, K.; Wang, F.; Chen, J.; Liu, C.-F. Relationship Between 25-Hydroxyvitamin D, Bone Density, and Parkinson’s Disease Symptoms. Acta Neurol. Scand. 2019, 140, 274–280. [Google Scholar] [CrossRef]
  49. Fahmy, E.M.; Elawady, M.E.; Sharaf, S.; Heneidy, S.; Ismail, R.S. Vitamin D Status in Idiopathic Parkinson’s Disease: An Egyptian Study. Egypt. J. Neurol. Psychiatry Neurosurg. 2020, 56, 45. [Google Scholar] [CrossRef]
  50. Ogura, H.; Hatip-Al-Khatib, I.; Suenaga, M.; Hatip, F.B.; Mishima, T.; Fujioka, S.; Ouma, S.; Matsunaga, Y.; Tsuboi, Y. Circulatory 25(OH)D and 1,25(OH)2D as Differential Biomarkers Between Multiple System Atrophy and Parkinson’s Disease Patients. eNeurologicalSci 2021, 25, 100369. [Google Scholar] [CrossRef] [PubMed]
  51. Barichella, M.; Cereda, E.; Iorio, L.; Pinelli, G.; Ferri, V.; Cassani, E.; Bolliri, C.; Caronni, S.; Pusani, C.; Schiaffino, M.G.; et al. Clinical Correlates of Serum 25-Hydroxyvitamin D in Parkinson’s Disease. Nutr. Neurosci. 2022, 25, 1128–1136. [Google Scholar] [CrossRef]
  52. Kakimoto, A.; Ogura, H.; Suenaga, M.; Mishima, T.; Fujioka, S.; Ouma, S.; Matsunaga, Y.; Tsuboi, Y. Role of Cytochrome P450 for Vitamin D Metabolisms in Patients with Neurodegenerative Disorders. Clin. Park. Relat. Disord. 2022, 7, 100162. [Google Scholar] [CrossRef]
  53. Novotnij, D.A.; Zhukova, N.G.; Shperling, L.P.; Stolyarova, V.A.; Zhukova, I.A.; Agasheva, A.E.; Shtaimets, S.V.; Druzhinina, O.A.; Shirokikh, I.V. Vitamin D and other Indicators of Calcium-Phosphorus Metabolism as Possible Predictors of Parkinson’s Disease. S.S. Korsakov J. Neurol. Psychiatry 2022, 122, 56–64. [Google Scholar] [CrossRef]
  54. Wu, H.; Khuram Raza, H.; Li, Z.; Li, Z.; Zu, J.; Xu, C.; Yang, D.; Cui, G. Correlation Between Serum 25(OH)D and Cognitive Impairment in Parkinson’s Disease. J. Clin. Neurosci. 2022, 100, 192–195. [Google Scholar] [CrossRef]
  55. Yakşi, E.; Yaşar, M.F. Bone Mineral Density and Vitamin D Levels in Parkinson’s Disease: A Retrospective Controlled Study. Northwestern Med. J. 2022, 2, 51–58. [Google Scholar] [CrossRef]
  56. Džoljić, E.; Matutinović, M.S.; Stojković, O.; Veličković, J.; Milinković, N.; Kostić, V.; Ignjatović, S. Vitamin D Serum Levels and Vitamin D Receptor Genotype in Patients with Parkinson’s Disease. Neuroscience 2023, 533, 53–62. [Google Scholar] [CrossRef] [PubMed]
  57. Sooragonda, B.G.; Sridharan, K.; Benjamin, R.N.; Prabhakar, A.T.; Sivadasan, A.; Kapoor, N.; Cherian, K.E.; Jebasingh, F.K.; Aaron, S.; Mathew, V.; et al. Do Bone Mineral Density, Trabecular Bone Score, and Hip Structural Analysis Differ in Indian Men with Parkinson’s Disease? A Case-Control Pilot Study from a Tertiary Center in Southern India. Ann. Indian Acad. Neurol. 2023, 26, 496–501. [Google Scholar] [CrossRef]
  58. Xia, M.; Zhou, Q. Correlation Between 25-Hydroxy-Vitamin D and Parkinson’s Disease. IBRO Neurosci. Rep. 2024, 16, 162–167. [Google Scholar] [CrossRef]
  59. Khan, M.S.H.; Mouri, U.K.; Islam, M.R.; Rahman, T.; Rahman, A.; Ahmed, J.U. Relationship of Vitamin D Deficiency with Non-Motor Functions of Parkinson’s Disease. Bangladesh J. Med. 2024, 35, 82–87. [Google Scholar] [CrossRef]
  60. Milanowski, J.; Nuszkiewicz, J.; Lisewska, B.; Lisewski, P.; Szewczyk-Golec, K. Adipokines, Vitamin D, and Selected Inflammatory Biomarkers among Parkinson’s Disease Patients with and Without Dyskinesia: A Preliminary Examination. Metabolites 2024, 14, 106. [Google Scholar] [CrossRef]
  61. Rahman, A.; Ahmed, M.J.; Russel, A.H.M.; Hossain, M.A.; Huq, M.N.; Khan, S.M.D.; Rahman, M.M.; Paul, B.; Saha, P.K. Association Between Serum Vitamin D Level and Parkinson’s Disease: Case Control Study in a Tertiary Level Hospital, Bangladesh. Bangladesh J. Med. 2024, 35, 180–186. [Google Scholar] [CrossRef]
  62. Xu, Y.; Wang, E.; Zhang, Q.; Liu, J.; Luo, W. Vitamin D and Focal Brain Atrophy in PD with Non-Dementia: A VBM Study. Front. Hum. Neurosci. 2024, 18, 1474148. [Google Scholar] [CrossRef]
  63. DuBose, S. Effects of Vitamin D Supplementation on Motor Symptoms of Patients with Parkinson’s Disease. Master’s Thesis, Rollins School of Public Health, Atlanta, GA, USA, 2011. [Google Scholar]
  64. Suzuki, M.; Yoshioka, M.; Hashimoto, M.; Murakami, M.; Noya, M.; Takahashi, D.; Urashima, M. Randomized, Double-Blind, Placebo-Controlled Trial of Vitamin D Supplementation in Parkinson Disease. Am. J. Clin. Nutr. 2013, 97, 1004–1013. [Google Scholar] [CrossRef]
  65. Habibi, A.H.; Anamoradi, A.; Shahidi, G.A.; Razmeh, S.; Alizadeh, E.; Kokhedan, K.M. Treatment of Levodopainduced Dyskinesia with Vitamin D: A Randomized, Double-Blind, Placebo-Controlled Trial. Neurol. Int. 2018, 10, 7737. [Google Scholar] [CrossRef]
  66. Hiller, A.L.; Murchison, C.F.; Lobb, B.M.; O’Connor, S.; O’Connor, M.; Quinn, J.F. A Randomized, Controlled Pilot Study of the Effects of Vitamin D Supplementation on Balance in Parkinson’s Disease: Does Age Matter? PLoS ONE 2018, 13, e0203637. [Google Scholar] [CrossRef]
  67. Barichella, M.; Cereda, E.; Pinelli, G.; Iorio, L.; Caroli, D.; Masiero, I.; Ferri, V.; Cassani, E.; Bolliri, C.; Caronni, S.; et al. Muscle-Targeted Nutritional Support for Rehabilitation in Patients with Parkinsonian Syndrome. Neurology 2019, 93, e485–e496. [Google Scholar] [CrossRef]
  68. Bytowska, Z.K.; Korewo-Labelle, D.; Berezka, P.; Kowalski, K.; Przewłócka, K.; Libionka, W.; Kloc, W.; Kaczor, J.J. Effect of 12-Week BMI-Based Vitamin D3 Supplementation in Parkinson’s Disease with Deep Brain Stimulation on Physical Performance, Inflammation, and Vitamin D Metabolites. Int. J. Mol. Sci. 2023, 24, 10200. [Google Scholar] [CrossRef] [PubMed]
  69. Zali, A.; Hajyani, S.; Salari, M.; Tajabadi-Ebrahimi, M.; Mortazavian, A.M.; Pakpour, B. Co-Administration of Probiotics and Vitamin D Reduced Disease Severity and Complications in Patients with Parkinson’s Disease: A Randomized Controlled Clinical Trial. Psychopharmacology 2024, 241, 1905–1914. [Google Scholar] [CrossRef] [PubMed]
  70. Li, D.; Ma, X.; Zhang, W.; Zhong, P.; Li, M.; Liu, S. Impact of Vitamin D3 Supplementation on Motor Functionality and the Immune Response in Parkinson’s Disease Patients with Vitamin D Deficiency. Sci. Rep. 2025, 15, 25154. [Google Scholar] [CrossRef]
  71. Kim, J.-S.; Kim, Y.-I.; Song, C.; Yoon, I.; Park, J.-W.; Choi, Y.-B.; Kim, H.-T.; Lee, K.-S. Association of Vitamin D Receptor Gene Polymorphism and Parkinson’s Disease in Koreans. J. Korean Med. Sci. 2005, 20, 495–498. [Google Scholar] [CrossRef] [PubMed]
  72. Han, X.; Xue, L.; Li, Y.; Chen, B.; Xie, A. Vitamin D Receptor Gene Polymorphism and its Association with Parkinson’s Disease in Chinese Han Population. Neurosci. Lett. 2012, 525, 29–33. [Google Scholar] [CrossRef]
  73. Liu, H.; Han, X.; Zheng, X.; Li, Y.; Xie, A. Association of Vitamin D Receptor Gene Polymorphisms with Parkinson Disease. Zhonghua Yi Xue Yi Chuan Xue Za Zhi 2013, 30, 13–16. [Google Scholar] [CrossRef]
  74. Lv, Z.; Tang, B.; Sun, Q.; Yan, X.; Guo, J. Association Study Between Vitamin d Receptor Gene Polymorphisms and Patients with Parkinson Disease in Chinese Han Population. Int. J. Neurosci. 2013, 123, 60–64. [Google Scholar] [CrossRef]
  75. Török, R.; Török, N.; Szalardy, L.; Plangar, I.; Szolnoki, Z.; Somogyvari, F.; Vecsei, L.; Klivenyi, P. Association of Vitamin D Receptor Gene Polymorphisms and Parkinson’s Disease in Hungarians. Neurosci. Lett. 2013, 551, 70–74. [Google Scholar] [CrossRef]
  76. Lin, C.-H.; Chen, K.-H.; Chen, M.-L.; Lin, H.-I.; Wu, R.-M. Vitamin D Receptor Genetic Variants and Parkinson’s Disease in a Taiwanese Population. Neurobiol. Aging 2014, 35, 1212.e11–1212.e13. [Google Scholar] [CrossRef]
  77. Gatto, N.M.; Sinsheimer, J.S.; Cockburn, M.; Escobedo, L.A.; Bordelon, Y.; Ritz, B. Vitamin D Receptor Gene Polymorphisms and Parkinson’s Disease in a Population with High Ultraviolet Radiation Exposure. J. Neurol. Sci. 2015, 352, 88–93. [Google Scholar] [CrossRef]
  78. Fazeli, A.; Motallebi, M.; Jamshidi, J.; Movafagh, A.; Ghaedi, H.; Emamalizadeh, B.; Kashani, K.; Darvish, H. Vitamin D Receptor Gene rs4334089 Polymorphism and Parkinson’s Disease in Iranian Population. Basal Ganglia 2016, 6, 157–160. [Google Scholar] [CrossRef]
  79. Kang, S.Y.; Park, S.; Oh, E.; Park, J.; Youn, J.; Kim, J.S.; Kim, J.-U.; Jang, W. Vitamin D Receptor Polymorphisms and Parkinson’s Disease in a Korean Population: Revisited. Neurosci. Lett. 2016, 628, 230–235. [Google Scholar] [CrossRef] [PubMed]
  80. Meamar, R.; Javadirad, S.M.; Chitsaz, N.; Asadian Ghahfarokhi, M.; Kazemi, M.; Ostadsharif, M. Vitamin D Receptor Gene Variants in Parkinson’s Disease Patients. Egypt. J. Med. Human Genet. 2017, 18, 225–230. [Google Scholar] [CrossRef]
  81. Mohammadzadeh, R.; Pazhouhesh, R. Association of VDR FokI and ApaI Genetic Polymorphisms with Parkinson’s Disease Risk in South Western Iranian Population. Acta Medica Int. 2016, 3, 111–115. [Google Scholar] [CrossRef]
  82. Gezen-Ak, D.; Alaylıoğlu, M.; Genç, G.; Gündüz, A.; Candaş, E.; Bilgiç, B.; Atasoy, İ.L.; Apaydın, H.; Kızıltan, G.; Gürvit, H.; et al. GC and VDR SNPs and Vitamin D Levels in Parkinson’s Disease: The Relevance to Clinical Features. Neuromolecular Med. 2017, 19, 24–40. [Google Scholar] [CrossRef]
  83. Tanaka, K.; Miyake, Y.; Fukushima, W.; Kiyohara, C.; Sasaki, S.; Tsuboi, Y.; Oeda, T.; Shimada, H.; Kawamura, N.; Sakae, N.; et al. Vitamin D Receptor Gene Polymorphisms, Smoking, and Risk of Sporadic Parkinson’s Disease in Japan. Neurosci. Lett. 2017, 643, 97–102. [Google Scholar] [CrossRef] [PubMed]
  84. Hu, W.; Wang, L.; Chen, B.; Wang, X. Vitamin D Receptor rs2228570 Polymorphism and Parkinson’s Disease Risk in a Chinese Population. Neurosci. Lett. 2020, 717, 134722. [Google Scholar] [CrossRef]
  85. Agliardi, C.; Guerini, F.R.; Zanzottera, M.; Bolognesi, E.; Meloni, M.; Riboldazzi, G.; Zangaglia, R.; Sturchio, A.; Casali, C.; Di Lorenzo, C.; et al. The VDR FokI (rs2228570) Polymorphism is Involved in Parkinson’s Disease. J. Neurol. Sci. 2021, 428, 117606. [Google Scholar] [CrossRef]
  86. Fahmy, E.M.; Elawady, M.E.; Sharaf, S.; Heneidy, S.; Ismail, R.S. Vitamin D Receptor Gene Polymorphisms and Idiopathic Parkinson Disease: An Egyptian Study. Egypt. J. Neurol. Psychiatry Neurosurg. 2021, 57, 102. [Google Scholar] [CrossRef]
  87. Redenšek, S.; Kristanc, T.; Blagus, T.; Trošt, M.; Dolžan, V. Genetic Variability of the Vitamin D Receptor Affects Susceptibility to Parkinson’s Disease and Dopaminergic Treatment Adverse Events. Front. Aging Neurosci. 2022, 14, 853277. [Google Scholar] [CrossRef]
  88. Canales-Cortés, S.; Rodríguez-Arribas, M.; Galindo, M.F.; Jordan, J.; Casado-Naranjo, I.; Fuentes, J.M.; Yakhine-Diop, S.M.S. Vitamin D Receptor Polymorphisms in a Spanish Cohort of Parkinson’s Disease Patients. Genet. Test. Mol. Biomark. 2024, 28, 59–64. [Google Scholar] [CrossRef]
  89. Kundu, N.C.; Kundu, A.; Khalil, M.I.; Joy, K.M.N.I.; Sen, M.; Hasan, Z.; Sahabuddin, M.; Rafi, M.A.; Hasan, M.J. A Case-Control Study on Vitamin D Receptor Gene Polymorphisms in Patients with Parkinson’s Disease in Bangladesh. Sci. Rep. 2025, 15, 12333. [Google Scholar] [CrossRef]
  90. Gröninger, M.; Sabin, J.; Kaaks, R.; Amiano, P.; Aune, D.; Castro, N.C.; Guevara, M.; Hansen, J.; Homann, J.; Masala, G.; et al. Associations of Milk, Dairy Products, Calcium and Vitamin D Intake with Risk of Developing Parkinson’s Disease Within the EPIC4ND Cohort. Eur. J. Epidemiol. 2024, 39, 1251–1265. [Google Scholar] [CrossRef]
  91. Veronese, N.; Nova, A.; Fazia, T.; Riggi, E.; Yang, L.; Piccio, L.; Huang, B.-H.; Ahmadi, M.; Barbagallo, M.; Notarnicola, M.; et al. Contribution of Nutritional, Lifestyle, and Metabolic Risk Factors to Parkinson’s Disease. Mov. Disord. 2024, 39, 1203–1212. [Google Scholar] [CrossRef]
  92. Gatto, N.M.; Paul, K.C.; Sinsheimer, J.S.; Bronstein, J.M.; Bordelon, Y.; Rausch, R.; Ritz, B. Vitamin D Receptor Gene Polymorphisms and Cognitive Decline in Parkinson’s Disease. J. Neurol. Sci. 2016, 370, 100–106. [Google Scholar] [CrossRef] [PubMed]
  93. Guo, Y.; Zhuang, X.-D.; Xian, W.-B.; Wu, L.-L.; Huang, Z.-N.; Hu, X.; Zhang, X.-S.; Chen, L.; Liao, X.-X. Serum Klotho, Vitamin D, and Homocysteine in Combination Predict the Outcomes of Chinese Patients with Multiple System Atrophy. CNS Neurosci. Ther. 2017, 23, 657–666. [Google Scholar] [CrossRef] [PubMed]
  94. Thaler, A.; Omer, N.; Giladi, N.; Gurevich, T.; Bar-Shira, A.; Gana-Weisz, M.; Goldstein, O.; Kestenbaum, M.; Cedarbaum, J.M.; Orr-Urtreger, A.; et al. Biochemical Markers for Severity and Risk in GBA and LRRK2 Parkinson’s Disease. J. Neurol. 2021, 268, 1517–1525. [Google Scholar] [CrossRef] [PubMed]
  95. Moghaddasi, M.; Mamarabadi, M.; Aghaii, M. Serum 25-Hydroxyvitamin D3 Concentration in Iranian Patients with Parkinson’s Disease. Iran. J. Neurol. 2013, 12, 56–59. [Google Scholar]
  96. Peterson, A.L.; Mancini, M.; Horak, F.B. The Relationship Between Balance Control and Vitamin D in Parkinson’s Disease-a Pilot Study. Mov. Disord. 2013, 28, 1133–1137. [Google Scholar] [CrossRef]
  97. Peterson, A.L.; Murchison, C.; Zabetian, C.; Leverenz, J.B.; Watson, G.S.; Montine, T.; Carney, N.; Bowman, G.L.; Edwards, K.; Quinn, J.F. Memory, Mood, and Vitamin D in Persons with Parkinson’s Disease. J. Park. Dis. 2013, 3, 547–555. [Google Scholar] [CrossRef] [PubMed]
  98. Canlı, M.; Yetiş, A.; Kuzu, Ş.; Valamur, İ.; Duran, S.; Sahin, B.E.; Kocaman, H.; Yıldız, N.T.; Alkan, H. Investigation of the Relationship Between Fatigue with Vitamin D, Disease Stage, Anxiety and Physical Activity Level in Patients with Parkinson’s Disease. J. Health Sci. Med. 2024, 7, 296–300. [Google Scholar] [CrossRef]
  99. Jang, W.; Park, J.; Kim, J.S.; Youn, J.; Oh, E.; Kwon, K.Y.; Jo, K.D.; Lee, M.K.; Kim, H.-T. Vitamin D Deficiency in Parkinson’s Disease Patients with Orthostatic Hypotension. Acta Neurol. Scand. 2015, 132, 242–250. [Google Scholar] [CrossRef] [PubMed]
  100. Kwon, K.Y.; Jo, K.D.; Lee, M.K.; Oh, M.; Kim, E.N.; Park, J.; Kim, J.S.; Youn, J.; Oh, E.; Kim, H.-T.; et al. Low Serum Vitamin D Levels May Contribute to Gastric Dysmotility in de novo Parkinson’s Disease. Neurodegener. Dis. 2016, 16, 199–205. [Google Scholar] [CrossRef]
  101. Kim, J.E.; Oh, E.; Park, J.; Youn, J.; Kim, J.S.; Jang, W. Serum 25-Hydroxyvitamin D3 Level May be Associated with Olfactory Dysfunction in de Novo Parkinson’s Disease. J. Clin. Neurosci. 2018, 57, 131–135. [Google Scholar] [CrossRef]
  102. Meamar, R.; Shaabani, P.; Tabibian, S.R.; Aghaye Ghazvini, M.R.; Feizi, A. The Effects of Uric Acid, Serum Vitamin D3, and Their Interaction on Parkinson’s Disease Severity. Park. Dis. 2015, 2015, 463483. [Google Scholar] [CrossRef] [PubMed]
  103. Lawton, M.; Baig, F.; Toulson, G.; Morovat, A.; Evetts, S.G.; Ben-Shlomo, Y.; Hu, M.T. Blood Biomarkers with Parkinson’s Disease Clusters and Prognosis: The Oxford Discovery Cohort. Mov. Disord. 2020, 35, 279–287. [Google Scholar] [CrossRef]
  104. Lv, L.; Zhang, H.; Tan, X.; Qin, L.; Peng, X.; Bai, R.; Xiao, Q.; Tan, C.; Liao, H.; Yan, W.; et al. Assessing the Effects of Vitamin D on Neural Network Function in Patients with Parkinson’s Disease by Measuring the Fraction Amplitude of Low-Frequency Fluctuation. Front. Aging Neurosci. 2021, 13, 763947. [Google Scholar] [CrossRef]
  105. Chitsaz, A.; Maracy, M.; Basiri, K.; Izadi Boroujeni, M.; Tanhaei, A.P.; Rahimi, M.; Meamar, R. 25-Hydroxyvitamin d and Severity of Parkinson’s Disease. Int. J. Endocrinol. 2013, 2013, 689149. [Google Scholar] [CrossRef] [PubMed]
  106. Arici Duz, O.; Helvaci Yilmaz, N. Nocturnal Blood Pressure Changes in Parkinson’s Disease: Correlation with Autonomic Dysfunction and Vitamin D Levels. Acta Neurol. Belg. 2020, 120, 915–920. [Google Scholar] [CrossRef]
  107. Lien, C.-Y.; Lu, C.-H.; Chang, C.-C.; Chang, W.-N. Correlation Between Hypovitaminosis D and Nutritional Status with The Severity of Clinical Symptoms and Impaired Cognitive Function in Patients with Parkinson’s Disease. Acta Neurol. Taiwan. 2021, 30, 63–73. [Google Scholar]
  108. Santangelo, G.; Raimo, S.; Erro, R.; Picillo, M.; Amboni, M.; Pellecchia, M.T.; Pivonello, C.; Barone, P.; Vitale, C. Vitamin D as a Possible Biomarker of Mild Cognitive Impairment in Parkinsonians. Aging Ment. Health 2021, 25, 1998–2002. [Google Scholar] [CrossRef]
  109. Russel, A.H.M.; Ali, Z.; Saha, P.K.; Rahman, A.; Haque, J.A.; Akhter, A.; Ahmed, M.J.; Rahman, M.M.; Khan, S.M.D.; Paul, B.; et al. Association of 99MTC-Hmpao Spect Measured Regional Cerebral Blood Flow and Serum Vitamin D Level with Clinical Staging of Parkinson’s Disease. Bangladesh J. Med. 2023, 34, 112–120. [Google Scholar] [CrossRef]
  110. Zhang, J.; Sokal, I.; Peskind, E.R.; Quinn, J.F.; Jankovic, J.; Kenney, C.; Chung, K.A.; Millard, S.P.; Nutt, J.G.; Montine, T.J. CSF Multianalyte Profile Distinguishes Alzheimer and Parkinson Diseases. Am. J. Clin. Pathol. 2008, 129, 526–529. [Google Scholar] [CrossRef]
  111. Tassorelli, C.; Berlangieri, M.; Buscone, S.; Bolla, M.; De Icco, R.; Baricich, A.; Pacchetti, C.; Cisari, C.; Sandrini, G. Falls, Fractures and Bone Density in Parkinson’s Disease—A Cross-Sectional Study. Int. J. Neurosci. 2017, 127, 299–304. [Google Scholar] [CrossRef]
  112. Deng, X.; Saffari, S.E.; Liu, N.; Xiao, B.; Allen, J.C.; Ng, S.Y.E.; Chia, N.; Tan, Y.J.; Choi, X.; Heng, D.L.; et al. Biomarker Characterization of Clinical Subtypes of Parkinson Disease. NPJ Park. Dis. 2022, 8, 109. [Google Scholar] [CrossRef]
  113. Alexoudi, A.; Kesidou, L.; Gatzonis, S.; Charalampopoulos, C.; Tsoga, A. Effectiveness of the Combination of Probiotic Supplementation on Motor Symptoms and Constipation in Parkinson’s Disease. Cureus 2023, 15, e49320. [Google Scholar] [CrossRef]
  114. Huang, Y.; Tian, S.; Qiu, K.; Xie, J.; Pan, A.; Liu, G.; Liao, Y. Outdoor Light Spending Time, Genetic Predisposition and Incident Parkinson’s Disease: The Mediating Effect of Lifestyle and Vitamin D. J. Health Popul. Nutr. 2025, 44, 235. [Google Scholar] [CrossRef]
  115. Hu, L.; Shi, Y.; Zou, X.; Lai, Z.; Lin, F.; Cai, G.; Liu, X. Association of Time Spent Outdoors with the Risk of Parkinson’s Disease: A Prospective Cohort Study of 329,359 Participants. BMC Neurol. 2024, 24, 10. [Google Scholar] [CrossRef]
  116. Kenborg, L.; Lassen, C.F.; Ritz, B.; Schernhammer, E.S.; Hansen, J.; Gatto, N.M.; Olsen, J.H. Outdoor Work and Risk for Parkinson’s Disease: A Population-Based Case-Control Study. Occup. Environ. Med. 2011, 68, 273–278. [Google Scholar] [CrossRef] [PubMed]
  117. Kravietz, A.; Kab, S.; Wald, L.; Dugravot, A.; Singh-Manoux, A.; Moisan, F.; Elbaz, A. Association of UV Radiation with Parkinson Disease Incidence: A Nationwide French Ecologic Study. Environ. Res. 2017, 154, 50–56. [Google Scholar] [CrossRef]
  118. Zhu, D.; Liu, G.; Lv, Z.; Wen, S.; Bi, S.; Wang, W. Inverse Associations of Outdoor Activity and Vitamin D Intake with the Risk of Parkinson’s Disease. J. Zhejiang Univ. Sci. B 2014, 15, 923–927. [Google Scholar] [CrossRef] [PubMed]
  119. Kwon, E.; Gallagher, L.G.; Searles Nielsen, S.; Franklin, G.M.; Littell, C.T.; Longstreth, W.T.J.; Swanson, P.D.; Checkoway, H. Parkinson’s Disease and History of Outdoor Occupation. Park. Relat. Disord. 2013, 19, 1164–1166. [Google Scholar] [CrossRef] [PubMed]
  120. Furman, I.; Baudet, C.; Brachet, P. Differential Expression of M-CSF, LIF, and TNF-Alpha Genes in Normal and Malignant Rat Glial Cells: Regulation by Lipopolysaccharide and Vitamin D. J. Neurosci. Res. 1996, 46, 360–366. [Google Scholar] [CrossRef]
  121. Puchacz, E.; Stumpf, W.E.; Stachowiak, E.K.; Stachowiak, M.K. Vitamin D Increases Expression of the Tyrosine Hydroxylase Gene in Adrenal Medullary Cells. Mol. Brain Res. 1996, 36, 193–196. [Google Scholar] [CrossRef]
  122. Musiol, I.M.; Feldman, D. 1,25-Dihydroxyvitamin D3 Induction of Nerve Growth Factor in L929 Mouse Fibroblasts: Effect of Vitamin D Receptor Regulation and Potency of Vitamin D3 Analogs. Endocrinology 1997, 138, 12–18. [Google Scholar] [CrossRef]
  123. Garcion, E.; Sindji, L.; Leblondel, G.; Brachet, P.; Darcy, F. 1,25-Dihydroxyvitamin D3 Regulates the Synthesis of Gamma-Glutamyl Transpeptidase and Glutathione Levels in Rat Primary Astrocytes. J. Neurochem. 1999, 73, 859–866. [Google Scholar] [CrossRef]
  124. Shinpo, K.; Kikuchi, S.; Sasaki, H.; Moriwaka, F.; Tashiro, K. Effect of 1,25-Dihydroxyvitamin D3 on Cultured Mesencephalic Dopaminergic Neurons to the Combined Toxicity Caused by L-Buthionine Sulfoximine and 1-Methyl-4-Phenylpyridine. J. Neurosci. Res. 2000, 62, 374–382. [Google Scholar] [CrossRef]
  125. Cui, X.; McGrath, J.J.; Burne, T.H.J.; Mackay-Sim, A.; Eyles, D.W. Maternal Vitamin D Depletion Alters Neurogenesis in the Developing Rat Brain. Int. J. Dev. Neurosci. 2007, 25, 227–232. [Google Scholar] [CrossRef]
  126. Orme, R.P.; Bhangal, M.S.; Fricker, R.A. Calcitriol Imparts Neuroprotection in Vitro to Midbrain Dopaminergic Neurons by Upregulating GDNF Expression. PLoS ONE 2013, 8, e62040. [Google Scholar] [CrossRef]
  127. Jang, W.; Kim, H.J.; Li, H.; Jo, K.D.; Lee, M.K.; Song, S.H.; Yang, H.O. 1,25-Dyhydroxyvitamin D3 Attenuates Rotenone-Induced Neurotoxicity in SH-SY5Y Cells Through Induction of Autophagy. Biochem. Biophys. Res. Commun. 2014, 451, 142–147. [Google Scholar] [CrossRef]
  128. Jang, W.; Park, H.-H.; Lee, K.-Y.; Lee, Y.J.; Kim, H.-T.; Koh, S.-H. 1,25-Dyhydroxyvitamin D3 Attenuates L-DOPA-Induced Neurotoxicity in Neural Stem Cells. Mol. Neurobiol. 2015, 51, 558–570. [Google Scholar] [CrossRef]
  129. Pertile, R.A.N.; Cui, X.; Eyles, D.W. Vitamin D signaling and the Differentiation of Developing Dopamine Systems. Neuroscience 2016, 333, 193–203. [Google Scholar] [CrossRef]
  130. Cataldi, S.; Arcuri, C.; Hunot, S.; Légeron, F.-P.; Mecca, C.; Garcia-Gil, M.; Lazzarini, A.; Codini, M.; Beccari, T.; Tasegian, A.; et al. Neutral Sphingomyelinase Behaviour in Hippocampus Neuroinflammation of MPTP-Induced Mouse Model of Parkinson’s Disease and in Embryonic Hippocampal Cells. Mediat. Inflamm. 2017, 2017, 2470950. [Google Scholar] [CrossRef] [PubMed]
  131. Rcom-H’cheo-Gauthier, A.N.; Meedeniya, A.C.B.; Pountney, D.L. Calcipotriol Inhibits α-Synuclein Aggregation in SH-SY5Y Neuroblastoma Cells by a Calbindin-D28k-Dependent Mechanism. J. Neurochem. 2017, 141, 263–274. [Google Scholar] [CrossRef] [PubMed]
  132. Cataldi, S.; Arcuri, C.; Hunot, S.; Mecca, C.; Codini, M.; Laurenti, M.E.; Ferri, I.; Loreti, E.; Garcia-Gil, M.; Traina, G.; et al. Effect of Vitamin D in HN9.10e Embryonic Hippocampal Cells and in Hippocampus from MPTP-Induced Parkinson’s Disease Mouse Model. Front. Cell. Neurosci. 2018, 12, 31. [Google Scholar] [CrossRef]
  133. Pertile, R.A.N.; Cui, X.; Hammond, L.; Eyles, D.W. Vitamin D Regulation of GDNF/Ret Signaling in Dopaminergic Neurons. FASEB J. 2018, 32, 819–828. [Google Scholar] [CrossRef] [PubMed]
  134. Hu, J.; Wu, J.; Wan, F.; Kou, L.; Yin, S.; Sun, Y.; Li, Y.; Zhou, Q.; Wang, T. Calcitriol Alleviates MPP+- and MPTP-Induced Parthanatos Through the VDR/PARP1 Pathway in the Model of Parkinson’s Disease. Front. Aging Neurosci. 2021, 13, 657095. [Google Scholar] [CrossRef]
  135. Zhang, Y.; Ji, W.; Zhang, S.; Gao, N.; Xu, T.; Wang, X.; Zhang, M. Vitamin D Inhibits the Early Aggregation of α-Synuclein and Modulates Exocytosis Revealed by Electrochemical Measurements. Angew. Chem. Int. Ed. 2022, 61, e202111853. [Google Scholar] [CrossRef]
  136. de Siqueira, E.A.; Magalhães, E.P.; de Assis, A.L.C.; Sampaio, T.L.; Lima, D.B.; Marinho, M.M.; Martins, A.M.C.; de Andrade, G.M.; de Barros Viana, G.S. 1α,25-Dihydroxyvitamin D3 (VD3) Shows a Neuroprotective Action Against Rotenone Toxicity on PC12 Cells: An In Vitro Model of Parkinson’s Disease. Neurochem. Res. 2023, 48, 250–262. [Google Scholar] [CrossRef]
  137. Sanchez, B.; Lopez-Martin, E.; Segura, C.; Labandeira-Garcia, J.L.; Perez-Fernandez, R. 1,25-Dihydroxyvitamin D3 Increases Striatal GDNF mRNA and Protein Expression in Adult Rats. Mol. Brain Res. 2002, 108, 143–146. [Google Scholar] [CrossRef]
  138. Lin, A.M.Y.; Fan, S.F.; Yang, D.M.; Hsu, L.L.; Yang, C.H.J. Zinc-Induced Apoptosis in Substantia Nigra of Rat Brain: Neuroprotection by Vitamin D3. Free Radic. Biol. Med. 2003, 34, 1416–1425. [Google Scholar] [CrossRef]
  139. Kalueff, A.V.; Lou, Y.-R.; Laaksi, I.; Tuohimaa, P. Impaired Motor Performance in Mice Lacking Neurosteroid Vitamin D Receptors. Brain Res. Bull. 2004, 64, 25–29. [Google Scholar] [CrossRef]
  140. Burne, T.H.J.; McGrath, J.J.; Eyles, D.W.; Mackay-Sim, A. Behavioural Characterization of Vitamin D Receptor Knockout Mice. Behav. Brain Res. 2005, 157, 299–308. [Google Scholar] [CrossRef]
  141. Kim, J.-S.; Ryu, S.-Y.; Yun, I.; Kim, W.-J.; Lee, K.-S.; Park, J.-W.; Kim, Y.-I. 1alpha,25-Dihydroxyvitamin D3 Protects Dopaminergic Neurons in Rodent Models of Parkinson’s Disease through Inhibition of Microglial Activation. J. Clin. Neurol. 2006, 2, 252–257. [Google Scholar] [CrossRef]
  142. Smith, M.P.; Fletcher-Turner, A.; Yurek, D.M.; Cass, W.A. Calcitriol Protection Against Dopamine Loss Induced by Intracerebroventricular Administration of 6-Hydroxydopamine. Neurochem. Res. 2006, 31, 533–539. [Google Scholar] [CrossRef] [PubMed]
  143. Sanchez, B.; Relova, J.L.; Gallego, R.; Ben-Batalla, I.; Perez-Fernandez, R. 1,25-Dihydroxyvitamin D3 Administration to 6-Hydroxydopamine-Lesioned Rats Increases Glial Cell Line-Derived Neurotrophic Factor and Partially Restores Tyrosine Hydroxylase Expression in Substantia Nigra and Striatum. J. Neurosci. Res. 2009, 87, 723–732. [Google Scholar] [CrossRef] [PubMed]
  144. Cui, X.; Pelekanos, M.; Burne, T.H.J.; McGrath, J.J.; Eyles, D.W. Maternal Vitamin D Deficiency Alters the Expression of Genes Involved in Dopamine Specification in the Developing Rat Mesencephalon. Neurosci. Lett. 2010, 486, 220–223. [Google Scholar] [CrossRef] [PubMed]
  145. Cass, W.A.; Peters, L.E.; Fletcher, A.M.; Yurek, D.M. Evoked Dopamine Overflow is Augmented in the Striatum of Calcitriol Treated Rats. Neurochem. Int. 2012, 60, 186–191. [Google Scholar] [CrossRef] [PubMed]
  146. Dean, E.D.; Mexas, L.M.; Cápiro, N.L.; McKeon, J.E.; DeLong, M.R.; Pennell, K.D.; Doorn, J.A.; Tangpricha, V.; Miller, G.W.; Evatt, M.L. 25-Hydroxyvitamin D Depletion Does Not Exacerbate MPTP-Induced Dopamine Neuron Damage in Mice. PLoS ONE 2012, 7, e39227. [Google Scholar] [CrossRef]
  147. Hashemvarzi, S.A.; Samadi, B.; Khazaeli, N. Preconditioning Effect of Aerobic Exercise with Vitamin D3 Intake on VEGF Levels in 6-OHDA-Lesioned Rat Model of Parkinson’s Disease. J. Basic. Clin. Pathophysiol. 2017, 5, 1–8. [Google Scholar]
  148. Jiang, P.; Zhang, W.-Y.; Li, H.-D.; Cai, H.-L.; Xue, Y. Repeated Haloperidol Administration Has No Effect on Vitamin D Signaling but Increase Retinoid X Receptors and Nur77 Expression in Rat Prefrontal Cortex. Cell. Mol. Neurobiol. 2013, 33, 309–312. [Google Scholar] [CrossRef]
  149. Cass, W.A.; Peters, L.E.; Fletcher, A.M.; Yurek, D.M. Calcitriol Promotes Augmented Dopamine Release in the Lesioned Striatum of 6-Hydroxydopamine Treated Rats. Neurochem. Res. 2014, 39, 1467–1476. [Google Scholar] [CrossRef]
  150. Li, H.; Jang, W.; Kim, H.J.; Jo, K.D.; Lee, M.K.; Song, S.H.; Yang, H.O. Biochemical Protective Effect of 1,25-Dihydroxyvitamin D3 Through Autophagy Induction in the MPTP Mouse Model of Parkinson’s Disease. Neuroreport 2015, 26, 669–674. [Google Scholar] [CrossRef] [PubMed]
  151. Mark, K.A.; Dumas, K.J.; Bhaumik, D.; Schilling, B.; Davis, S.; Oron, T.R.; Sorensen, D.J.; Lucanic, M.; Brem, R.B.; Melov, S.; et al. Vitamin D Promotes Protein Homeostasis and Longevity via the Stress Response Pathway Genes skn-1, ire-1, and xbp-1. Cell Rep. 2016, 17, 1227–1237. [Google Scholar] [CrossRef]
  152. Calvello, R.; Cianciulli, A.; Nicolardi, G.; De Nuccio, F.; Giannotti, L.; Salvatore, R.; Porro, C.; Trotta, T.; Panaro, M.A.; Lofrumento, D.D. Vitamin D Treatment Attenuates Neuroinflammation and Dopaminergic Neurodegeneration in an Animal Model of Parkinson’s Disease, Shifting M1 to M2 Microglia Responses. J. Neuroimmune Pharmacol. 2017, 12, 327–339. [Google Scholar] [CrossRef]
  153. Cass, W.A.; Peters, L.E. Reduced Ability of Calcitriol to Promote Augmented Dopamine Release in the Lesioned Striatum of Aged Rats. Neurochem. Int. 2017, 108, 222–229. [Google Scholar] [CrossRef]
  154. Lima, L.A.R.; Lopes, M.J.P.; Costa, R.O.; Lima, F.A.V.; Neves, K.R.T.; Calou, I.B.F.; Andrade, G.M.; Viana, G.S.B. Vitamin D Protects Dopaminergic Neurons Against Neuroinflammation and Oxidative Stress in Hemiparkinsonian Rats. J. Neuroinflammation 2018, 15, 249. [Google Scholar] [CrossRef] [PubMed]
  155. Kim, H.; Shin, J.-Y.; Lee, Y.-S.; Yun, S.P.; Maeng, H.-J.; Lee, Y. Brain Endothelial P-Glycoprotein Level Is Reduced in Parkinson’s Disease via a Vitamin D Receptor-Dependent Pathway. Int. J. Mol. Sci. 2020, 21, 8538. [Google Scholar] [CrossRef]
  156. Ding, M.-Y.; Peng, Y.; Li, F.; Li, Z.-Q.; Wang, D.; Zhou, G.-C.; Wang, Y. Andrographolide Derivative as Antagonist of Vitamin D Receptor to Induce Lipidation of Microtubule Associate Protein 1 Light Chain 3 (LC3). Bioorg. Med. Chem. 2021, 51, 116505. [Google Scholar] [CrossRef]
  157. Araújo de Lima, L.; Oliveira Cunha, P.L.; Felicio Calou, I.B.; Tavares Neves, K.R.; Facundo, H.T.; Socorro de Barros Viana, G. Effects of Vitamin D (VD3) Supplementation on the Brain Mitochondrial Function of Male Rats, in the 6-OHDA-Induced Model of Parkinson’s Disease. Neurochem. Int. 2022, 154, 105280. [Google Scholar] [CrossRef]
  158. Bayo-Olugbami, A.; Nafiu, A.B.; Amin, A.; Ogundele, O.M.; Lee, C.C.; Owoyele, B.V. Vitamin D Attenuated 6-OHDA-Induced Behavioural Deficits, Dopamine Dysmetabolism, Oxidative Stress, and Neuro-Inflammation in Mice. Nutr. Neurosci. 2022, 25, 823–834. [Google Scholar] [CrossRef] [PubMed]
  159. Magdy, A.; Farrag, E.A.E.; Hamed, S.M.; Abdallah, Z.; El Nashar, E.M.; Alghamdi, M.A.; Ali, A.A.H.; Abd El-Kader, M. Neuroprotective and Therapeutic Effects of Calcitriol in Rotenone-Induced Parkinson’s Disease Rat Model. Front. Cell. Neurosci. 2022, 16, 967813. [Google Scholar] [CrossRef] [PubMed]
  160. da Costa, R.O.; Gadelha-Filho, C.V.J.; de Aquino, P.E.A.; Lima, L.A.R.; de Lucena, J.D.; Ribeiro, W.L.C.; Lima, F.A.V.; Neves, K.R.T.; de Barros Viana, G.S. Vitamin D (VD3) Intensifies the Effects of Exercise and Prevents Alterations of Behavior, Brain Oxidative Stress, and Neuroinflammation, in Hemiparkinsonian Rats. Neurochem. Res. 2023, 48, 142–160. [Google Scholar] [CrossRef]
  161. Imam, R.A.; Abdel-Hamed, M.R. Vitamin D3 Promotes Oligodendrogenesis and Modulates Synucleinopathy in Lead-Induced Nigral Pars Compacta Neurotoxicity in Rats. Folia Morphol. 2023, 82, 42–52. [Google Scholar] [CrossRef]
  162. Khosravi, F.; Mirzaei, S.; Hojati, V.; Hashemi, M.; Entezari, M. Co-Administration of Vitamins B12 and D During Pregnancy Have Strong Neuroprotective Effects in Parkinson Disease. Mol. Neurobiol. 2023, 60, 1986–1996. [Google Scholar] [CrossRef]
  163. Claro da Silva, T.; Hiller, C.; Gai, Z.; Kullak-Ublick, G.A. Vitamin D3 Transactivates the Zinc and Manganese Transporter SLC30A10 via the Vitamin D Receptor. J. Steroid Biochem. Mol. Biol. 2016, 163, 77–87. [Google Scholar] [CrossRef] [PubMed]
  164. Jiang, P.; Zhang, L.-H.; Cai, H.-L.; Li, H.-D.; Liu, Y.-P.; Tang, M.-M.; Dang, R.-L.; Zhu, W.-Y.; Xue, Y.; He, X. Neurochemical Effects of Chronic Administration of Calcitriol in Rats. Nutrients 2014, 6, 6048–6059. [Google Scholar] [CrossRef]
  165. Dygaĭ, A.M.; Sherstoboev, E.I.; Gol’dberg, E.D. The Role of SC-1(+)- and Thy-1(+)-Cells in the Regulation of Cytokine Production by Cells of the Bone Marrow Regenerating After Cytostatic Action. Biulleten’ Eksp. Biol. Med. 1998, 125, 374–377. [Google Scholar]
  166. Kesavapany, S.; Amin, N.; Zheng, Y.-L.; Nijhara, R.; Jaffe, H.; Sihag, R.; Gutkind, J.S.; Takahashi, S.; Kulkarni, A.; Grant, P.; et al. p35/cyclin-Dependent Kinase 5 Phosphorylation of Ras Guanine Nucleotide Releasing Factor 2 (RasGRF2) Mediates Rac-Dependent Extracellular Signal-Regulated Kinase 1/2 Activity, Altering RasGRF2 and Microtubule-Associated Protein 1b Distribution in Neurons. J. Neurosci. 2004, 24, 4421–4431. [Google Scholar] [CrossRef]
  167. Carpenter, T.O. Mineral Regulation of Vitamin D Metabolism. Bone Miner. 1989, 5, 259–269. [Google Scholar] [CrossRef] [PubMed]
  168. Kattar, S.D.; Gulati, A.; Margrey, K.A.; Keylor, M.H.; Ardolino, M.; Yan, X.; Johnson, R.; Palte, R.L.; McMinn, S.E.; Nogle, L.; et al. Discovery of MK-1468: A Potent, Kinome-Selective, Brain-Penetrant Amidoisoquinoline LRRK2 Inhibitor for the Potential Treatment of Parkinson’s Disease. J. Med. Chem. 2023, 66, 14912–14927. [Google Scholar] [CrossRef] [PubMed]
  169. Gul, F.H.; Bozkurt, N.M.; Nogay, N.H.; Unal, G. The Neuroprotective Effect of 1,25-Dyhydroxyvitamin D3 (Calcitriol) and Probiotics on the Rotenone-Induced Neurotoxicity Model in SH-SY5Y Cells. Drug Chem. Toxicol. 2025, 48, 72–83. [Google Scholar] [CrossRef] [PubMed]
  170. Joseph, K.M. Muralidhara Combined Oral Supplementation of Fish Oil and Quercetin Enhances Neuroprotection in a Chronic Rotenone Rat Model: Relevance to Parkinson’s Disease. Neurochem. Res. 2015, 40, 894–905. [Google Scholar] [CrossRef]
  171. Rahimmi, A.; Khademerfan, M. Selenium and Resveratrol Attenuate Rotenone-Induced Parkinson’s Disease in Zebrafish Model. Mol. Biol. Rep. 2025, 52, 823. [Google Scholar] [CrossRef]
  172. Pal, R.; Das, N.; Basu, B.R. Exploring Vitamin D and its Analogues as a Possible Therapeutic Intervention in Targeting Inflammatory Mediators in Neurodegenerative Disorders Like Parkinson’s Disease. bioRxiv 2025, 2025.01.10.632401. [Google Scholar] [CrossRef]
  173. Mazzetti, S.; Barichella, M.; Giampietro, F.; Giana, A.; Calogero, A.M.; Amadeo, A.; Palazzi, N.; Comincini, A.; Giaccone, G.; Bramerio, M.; et al. Astrocytes Expressing Vitamin D-Activating Enzyme Identify Parkinson’s Disease. CNS Neurosci. Ther. 2022, 28, 703–713. [Google Scholar] [CrossRef]
  174. Permadi, N.G.A.; Meidiary, A.; Laksmidewi, A.P. The Low Serum Vitamin D Levels as a Risk Factor for Cognitive Impairment in Parkinson’s Disease: A Literature Review. Int. J. Sci. Adv. 2025, 6, 240–248. [Google Scholar] [CrossRef]
  175. Almokhtar, M.; Wikvall, K.; Ubhayasekera, S.J.K.A.; Bergquist, J.; Norlin, M. Motor Neuron-Like NSC-34 Cells as a New Model for the Study of Vitamin D Metabolism in the Brain. J. Steroid Biochem. Mol. Biol. 2016, 158, 178–188. [Google Scholar] [CrossRef]
  176. Malaguarnera, L.; Marsullo, A.; Zorena, K.; Musumeci, G.; Di Rosa, M. Vitamin D3 Regulates LAMP3 Expression in Monocyte Derived Dendritic Cells. Cell. Immunol. 2017, 311, 13–21. [Google Scholar] [CrossRef]
  177. Rizk, S.K.; Ali, E.A.; Sheref, A.A.M.; Tayel, S.G.; El Derbaly, S.A. Vitamin D and Canagliflozin Combination Alleviates Parkinson’s Disease in rats Through Modulation of RAC1/NF-κB/Nrf2 Interaction. Immunopharmacol. Immunotoxicol. 2025, 47, 328–344. [Google Scholar] [CrossRef]
  178. Verma, R.; Kim, J.Y. 1,25-Dihydroxyvitamin D3 Facilitates M2 Polarization and Upregulates TLR10 Expression on Human Microglial Cells. Neuroimmunomodulation 2016, 23, 75–80. [Google Scholar] [CrossRef]
  179. Manjari, S.K.V.; Maity, S.; Poornima, R.; Yau, S.-Y.; Vaishali, K.; Stellwagen, D.; Komal, P. Restorative Action of Vitamin D3 on Motor Dysfunction Through Enhancement of Neurotrophins and Antioxidant Expression in the Striatum. Neuroscience 2022, 492, 67–81. [Google Scholar] [CrossRef]
  180. Kumar, P.T.; Antony, S.; Nandhu, M.S.; Sadanandan, J.; Naijil, G.; Paulose, C.S. Vitamin D3 Restores Altered Cholinergic and Insulin Receptor Expression in the Cerebral Cortex and Muscarinic M3 Receptor Expression in Pancreatic Islets of Streptozotocin Induced Diabetic Rats. J. Nutr. Biochem. 2011, 22, 418–425. [Google Scholar] [CrossRef] [PubMed]
  181. Chen, B.; Wen, X.; Jiang, H.; Wang, J.; Song, N.; Xie, J. Interactions Between Iron and α-Synuclein Pathology in Parkinson’s Disease. Free Radic. Biol. Med. 2019, 141, 253–260. [Google Scholar] [CrossRef] [PubMed]
  182. Gooch, H.; Cui, X.; Anggono, V.; Trzaskowski, M.; Tan, M.C.; Eyles, D.W.; Burne, T.H.J.; Jang, S.E.; Mattheisen, M.; Hougaard, D.M.; et al. 1,25-Dihydroxyvitamin D Modulates L-type Voltage-Gated Calcium Channels in a Subset of Neurons in the Developing Mouse Prefrontal Cortex. Transl. Psychiatry 2019, 9, 281. [Google Scholar] [CrossRef] [PubMed]
  183. Ogundele, O.M.; Nanakumo, E.T.; Ishola, A.O.; Obende, O.M.; Enye, L.A.; Balogun, W.G.; Cobham, A.E.; Abdulbasit, A. -NMDA R/+VDR Pharmacological Phenotype as a Novel Therapeutic Target in Relieving Motor-Cognitive Impairments in Parkinsonism. Drug Chem. Toxicol. 2015, 38, 415–427. [Google Scholar] [CrossRef]
  184. Fetahu, I.S.; Höbaus, J.; Kállay, E. Vitamin D and the Epigenome. Front. Physiol. 2014, 5, 164. [Google Scholar] [CrossRef]
  185. Xie, Y.; Chen, L.; Chen, J.; Chen, Y. Calcitriol Restrains Microglial M1 Polarization and Alleviates Dopaminergic Degeneration in Hemiparkinsonian Mice by Boosting Regulatory T-Cell Expansion. Brain Behav. 2024, 14, e3373. [Google Scholar] [CrossRef]
  186. Bayo-Olugbami, A.; Nafiu, A.B.; Amin, A.; Ogundele, O.M.; Lee, C.C.; Owoyele, B.V. Cholecalciferol (VD3) Attenuates L-DOPA-Induced Dyskinesia in Parkinsonian Mice Via Modulation of Microglia and Oxido-Inflammatory Mechanisms. Niger. J. Physiol. Sci. 2022, 37, 175–183. [Google Scholar] [CrossRef] [PubMed]
  187. Sirajo, M.U.; Oyem, J.C.; Badamasi, M.I. Supplementation with Vitamins D3 and a Mitigates Parkinsonism in a Haloperidol Mice Model. J. Chem. Neuroanat. 2024, 135, 102366. [Google Scholar] [CrossRef] [PubMed]
  188. Umar, M.S.; Ibrahim, B.M. Vitamin A and vitamin D3 Protect the Visual Apparatus During the Development of Dopamine-2 Receptor Knockout Mouse Model of Parkinsonism. J. Complement. Integr. Med. 2023, 20, 577–589. [Google Scholar] [CrossRef]
  189. Zhang, P.; Rhodes, J.S.; Garland, T.J.; Perez, S.D.; Southey, B.R.; Rodriguez-Zas, S.L. Brain Region-Dependent Gene Networks Associated with Selective Breeding for Increased Voluntary Wheel-Running Behavior. PLoS ONE 2018, 13, e0201773. [Google Scholar] [CrossRef]
  190. John Marshal, J.; Kuriakose, B.B.; Alhazmi, A.H.; Muthusamy, K. Mechanistic Insights into the Role of Vitamin D and Computational Identification of Potential Lead Compounds for Parkinson’s Disease. J. Cell. Biochem. 2023, 124, 434–445. [Google Scholar] [CrossRef]
  191. Bytowska, Z.K.; Korewo-Labelle, D.; Kowalski, K.; Libionka, W.; Przewłócka, K.; Kloc, W.; Kaczor, J.J. Impact of 12 Weeks of Vitamin D3 Administration in Parkinson’s Patients with Deep Brain Stimulation on Kynurenine Pathway and Inflammatory Status. Nutrients 2023, 15, 3839. [Google Scholar] [CrossRef] [PubMed]
  192. Alaylıoğlu, M.; Dursun, E.; Genç, G.; Şengül, B.; Bilgiç, B.; Gündüz, A.; Apaydın, H.; Kızıltan, G.; Gürvit, H.; Hanağası, H.; et al. Genetic Variants of Vitamin D Metabolism-Related DHCR7/NADSYN1 Locus and CYP2R1 Gene Are Associated with Clinical Features of Parkinson’s Disease. Int. J. Neurosci. 2022, 132, 439–449. [Google Scholar] [CrossRef]
  193. Hasuike, Y.; Endo, T.; Koroyasu, M.; Matsui, M.; Mori, C.; Yamadera, M.; Fujimura, H.; Sakoda, S. Bile Acid Abnormality Induced by Intestinal Dysbiosis Might Explain Lipid Metabolism in Parkinson’s Disease. Med. Hypotheses 2020, 134, 109436. [Google Scholar] [CrossRef]
  194. Oda, Y.; Sihlbom, C.; Chalkley, R.J.; Huang, L.; Rachez, C.; Chang, C.-P.B.; Burlingame, A.L.; Freedman, L.P.; Bikle, D.D. Two Distinct Coactivators, DRIP/Mediator and SRC/p160, are Differentially Involved in Vitamin D Receptor Transactivation During Keratinocyte Differentiation. Mol. Endocrinol. 2003, 17, 2329–2339. [Google Scholar] [CrossRef]
  195. Oak, A.S.W.; Bocheva, G.; Kim, T.-K.; Brożyna, A.A.; Janjetovic, Z.; Athar, M.; Tuckey, R.C.; Slominski, A.T. Noncalcemic Vitamin D Hydroxyderivatives Inhibit Human Oral Squamous Cell Carcinoma and Down-regulate Hedgehog and WNT/β-Catenin Pathways. Anticancer. Res. 2020, 40, 2467–2474. [Google Scholar] [CrossRef]
  196. Zhang, J.; Chalmers, M.J.; Stayrook, K.R.; Burris, L.L.; Wang, Y.; Busby, S.A.; Pascal, B.D.; Garcia-Ordonez, R.D.; Bruning, J.B.; Istrate, M.A.; et al. DNA Binding Alters Coactivator Interaction Surfaces of the Intact VDR-RXR Complex. Nat. Struct. Mol. Biol. 2011, 18, 556–563. [Google Scholar] [CrossRef]
  197. Feldman, D.; Malloy, P.J. Mutations in the Vitamin D Receptor and Hereditary Vitamin D-Resistant Rickets. BoneKEy Rep. 2014, 3, 510. [Google Scholar] [CrossRef]
  198. Gholipour, M.; Honarmand Tamizkar, K.; Niknam, A.; Hussen, B.M.; Eslami, S.; Sayad, A.; Ghafouri-Fard, S. Expression Analysis of Vitamin D Receptor and Its Related Long Non-Coding RNAs in Peripheral Blood of Patients with Parkinson’s Disease. Mol. Biol. Rep. 2022, 49, 5911–5917. [Google Scholar] [CrossRef]
  199. Hu, T.-M.; Chung, H.-S.; Ping, L.-Y.; Hsu, S.-H.; Tsai, H.-Y.; Chen, S.-J.; Cheng, M.-C. Differential Expression of Multiple Disease-Related Protein Groups Induced by Valproic Acid in Human SH-SY5Y Neuroblastoma Cells. Brain Sci. 2020, 10, 545. [Google Scholar] [CrossRef]
  200. Wang, L.; Maldonado, L.; Beecham, G.W.; Martin, E.R.; Evatt, M.L.; Ritchie, J.C.; Haines, J.L.; Zabetian, C.P.; Payami, H.; Pericak-Vance, M.A.; et al. DNA Variants in CACNA1C Modify Parkinson Disease Risk Only when Vitamin D Level is Deficient. Neurol. Genet. 2016, 2, e72. [Google Scholar] [CrossRef]
  201. Butler, M.W.; Burt, A.; Edwards, T.L.; Zuchner, S.; Scott, W.K.; Martin, E.R.; Vance, J.M.; Wang, L. Vitamin D Receptor Gene as a Candidate Gene for Parkinson Disease. Ann. Hum. Genet. 2011, 75, 201–210. [Google Scholar] [CrossRef] [PubMed]
  202. Zeng, Z.; Xiong, L.; Cen, Y.; Hong, G.; Shen, Y.; Luo, X. Associations of Dietary Intakes with Parkinson’s Disease: Findings from a Cross-Sectional Study. Int. J. Vitam. Nutr. Res. 2024, 95, 36422. [Google Scholar] [CrossRef] [PubMed]
  203. Yong, Y.; Dong, H.; Zhou, Z.; Zhu, Y.; Gu, M.; Li, W. Serum 25-Hydroxyvitamin D Concentrations and Their Impact on All-Cause Mortality in Parkinson’s Disease: Insights from National Health and Nutrition Examination Survey 1999–2020 data. Front. Nutr. 2024, 11, 1423651. [Google Scholar] [CrossRef]
  204. Turcu-Stiolica, A.; Naidin, M.-S.; Halmagean, S.; Ionescu, A.M.; Pirici, I. The Impact of the Dietary Intake of Vitamin B12, Folic Acid, and Vitamin D3 on Homocysteine Levels and the Health-Related Quality of Life of Levodopa-Treated Patients with Parkinson’s Disease-A Pilot Study in Romania. Diagnostics 2024, 14, 1609. [Google Scholar] [CrossRef] [PubMed]
  205. Evatt, M.L.; DeLong, M.R.; Kumari, M.; Auinger, P.; McDermott, M.P.; Tangpricha, V. High Prevalence of Hypovitaminosis D Status in Patients with Early Parkinson Disease. Arch. Neurol. 2011, 68, 314–319. [Google Scholar] [CrossRef]
  206. Baert, F.; Matthys, C.; Mellaerts, R.; Lemaître, D.; Vlaemynck, G.; Foulon, V. Dietary Intake of Parkinson’s Disease Patients. Front. Nutr. 2020, 7, 105. [Google Scholar] [CrossRef]
  207. Ferguson, C.C.; Knol, L.L.; Halli-Tierney, A.; Ellis, A.C. Dietary Supplement Use is High Among Individuals with Parkinson Disease. South. Med. J. 2019, 112, 621–625. [Google Scholar] [CrossRef] [PubMed]
  208. Demay, M.B.; Pittas, A.G.; Bikle, D.D.; Diab, D.L.; Kiely, M.E.; Lazaretti-Castro, M.; Lips, P.; Mitchell, D.M.; Murad, M.H.; Powers, S.; et al. Vitamin D for the Prevention of Disease: An Endocrine Society Clinical Practice Guideline. J. Clin. Endocrinol. Metab. 2024, 109, 1907–1947. [Google Scholar] [CrossRef] [PubMed]
  209. Philippou, E.; Hirsch, M.A.; Heyn, P.C.; van Wegen, E.E.H.; Darwish, H. Vitamin D and Brain Health in Alzheimer and Parkinson Disease. Arch. Phys. Med. Rehabil. 2024, 105, 809–812. [Google Scholar] [CrossRef]
  210. Chatterjee, R.; Chatterjee, K.; Sen, C. Reversible Parkinsonism Due to Vitamin D Toxicity. J. Neurosci. Rural. Pract. 2017, 8, 305–306. [Google Scholar] [CrossRef]
  211. Kosakai, A.; Ito, D.; Nihei, Y.; Yamashita, S.; Okada, Y.; Takahashi, K.; Suzuki, N. Degeneration of Mesencephalic Dopaminergic Neurons in Klotho Mouse Related to Vitamin D Exposure. Brain Res. 2011, 1382, 109–117. [Google Scholar] [CrossRef]
  212. Kuhn, W.; Karp, G.; Müller, T. No Vitamin D Deficiency in Patients with Parkinson’s Disease. Degener. Neurol. Neuromuscul. Dis. 2022, 12, 127–131. [Google Scholar] [CrossRef]
  213. Fullard, M.E.; Xie, S.X.; Marek, K.; Stern, M.; Jennings, D.; Siderowf, A.; Willis, A.W.; Chen-Plotkin, A.S. Vitamin D in the Parkinson Associated Risk Syndrome (PARS) study. Mov. Disord. 2017, 32, 1636–1640. [Google Scholar] [CrossRef] [PubMed]
  214. Shrestha, S.; Lutsey, P.L.; Alonso, A.; Huang, X.; Mosley, T.H.J.; Chen, H. Serum 25-Hydroxyvitamin D Concentrations in Mid-Adulthood and Parkinson’s Disease Risk. Mov. Disord. 2016, 31, 972–978. [Google Scholar] [CrossRef]
  215. Luthra, N.S.; Kim, S.; Zhang, Y.; Christine, C.W. Characterization of Vitamin D Supplementation and Clinical Outcomes in a Large Cohort of Early Parkinson’s Disease. J. Clin. Mov. Disord. 2018, 5, 7. [Google Scholar] [CrossRef]
  216. Miyake, Y.; Tanaka, K.; Fukushima, W.; Sasaki, S.; Kiyohara, C.; Tsuboi, Y.; Yamada, T.; Oeda, T.; Miki, T.; Kawamura, N.; et al. Lack of Association of Dairy Food, Calcium, and Vitamin D Intake with the Risk of Parkinson’s Disease: A Case-Control Study in Japan. Park. Relat. Disord. 2011, 17, 112–116. [Google Scholar] [CrossRef]
  217. Larsson, S.C.; Singleton, A.B.; Nalls, M.A.; Richards, J.B. No Clear Support for a Role for Vitamin D in Parkinson’s Disease: A Mendelian Randomization Study. Mov. Disord. 2017, 32, 1249–1252. [Google Scholar] [CrossRef]
  218. Anderson, C.; Checkoway, H.; Franklin, G.M.; Beresford, S.; Smith-Weller, T.; Swanson, P.D. Dietary Factors in Parkinson’s Disease: The Role of Food Groups and Specific Foods. Mov. Disord. 1999, 14, 21–27. [Google Scholar] [CrossRef] [PubMed]
  219. Chen, H.; Zhang, S.M.; Hernán, M.A.; Willett, W.C.; Ascherio, A. Diet and Parkinson’s Disease: A Potential Role of Dairy Products in Men. Ann. Neurol. 2002, 52, 793–801. [Google Scholar] [CrossRef] [PubMed]
  220. Altun, I.G.; Barut, B.Ö.; Bulut, A.; Padir, N.; Inan, R. The Relationship Between Pain and Vitamin D in Parkinson’s Disease. South. Clin. Istanb. Eurasia 2021, 32, 395–399. [Google Scholar] [CrossRef]
  221. Voltan, G.; Cannito, M.; Ferrarese, M.; Ceccato, F.; Camozzi, V. Vitamin D: An Overview of Gene Regulation, Ranging from Metabolism to Genomic Effects. Genes 2023, 14, 1691. [Google Scholar] [CrossRef]
  222. Jiang, W.; Ju, C.; Jiang, H.; Zhang, D. Dairy Foods Intake and Risk of Parkinson’s Disease: A Dose-Response Meta-Analysis of Prospective Cohort Studies. Eur. J. Epidemiol. 2014, 29, 613–619. [Google Scholar] [CrossRef]
  223. Lv, L.; Tan, X.; Peng, X.; Bai, R.; Xiao, Q.; Zou, T.; Tan, J.; Zhang, H.; Wang, C. The Relationships of Vitamin D, Vitamin D Receptor Gene Polymorphisms, and Vitamin D Supplementation with Parkinson’s Disease. Transl. Neurodegener. 2020, 9, 34. [Google Scholar] [CrossRef] [PubMed]
  224. 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]
  225. Wang, C.; Wang, Y.-L.; Xu, Q.-H. Integrating Network Pharmacology with Molecular Docking and Dynamics to Uncover Therapeutic Targets and Signaling Mechanisms of Vitamin D3 in Parkinson’s Disease. Mol. Divers. 2025, 29, 5661–5677. [Google Scholar] [CrossRef] [PubMed]
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.

Article Metrics

Citations

Article Access Statistics

Multiple requests from the same IP address are counted as one view.