Skip to main content

The relationships of vitamin D, vitamin D receptor gene polymorphisms, and vitamin D supplementation with Parkinson’s disease

Abstract

In recent years, many studies have investigated the correlations between Parkinson’s disease (PD) and vitamin D status, but the conclusion remains elusive. The present review focuses on the associations between PD and serum vitamin D levels by reviewing studies on the associations of PD with serum vitamin D levels and vitamin D receptor (VDR) gene polymorphisms from PubMed, Web of Science, Cochrane Library, and Embase databases. We found that PD patients have lower vitamin D levels than healthy controls and that the vitamin D concentrations are negatively correlated with PD risk and severity. Furthermore, higher vitamin D concentrations are linked to better cognitive function and mood in PD patients. Findings on the relationship between VDR gene polymorphisms and the risk of PD are inconsistent, but the FokI (C/T) polymorphism is significantly linked with PD. The occurrence of FokI (C/T) gene polymorphism may influence the risk, severity, and cognitive ability of PD patients, while also possibly influencing the effect of Vitamin D3 supplementation in PD patients. In view of the neuroprotective effects of vitamin D and the close association between vitamin D and dopaminergic neurotransmission, interventional prospective studies on vitamin D supplementation in PD patients should be conducted in the future.

Introduction

Parkinson’s disease (PD) is the most common type of parkinsonism, a syndrome characterized by bradykinesia, postural instability, rigidity, and resting tremor [1]. The pathophysiological cause of PD is the progressive loss of dopaminergic (DA) neurons in the substantia nigra (SN) of the midbrain [2,3,4] and the formation of Lewy bodies, which are neuronal inclusions mainly consisting of α-synuclein protein aggregations [5,6,7]. In high-income countries, the annual incidence of PD is 14 per 100,000 in the general population, and rises to 160 per 100,000 in the population aged 65 years or older [8]. A systematic review estimated that there were 6.1 million PD patients worldwide in the year 2016, a significant increase from 2.5 million in 1990, with further projected increases in the future. Moreover, the increase cannot entirely be explained by the growth of number of older people [9]. The etiology of PD remains unknown and is presumably multifactorial [10]. The exact mechanism of neurodegeneration in PD is not yet fully elucidated, but it likely involves a series of events including interactions between genetic and environmental factors, oxidative stress, mitochondrial dysfunction, inflammation, immune regulation, and others [11,12,13,14,15,16,17,18,19,20,21,22]. Due to the unclear etiology, no medications have been proven to cure PD [1]. Therefore, there is a critical demand for new and targeted drugs that focus on protecting DA neurons from degeneration in PD.

Vitamin D obtained via sun exposure or through the diet is converted by 25-hydroxylase mainly located in the liver into 25-hydroxyvitamin D (25(OH)D), the major circulating form of vitamin D. The 25(OH)D can be used as a serum marker to measure vitamin D levels in PD patients, however, it is biologically inactive and must be transformed into the active form 1,25-dihydroxyvitamin D3 (1,25(OH)2D3) by 25-hydroxyvitamin D-1α-hydroxylase (CYP27B1) in the kidney. Increased concentrations of 1,25(OH)2D3 can raise the expression of 25-hydroxyvitamin D-24-hydroxylase (CYP24A1) to catabolize 1,25(OH)2D3 into calcitroic acid [23,24,25]. The biological functions of 1,25(OH)2D3 are mediated by the vitamin D receptor (VDR), a member of the nuclear receptor superfamily of transcription factors [25, 26]. Upon ligand binding, the VDR interacts with the retinoid X receptor (RXR) to form a heterodimer, which then binds to vitamin D response elements (VDREs) in target genes to promote their expression [27, 28]. It has been predicted that 1,25(OH)2D3 regulates more than 200 genes, influencing a variety of cellular processes (Fig. 1) [29]. There is ample evidence from in vitro and animal studies that vitamin D plays an important role in cell proliferation and differentiation, neurotrophic regulation and neuroprotection, neurotransmission, immune regulation, and neuroplasticity [30,31,32,33]. Studies have confirmed the presence of vitamin D metabolites, their metabolizing enzymes CYP27B1 and CYP24A1, as well as VDR in the human brain. This indicates that the human brain can regulate 1,25(OH)2D3 levels, and vitamin D may play a key role in the maintenance of normal nervous system function [34, 35]. Moreover, VDR and CYP27B1 expression is most abundant in the substantia nigra (SN; a brain region rich in dopaminergic neurons) according to immunofluorescence [36]. Studies have also found that the earliest time of VDR expression in the midbrain is on embryonic day 12 (E12), which coincides with the time of development of a majority of dopaminergic neurons in the SN region [37, 38]. Vitamin D is a fat-soluble hormone that can pass the blood-brain barrier, which supports its significance in the central nervous system [39].

Fig. 1
figure 1

Vitamin D metabolism. DBP, Vitamin D-binding protein; RXR, retinoid X receptor; VDREs, vitamin D response elements; VDR, vitamin D receptor; 1-OHase, 25-hydroxyvitamin D-1α- hydroxylase; 24-OHase, 25-hydroxyvitamin D-24-hydroxylase

Considering the neuroprotective effect of vitamin D in the human brain, researchers have proposed a ‘two-hit hypotheses’ to explain how low vitamin D levels make the nervous system more susceptible to secondary harmful effects, which may aggravate the development of diseases such as PD and cerebrovascular disease [40,41,42]. Notably, many studies have indicated that vitamin D metabolism may be directly or indirectly related to the pathogenesis of PD [30, 31, 33, 39, 42,43,44]. Further, vitamin D can act as a neuroprotective agent to provide partial protection for DA neurons [45]. Accordingly, inadequate vitamin D status may play a significant role in PD, resulting in a progressive loss of DA neurons in the human brain [46]. However, experimental data from the Asymptomatic Parkinson Associated Vitamin D Intake Risk Syndrome cohort were not consistent with the hypothesis that chronically inadequate levels of vitamin D threaten the integrity of the DA system, leading to the pathogenesis of PD [47]. The conundrum of the connection between serum vitamin D levels and PD therefore remains unsolved. In this paper, we review the serum vitamin D concentrations in PD patients, the relationships of serum vitamin D concentrations and VDR gene polymorphisms with PD risk, the relationship between serum vitamin D concentrations and clinical manifestations of PD patients, as well as the preventive and therapeutic roles of vitamin D in PD.

Main text

PD patients often have low serum vitamin D concentrations

Vitamin D insufficiency is prevalent in the elderly worldwide [48], and it is also a common health problem in neurodegenerative diseases such as PD and Alzheimer’s disease (AD). Recently, it has been reported that vitamin D insufficiency is more common among PD patients than healthy controls [49,50,51]. If the insufficiency of vitamin D is a consequence of neurodegenerative disease, the incidence of vitamin D insufficiency in AD and PD patients should be similar, but a study revealed that vitamin D insufficiency in PD patients was more pronounced than that in AD patients and controls (55% versus 41% and 36%, respectively) [52]. The 1,25(OH)2D3 levels were normal in all PD patients, whereas the serum levels of 25(OH)D were insufficient (< 20 ng/mL) in 49% of patients in a prospective cohort study [53]. This can be explained by the fact that circulating 25(OH)D levels are 1000 times higher than 1,25(OH)2D3, and that the 25(OH)D can be converted into 1,25(OH)2D3 by 1a-OHase [29, 53].

PD patients experience mobility problems more frequently, and the typical course of PD is longer than that of AD. Both factors may decrease sunlight exposure, thus reducing the cutaneous synthesis of vitamin D. Many studies have reported that the more severe the motor symptoms, the lower the 25(OH)D concentrations in PD patients [53,54,55,56,57]. The reduced mobility and sunlight deprivation may be responsible for the higher incidence of vitamin D deficiency in PD patients. However, compared with controls, there are significantly lower levels of 25(OH)D in PD patients with sufficient sunlight exposure [51]. This may be because that as some vitamin D should be from the diet [29], the gastrointestinal dysfunction in PD patients may result in chronically inadequate vitamin D intake [58, 59]. Interestingly, a longitudinal cohort study discovered that the 25(OH)D concentrations were slightly increased over the study period, which means that these patients did not have digestive dysfunction. The study also reported that there was a high incidence of vitamin D insufficiency in subjects with early PD who did not require symptomatic therapy [60].

However, another study showed that compared with controls, the PD patients had slightly yet not significantly lower serum vitamin D concentrations [61]. There may be two factors that affect the results. First, the case-control study was conducted in the Faroe Islands at a high latitude and with harsh climate and frequent cloud cover, which may have decreased sunlight exposure. Second, Faroese food is not rich in vitamin D, which resulted in the common vitamin D insufficiency in the Faroe Islands.

The relationship between vitamin D deficiency and PD risk

Lower vitamin D levels may be a result of PD due to the limited mobility and digestive symptoms of PD patients. However, several studies suggested that vitamin D deficiency may be associated with the etiology of PD [46, 54]. A study showed that the prevalence of vitamin D deficiency was higher among patients with PD, even if they had normal ambulation and gastrointestinal functions [60]. Newmark and colleagues concluded that chronic vitamin D deficiency is likely to be linked to the pathogenesis or progression of PD rather than only being a consequence of the disease [46, 54]. This hypothesis was also supported by a 29-year prospective study in Finland, which confirmed that those who had higher vitamin D concentrations were less likely to develop PD. When comparing probands in the highest and the lowest quartiles of vitamin D levels, the relative risk of PD was 0.33 for the highest quartile (95% confidence interval, 0.14–0.80) [62]. In line with these findings, a large case-control sample study revealed a negative correlation between PD risk and the level of 25(OH)D, and additionally showed an inverse correlation between 25(OH)D2 and PD [59]. Moreover, Danish and Chinese case-control studies both suggested that outdoor work can reduce the risk of PD in later life [63,64,65]. One possible protective mechanism of outdoor work is to increase sunlight exposure, which contributes to vitamin D3 synthesis in the skin. Interestingly, a nationwide ecological study in France showed that increasing sunlight exposure can reduce the risk of PD in the young population [66]. However, a population-based prospective study with 17 years of follow-up and a Mendelian randomization study did not explore the association between 25(OH)D concentrations and the prevalence of PD [67, 68].

The relationship between VDR gene polymorphisms and PD risk

The VDR is the key mediator of the functions of vitamin D. A transcriptome-wide scan indicated that the VDR gene expression is increased in the blood cells of early-stage PD patients [69]. Consequently, it is reasonable that the VDR polymorphic variants might also have an effect on the pathogenesis of PD. In recent years, the polymorphisms of BsmI (rs1544410), FokI (rs10735810), ApaI (rs7975232), and TaqI (rs731236) have been most widely studied in research on the correlation between VDR gene variants and PD, but the results were inconsistent [70,71,72]. A polymerase chain reaction-based restriction analysis of VDR gene polymorphisms in Korea indicated that the BsmI (B/b) polymorphism is a candidate allele influencing the pathogenesis of PD. The study further showed that the bb genotype was more common in the group with predominant postural instability and gait disorders than in the tremor-predominant group and the healthy controls [73]. In addition, Hungarian, Japanese and Chinese studies suggested that the FokI (C/T) polymorphism located in exon 2 in the 5′ coding region of the gene was significantly linked with PD, and the C allele can increase the risk of PD [74,75,76,77].

The most significant start codon polymorphism of the VDR gene is the functional FokI polymorphism, which results in different translation initiation sites, one producing a long version of the VDR protein (the T-allele) and the other producing a protein shortened by three amino acids (the C-allele) [70]. In spite of the small difference, the functional characteristics of the two forms of VDR (C-VDR and T-VDR) are significantly different. Compared with T-VDR, the C-VDR has a better capacity for intestinal calcium absorption [70, 78]. Therefore, the C allele may forecast higher vitamin D levels and reduce the risk of PD. However, research findings have suggested that the C allele is a risk factor for PD rather than being a protective factor [74,75,76,77]. Suzuki et al. revealed that there was a stronger association of the FokI CC genotype with milder forms of PD (odds ratio, 0.32; 95% confidence interval, 0.16–0.66) [53]. Moreover, the Parkinson Environment Gene study, a population-based case-control study of PD in the Central Valley of California, showed that FokI polymorphism was linked to cognitive decline in PD [79].

In 2015, a study in California with higher ultraviolet radiation levels than in previous studies of VDR gene polymorphisms showed that the major allele TaqI TT genotype and the ApaI GG genotype are associated with decreased risk of PD [80]. However, some studies did not find any association between the VDR genotypes (BsmI, FokI, ApaI, and TaqI loci) and PD risk [61, 81]. The different results may be explained by a number of reasons. First, the effect of VDR gene polymorphism on PD risk may be related to the vitamin D levels. Second, these studies involved different ethnicities, environmental factors, gene-gene and gene-environment interactions, or small sample sizes. Therefore, future studies should shift to the interactions of vitamin D levels and VDR gene polymorphisms in PD, and take into account the environmental factors.

The relationship between vitamin D level and clinical manifestations of PD patients

There is accumulating evidence that the PD patients have an increased prevalence of osteoporosis and osteopenia [82,83,84], and PD is recognized as a cause of secondary osteoporosis [85]. In a study conducted in Korea, researchers found that 6542 (18.3%) of 35,663 PD patients experienced osteoporosis, and that fractures occurred most commonly within 6 months after PD onset and decreased after 3 years from PD diagnosis [86]. Other studies have shown that bone loss and fractures in PD patients are multifactorial [87,88,89], with causes including vitamin D deficiency [90]. As PD patients experience more bone loss, more falls and more fractures (particularly at the hip) [91] than controls, osteoporosis should be screened and treated early [51, 82, 84, 92,93,94], particularly for older female patients within 3 years of PD diagnosis [86]. A meta-analysis of randomized controlled trials found that daily supplementation of 700 IU to 1000 IU of vitamin D could reduce the risk of falls by 19%, and that this advantage may not be dependent on additional calcium supplementation [95].

Many recent studies have confirmed the significant negative correlation between the severity of PD evaluated using the Hoehn & Yahr scale or Unified Parkinson’s Disease Rating Scale (UPDRS) and the circulating serum 25(OH)D levels [53,54,55,56,57, 96]. Prospective observational studies have also found a negative association between the vitamin D status at baseline and severity of PD motor symptoms during disease progression [41, 97]. Therefore, supplementation of vitamin D may delay the worsening of symptoms in PD patients. A cross-sectional, observational study supported the relationship between postural balance and serum vitamin D levels. Further analysis showed that among balance measures, vitamin D levels were associated with an automatic posture response to backwards translation, particularly with response strength and weight-bearing asymmetry [98].

PD patients often ignore nonmotor symptoms, which may, however, have been present for years before diagnosis. A large population-based sample of French older people found a strong relationship of lower 25(OH)D concentrations with cognitive decline, as well as increased risk of dementia and AD over 12 years of follow-up [99]. Likewise, in a sample of PD patients without cognitive impairment, higher vitamin D levels were associated with better cognition and mood [100]. The impact of vitamin D on cognition can partially be explained by its effect on amyloid beta (Aβ) [101], which has been shown to deposit in PD as well, probably leading or contributing to cognitive decline [102]. Interestingly, vitamin D has been reported to affect the Aβ-producing enzymes BACE1 and γ-secretase to reduce Aβ anabolism and elevate Aβ catabolism. Furthermore, vitamin D3 could reduce the cytotoxicity of Aβ peptide by ameliorating the decrease of the sphingosine-1-phosphate/ceramide ratio caused by Aβ [103]. A randomized double-blind trial found that vitamin D supplementation can effectively reduce the levels of Aβ, amyloid precursor protein (APP), BACE1, APP mRNA, and BACE1 mRNA [104]. Vitamin D and its receptors are important components of neuronal amyloid processing pathways [105]. Mayne et al. found that vitamin D deficiency may affect synaptic plasticity, leading to a decline of cognition [31]. Studies have shown that vitamin D signaling can affect the expression of L-type voltage-gated calcium channels, which are involved in neurotransmitter release, neuronal excitability change, learning and memory, etc. [31, 106]. Treatment of aging rodents with high-dose vitamin D3 could prevent cognitive decline and enhance hippocampal synaptic excitability [107].

Many studies have shown that brain regions involved in regulating olfactory function are closely related to cognitive decline, and the severity of olfactory disorder in PD patients may precede dementia [108,109,110]. In 2018, Kim et al. firstly demonstrated that the 25(OH)D3 levels were correlated with the severity of olfactory dysfunction in PD [111]. According to the Braak model of neuropathological staging of PD, the early stages 1 and 2 start from the medulla and the olfactory bulb [112], supporting a relationship between vitamin D and early PD. In addition to being associated with dementia and olfactory function in PD patients, serum 25(OH)D3 concentrations can also affect the gastric emptying time [113] and orthostatic hypotension [114].

Preventive and therapeutic effects of vitamin D in PD

The pathophysiology of PD is affected by 1,25(OH)2D3 via genomic (Table 1) and non-genomic routes (rapid vitamin D-dependent membrane-associated effects) [120]. 1,25(OH)2D3 can increase or decrease the expression of a number of genes, thereby affecting intracellular signaling pathways. Recent pieces of evidence suggest that there is an inverse correlation between vitamin D concentrations and PD risk.

Table 1 Effects of 1,25-dihydroxyvitamin D3 exposure on gene expression in PD

1,25(OH)2D3 affects PD by genomic actions mediated by VDR

Neuroprotective effects of vitamin D

Glial cell-derived neurotrophic factors (GDNFs) facilitate neuronal regrowth and protect dopaminergic nerve terminals, which make them a very promising candidate for neuro-restoration therapy of PD [121, 122]. GDNF binds to the GDNF family receptor alpha 1 (GFRa1) and then associates with the proto-oncogene tyrosine-protein kinase receptor Ret (C-Ret). This complex enables GDNF to exert intracellular signaling in DA neurons [115]. However, GDNF cannot pass the blood-brain barrier, and injecting GDNF into the CNS has many negative effects [121]. As a fat-soluble vitamin, 1,25(OH)2D3 can pass the blood-brain barrier, which consolidates the importance of this hormone in PD [39]. Upon VDR binding, 1,25(OH)2D3 directly upregulates the transcription of genes targeted by C-Ret and GDNF. There is a positive feedback between GDNF and C-Ret, and both can suppress GFRa1 production [115]. Depletion of 1,25(OH)2D3 results in decreased expression of GDNF, Nurr1 and p57kip2 [123, 124], which may alter the differentiation and maturation of DA neurons in the developing rat brain [116, 117]. Nurr1 is also crucial for the expression of C-Ret [115], which in turn triggers Src-family kinases and tyrosine kinase, activating several downstream signaling cascades, including the phosphoinositide 3-kinase (PI3K) pathway, the phospholipase Cγ (PLC-γ) pathway, and the p42/p44 mitogen-activated protein kinase (MAPK) pathway [125, 126]. The activation of the MAPK pathway requires a basal activity of the PI3K pathway [126]. The activation of these pathways may promote the survival and differentiation of midbrain DA neurons (Fig. 2). In conclusion, 1,25(OH)2D3 exerts its neuroprotective effects by increasing the expression of GDNF gene and then activating several downstream intracellular signaling cascades.

Fig. 2
figure 2

1,25(OH)2D3 exerts neuroprotective effects via GDNF. CAT, catalase; C-Ret, proto-oncogene tyrosine-protein kinase receptor Ret; DA, dopaminergic; GFR α1, GDNF family receptor alpha 1; GPx, glutathione peroxidase; MAPK pathway, p42/p44 mitogen-activated protein kinase pathway; PLC-γ pathway, phospholipase Cγ pathway; PI3K pathway, phosphoinositide 3-kinase pathway; ROS, reactive oxygen species; RXR, retinoid X receptor; SOD, superoxide dismutase; VDR, vitamin D receptor; VDREs, vitamin D response elements. The arrows indicate signaling components that are either enhanced (red arrows) or reduced (green arrows)

In addition, 1,25(OH)2D3 is also an antioxidant, which may further contribute to its neuroprotective effects. Studies have demonstrated that 1,25(OH)2D3 increases the expression of GDNF, a powerful antioxidant that can reduce reactive oxygen species (ROS). GDNF markedly increases the levels of superoxide dismutase, glutathione peroxidase and catalase in the striatum (Fig. 2) [127]. In addition to the upregulation of GDNF expression, 1,25(OH)2D3 can also exert its antioxidant effect through genomic and/or nongenomic activation. Under inflammatory stimulation, microglial cells can produce 1,25(OH)2D3 in situ, where it potentiates the mRNA expression of gamma-glutamyl transferase (γ-GT) and γ-GT activity induced by proinflammatory stimuli. γ-GT mediates the import of glutathione (GSH) into the cell, after which the intracellular GSH reduces the production of reactive nitrogen species and hydrogen peroxide [128]. In addition, 1,25(OH)2D3 also increases the expression of the nuclear factor erythroid 2-related factor 2 (Nrf2). When ROS rise, they bind to antioxidant response elements (AREs) in the nucleus, enhancing the expression of antioxidant genes, detoxifying enzymes and various signaling components. By increasing the expression of Fos and JUN, Nrf2 also increases the expression of both VDR and RXR [129]. Moreover, 1,25(OH)2D3 can directly inhibit lipid peroxidation as a membrane antioxidant, which protects the membranes of normal cells from ROS-induced oxidative damage [130, 131]. Therefore, 1,25(OH)2D3 contributes to the enhancement of antioxidative systems by increasing the expression of GDNF, γ-GT and Nrf2 (Fig. 3).

Fig. 3
figure 3

1,25(OH)2D3 also exerts neuroprotective effects through genomic and/or non-genomic activation. ARE, antioxidant response element; GFR α1, GDNF family receptor alpha 1; γ-GT, gamma-glutamyl transferase; GSH, glutathione; GSNOH, S-nitrosoglutathione; RXR, retinoid X receptor; ROS, reactive oxygen species; PMCA, plasma membrane Ca2+ ATP-ase; Nrf2, nuclear factor erythroid 2-related factor 2; VDR, vitamin D receptor; VDREs, vitamin D response elements. The arrows indicate signaling components that are either enhanced (red arrows) or reduced (green arrows)

It has also been reported that 1,25(OH)2D3 has anti-inflammatory properties. It can attenuate pro-inflammatory and upregulate anti-inflammatory processes [132]. In the 6-OHDA-induced PD model, pre- or post-treatment with 1,25(OH)2D3 reduced tissue immunopositivity for TNF-α, partially restored tyrosine hydroxylase (TH) immunoreactivity, and prevented the decrease of VDR immunoreactivity in the lesioned striatum [133, 134]. Cell culture studies revealed that the increased intracellular free calcium can induce the aggregation of α-synuclein, and proved that the increase of intracellular calcium and oxidative stress can act cooperatively to promote α-synuclein aggregation [135,136,137,138]. By reducing the expression of L-type Ca2+ channels and increasing the expression of the plasma membrane Ca2+ ATP-ase, NCX1, anti-apoptotic factor Bcl-2 and buffering protein calbindin D28k, 1,25(OH)2D3 can maintain the low cytosolic Ca2+ concentrations and thereby protect against Ca2+-induced oxidative damage in SN dopaminergic neurons [106, 129]. High concentrations of divalent metal ions exhibit toxic effects that may cause an elevation of ROS levels and mitochondrial dysfunction, and even induce neuronal cell death. Notably, 1,25(OH)2D3 can maintain zinc, iron and manganese homeostasis by regulating the expression of related genes. It can transactivate SLC30A10 to increase the expression of zinc and manganese transporter ZnT10. It can also decrease the expression of SLC39A2, which encodes the ZIP (SLC39) protein implicated in zinc, iron and/or manganese transport. The ZnT (SLC30) protein reduces the cytoplasmic concentrations of metal ions, while ZIP (SLC39) transporters lead to an increase (Fig. 3) [118]. In short, 1,25(OH)2D3 can maintain the homeostasis of calcium, zinc, iron and manganese by regulating the expression of some genes, thereby reducing oxidative stress and mitochondrial damage.

Vitamin D is closely associated with dopaminergic neurotransmission

Studies have shown that 1,25(OH)2D3 and VDR are directly involved in regulating the expression of genes in dopaminergic neurons [139]. Many studies have found that VDR protein levels and TH expression are enhanced in the brains of rats following 1,25(OH)2D administration [119, 140]. Notably, TH is the rate-limiting enzyme of dopamine synthesis. It has been reported that a likely mediator of the regulation of TH expression by vitamin D is N-cadherin [141]. In line with these findings, researchers found that pre- or post-treatments with 1,25(OH)2D3 restored the decreased DA content, and increased the expression of TH and dopamine transporter in 6-OHDA-lesioned rats according to striatal neurochemical and immunohistochemical assays [133]. GDNF can act directly on DA neurons to enhance their activity and increase DA release [127]. In a word, 1,25(OH)2D3 may participate in dopaminergic neurotransmission via TH expression regulation and the direct effect of GDNF on DA neurons, which mediates the relationship between vitamin D concentrations and the severity of PD.

1,25(OH)2D3 affects PD by rapid vitamin D-dependent membrane-associated effects

Protein disulfide isomerase 3 (PDIA3), also known as the endoplasmic reticulum stress protein 57 (ERp57), acts as another 1,25(OH)2D3 membrane receptor [142]. Compared to the kidney and liver, PDIA3 is highly expressed in all types of brain cells and can be considered as the main VDR in the brain [143]. It is a multifunctional protein that can not only control the quality of protein processing, but also maintain Ca2+ homeostasis and regulate cellular stress responses [144, 145]. In the PD model induced by 6-OHDA, the level of PDIA3 protein in the striatum is increased, which may be a cellular response to oxidative stress. In this case, PDIA3 may act as a chaperone to prevent the misfolding and aggregation of α-synuclein [144]. After generation of ROS by 6-OHDA, protein oxidation occurs first, and early in the protein oxidation process, the PDIA3 rapidly forms juxtanuclear aggresome-like structures (ERp57/DNA) in dopaminergic cells, which may induce downstream sequelae such as the unfolded protein response, cell stress, and apoptosis [146, 147]. ERp57 has an affinity for Ref-1, which has a synergistic effect and jointly regulates the gene expression mediated by redox-sensitive transcription factors and the adaptive responses of cells to oxidative damage [147]. As a result, it is likely that vitamin D functions in the PD through PDIA3.

Vitamin D supplementation in PD patients

The most important studies on vitamin D supplementation in PD patients are shown in Table 2. An interventional trial supported the role of vitamin D in postural balance of PD patients and suggested that daily supplementation of vitamin D could improve the balance of younger PD patients [148]. Other studies have also confirmed that supplemental 25(OH)D has beneficial effects on strength and balance in older adults [149]. Therefore, there is a debate on whether vitamin D supplementation can specifically delay the progression of motor symptoms in PD patients, or only lead to a non-specific improvement in muscle strength and balance. However, the complex automatic postural response not only requires muscle function, but also involves the spinal cord, midbrain/brainstem, and cerebellum/basal ganglia/cerebral cortex [150, 151]. In addition, vitamin D3 supplementation has an age-dependent effect on PD [148]. Another randomized controlled trial of vitamin D supplementation found that vitamin D3 supplementation may retard the progression of PD for a short period in patients with FokI CT and TT genotypes [41]. Therefore, the extensive roles of vitamin D in the skeletal muscle and neural systems suggest that vitamin D can affect the symptoms of PD.

Table 2 Vitamin D3 supplementation in PD patients

Conclusions and future directions

In summary, the most consistent view at present is that the concentration of vitamin D is low in PD patients. Higher vitamin D concentrations are linked to reduced risk and severity of PD, as well as better cognition and mood of the patients. Furthermore, the VDR gene phenotypes may influence the risk and severity of PD, as well as the effect of vitamin D supplementation in PD patients. Although there are limited data on the effectiveness of vitamin D3 supplementation in PD patients, related studies have highlighted the effectiveness of vitamin D3 supplementation in preventing osteoporotic fractures in the aging population and retarding the progression of PD for a short period.

A recent study has found that vitamin D has the potential to be used as a biomarker for PD [152], inspiring great interest in the relationship between PD and vitamin D. Vitamin D can improve protein homeostasis and slow down the aging process [153], but vitamin D insufficiency is prevalent worldwide [48]. Moreover, vitamin D supplementation is readily available, affordable and safe. The earliest detectable side-effects of vitamin D supplementation are hypercalciuria and hypercalcemia, which are only a concern when 25(OH)D levels exceed 88 ng/mL (220 nmol/L) [154, 155]. Vitamin D supplementation in PD patients at a dose of 1200 IU/day for 12 months [41] or 10,000 IU/day for 16 weeks [148] did not lead to obvious adverse events such as hypercalcemia (Table 2). Therefore, vitamin D supplementation in PD patients seems to be promising, although the dose of vitamin D that may cause toxicity remains unclear. Despite the limited long-term safety data, in 2010, the Institute of Medicine (IOM) defined a safe upper limit dosage for vitamin D of 4000 IU/day, although practitioners should keep in mind the intake of other dietary supplements [156]. The possibility of neuroprotection is the most exciting aspect of vitamin D therapy in PD. Considering the neuroprotective effects of vitamin D and the role of vitamin D in dopaminergic neurotransmission, interventional prospective studies on vitamin D supplementation in PD patients should be conducted in the future.

Availability of data and materials

All data generated or analyzed during this study are included in this published article.

Abbreviations

ARE:

Antioxidant response element

Aβ:

Amyloid beta

APP:

Amyloid precursor protein

C-Ret:

Proto-oncogene tyrosine-protein kinase receptor Ret

DA:

Dopaminergic

ERp57:

Endoplasmic reticulum stress protein 57

γ-GT:

Gamma-glutamyl transferase

GSH:

Glutathione

GDNF:

Glia-derived neurotrophic factors

GFRa1:

GDNF family receptor alpha 1

MAPK pathway:

p42/p44 mitogen-activated protein kinase pathway

Nrf2:

Nuclear factor erythroid 2-related factor 2

PD:

Parkinson’s disease

PLC-γ pathway:

Phospholipase Cγ pathway

PMCA:

Plasma membrane Ca2+ ATP-ase

PI3K pathway:

Phosphoinositide 3-kinase pathway

PDIA3:

Protein disulfide isomerase 3

ROS:

Reactive oxygen species

RXR:

Retinoid X receptor

SN:

Substantia nigra

TH:

Tyrosine hydroxylase

UPDRS:

Unified Parkinson’s Disease Rating Stage

VDR:

Vitamin D receptor

VDREs:

Vitamin D response elements

25(OH)D:

25-hydroxyvitamin D

1,25(OH)2D3 :

1,25-dihydroxyvitamin D3

1-OHase:

25-hydroxyvitamin D-1α-hydroxylase

24-OHase:

25-hydroxyvitamin D-24-hydroxylase

References

  1. Armstrong MJ, Okun MS. Diagnosis and treatment of Parkinson disease: a review. JAMA. 2020;323:548–60.

    Article  PubMed  Google Scholar 

  2. Benazzouz A, Mamad O, Abedi P, Bouali-Benazzouz R, Chetrit J. Involvement of dopamine loss in extrastriatal basal ganglia nuclei in the pathophysiology of Parkinson's disease. Front Aging Neurosci. 2014;6:87.

    Article  PubMed  PubMed Central  Google Scholar 

  3. Pakkenberg B, Moller A, Gundersen HJ, Mouritzen Dam A, Pakkenberg H. The absolute number of nerve cells in substantia nigra in normal subjects and in patients with Parkinson’s disease estimated with an unbiased stereological method. J Neurol Neurosurg Psychiatry. 1991;54:30–3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Hornykiewicz O. The discovery of dopamine deficiency in the parkinsonian brain. J Neural Transm Suppl. 2006;9:20015.

    Google Scholar 

  5. Wakabayashi K, Tanji K, Mori F, Takahashi H. The Lewy body in Parkinson's disease: molecules implicated in the formation and degradation of alpha-synuclein aggregates. Neuropathology. 2007;27:494–506.

    Article  PubMed  Google Scholar 

  6. Shoji TMM, Imai Y, Inoue H, Kawarabayashi T, Matsubara E, Sasaki YHA, et al. Pael-R is accumulated in Lewy bodies of Parkinson's disease. Ann Neurol. 2004;55:439–42.

    Article  PubMed  CAS  Google Scholar 

  7. Bodner RA, Outeiro TF, Altmann S, Maxwell MM, Cho SH, Hyman BT, et al. Pharmacological promotion of inclusion formation: a therapeutic approach for Huntington’s and Parkinson’s diseases. Proc Natl Acad Sci U S A. 2006;103:4246–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Ascherio A, Schwarzschild MA. The epidemiology of Parkinson's disease: risk factors and prevention. Lancet Neurol. 2016;15:1257–72.

    Article  PubMed  Google Scholar 

  9. Dorsey ER, Elbaz A, Nichols E, Abd-Allah F, Abdelalim A, Adsuar JC, et al. Global, regional, and national burden of Parkinson’s disease, 1990–2016: a systematic analysis for the global burden of disease study 2016. Lancet Neurol. 2018;17:939–53.

    Article  Google Scholar 

  10. Logroscino G. The role of early life environmental risk factors in Parkinson disease: what is the evidence? Environ Health Perspect. 2005;113:1234–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Feany MB. New genetic insights into Parkinson's disease. N Engl J Med. 2004;351:1937–40.

    Article  CAS  PubMed  Google Scholar 

  12. Meredith GE, Halliday GM, Totterdell S. A critical review of the development and importance of proteinaceous aggregates in animal models of Parkinson's disease: new insights into Lewy body formation. Parkinsonism Relat Disord. 2004;10:191–202.

    Article  PubMed  Google Scholar 

  13. Cory-Slechta DA, Thiruchelvam M, Barlow BK, Richfield EK. Developmental pesticide models of the Parkinson disease phenotype. Environ Health Perspect. 2005;113:1263–70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Ahlskog JE. Challenging conventional wisdom: the etiologic role of dopamine oxidative stress in Parkinson’s disease. Mov Disord. 2005;20:271–82.

    Article  PubMed  Google Scholar 

  15. Greenamyre JT, Hastings TG. Parkinson’s--divergent causes, convergent mechanisms. Science. 2004;304:1120–2.

    Article  CAS  PubMed  Google Scholar 

  16. Ghavami S, Shojaei S, Yeganeh B, Ande SR, Jangamreddy JR, Mehrpour M, et al. Autophagy and apoptosis dysfunction in neurodegenerative disorders. Prog Neurobiol. 2014;112:24–49.

    Article  CAS  PubMed  Google Scholar 

  17. Dehay B, Martinez-Vicente M, Caldwell GA, Caldwell KA, Yue Z, Cookson MR, et al. Lysosomal impairment in Parkinson’s disease. Mov Disord. 2013;28:725–32.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Bové J, Prou D, Perier C, Przedborski S. Toxin-induced models of Parkinson’s disease. NeuroRx. 2005;2:484–94.

    Article  PubMed  PubMed Central  Google Scholar 

  19. Lim KL, Zhang C. Molecular events underlying Parkinson’s disease – an interwoven tapestry. Front Neurol. 2013;4:33.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  20. Xiong H, Wang D, Chen L, Choo YS, Ma H, Tang C, et al. Parkin, PINK1, and DJ-1 form a ubiquitin E3 ligase complex promoting unfolded protein degradation. J Clin Invest. 2009;119:650–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Selvaraj S, Sun Y, Watt JA, Wang S, Lei S, Birnbaumer L, et al. Neurotoxin-induced ER stress in mouse dopaminergic neurons involves downregulation of TRPC1 and inhibition of AKT/mTOR signaling. J Clin Invest. 2012;122:1354–67.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Maiti P, Manna J, Dunbar GL. Current understanding of the molecular mechanisms in Parkinson's disease: targets for potential treatments. Transl Neurodegener. 2017;6:28.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  23. Tuckey RC, Cheng CYS, Slominski AT. The serum vitamin D metabolome: what we know and what is still to discover. J Steroid Biochem Mol Biol. 2019;186:4–21.

    Article  CAS  PubMed  Google Scholar 

  24. Christakos S, Ajibade DV, Dhawan P, Fechner AJ, Mady LJ. Vitamin D: metabolism. Endocrinol Metab Clin N Am. 2010;39:243–53.

    Article  CAS  Google Scholar 

  25. Christakos S, Dhawan P, Verstuyf A, Verlinden L, Carmeliet G. Vitamin D: metabolism, molecular mechanism of action, and pleiotropic effects. Physiol Rev. 2016;96:365–408.

    Article  CAS  PubMed  Google Scholar 

  26. DeLuca HF. Overview of general physiologic features and functions of vitamin D. Am J Clin Nutr. 2004;80:1689S–96S.

    Article  CAS  PubMed  Google Scholar 

  27. Rachez C, Freedman LP. Mechanisms of gene regulation by vitamin D3 receptor: a network of coactivator interactions. Gene. 2000;246:9–21.

    Article  CAS  PubMed  Google Scholar 

  28. Takeyama K, Kato S. The vitamin D3 1alpha-hydroxylase gene and its regulation by active vitamin D3. Biosci Biotechnol Biochem. 2011;75:208–13.

    Article  CAS  PubMed  Google Scholar 

  29. Holick MF. Vitamin D deficiency. N Engl J Med. 2007;357:266–81.

    Article  CAS  PubMed  Google Scholar 

  30. Almeras L, Eyles D, Benech P, Laffite D, Villard C, Patatian A, et al. Developmental vitamin D deficiency alters brain protein expression in the adult rat: implications for neuropsychiatric disorders. Proteomics. 2007;7:769–80.

    Article  CAS  PubMed  Google Scholar 

  31. Mayne PE, Burne THJ. Vitamin D in synaptic plasticity, cognitive function, and neuropsychiatric illness. Trends Neurosci. 2019;42:293–306.

    Article  CAS  PubMed  Google Scholar 

  32. Bikle DD. Vitamin D metabolism, mechanism of action, and clinical applications. Chem Biol. 2014;21:319–29.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Groves NJ, McGrath JJ, Burne TH. Vitamin D as a neurosteroid affecting the developing and adult brain. Annu Rev Nutr. 2014;34:117–41.

    Article  CAS  PubMed  Google Scholar 

  34. Harms LR, Burne TH, Eyles DW, McGrath JJ. Vitamin D and the brain. Best Pract Res Clin Endocrinol Metab. 2011;25:657–69.

    Article  CAS  PubMed  Google Scholar 

  35. Eyles D, Burne T, McGrath J. Vitamin D in fetal brain development. Semin Cell Dev Biol. 2011;22:629–36.

    Article  CAS  PubMed  Google Scholar 

  36. Eyles DW, Smith S, Kinobe R, Hewison M, McGrath JJ. Distribution of the vitamin D receptor and 1 alpha-hydroxylase in human brain. J Chem Neuroanat. 2005;29:21–30.

    Article  CAS  PubMed  Google Scholar 

  37. Gates MA, Torres EM, White A, Fricker-Gates RA, Dunnett SB. Re-examining the ontogeny of substantia nigra dopamine neurons. Eur J Neurosci. 2006;23:1384–90.

    Article  PubMed  Google Scholar 

  38. Veenstra TD, Prufer K, Koenigsberger C, Brimijoin SW, Grande JP, Kumar R. 1,25-Dihydroxyvitamin D3 receptors in the central nervous system of the rat embryo. Brain Res. 1998;804:193–205.

    Article  CAS  PubMed  Google Scholar 

  39. DeLuca GC, Kimball SM, Kolasinski J, Ramagopalan SV, Ebers GC. Review: the role of vitamin D in nervous system health and disease. Neuropathol Appl Neurobiol. 2013;39:458–84.

    Article  CAS  PubMed  Google Scholar 

  40. Balden R, Selvamani A, Sohrabji F. Vitamin D deficiency exacerbates experimental stroke injury and dysregulates ischemia-induced inflammation in adult rats. Endocrinology. 2012;153:2420–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Suzuki M, Yoshioka M, Hashimoto M, Murakami M, Noya M, Takahashi D, et al. Randomized, double-blind, placebo-controlled trial of vitamin D supplementation in Parkinson disease. Am J Clin Nutr. 2013;97:1004–13.

    Article  CAS  PubMed  Google Scholar 

  42. Cui X, Gooch H, Groves NJ, Sah P, Burne TH, Eyles DW, et al. Vitamin D and the brain: key questions for future research. J Steroid Biochem Mol Biol. 2015;148:305–9.

    Article  CAS  PubMed  Google Scholar 

  43. Eyles DW, Feron F, Cui X, Kesby JP, Harms LH, Ko P, et al. Developmental vitamin D deficiency causes abnormal brain development. Psychoneuroendocrinology. 2009;34(Suppl 1):S247–57.

    Article  CAS  PubMed  Google Scholar 

  44. Garcion E, Wion-Barbot N, Montero-Menei CN, Berger FO, Wion D. New clues about vitamin D functions in the nervous system. Trends Endocrinol Metab. 2002;13:100–5.

    Article  CAS  PubMed  Google Scholar 

  45. Smith MP, Fletcher-Turner A, Yurek DM, Cass WA. Calcitriol protection against dopamine loss induced by intracerebroventricular administration of 6-hydroxydopamine. Neurochem Res. 2006;31:533–9.

    Article  CAS  PubMed  Google Scholar 

  46. Newmark HL, Newmark J. Vitamin D and Parkinson’s disease--a hypothesis. Mov Disord. 2007;22:461–8.

    Article  PubMed  Google Scholar 

  47. Fullard ME, Xie SX, Marek K, Stern M, Jennings D, Siderowf A, et al. Vitamin D in the Parkinson associated risk syndrome (PARS) study. Mov Disord. 2017;32:1636–40.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Dawson-Hughes B, Mithal A, Bonjour JP, Boonen S, Burckhardt P, Fuleihan GE, et al. IOF position statement: vitamin D recommendations for older adults. Osteoporos Int. 2010;21:1151–4.

    Article  CAS  PubMed  Google Scholar 

  49. Ding H, Dhima K, Lockhart KC, Locascio JJ, Hoesing AN, Duong K, et al. Unrecognized vitamin D3 deficiency is common in Parkinson disease: Harvard Biomarker Study. Neurology. 2013;81:1531–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Hatem AK, Lateef HF. The state of Vitamin D in Iraqi patients With Parkinson disease. Al Kindy Coll Med J. 2017;13:137–41.

    Article  Google Scholar 

  51. van den Bos F, Speelman AD, van Nimwegen M, van der Schouw YT, Backx FJ, Bloem BR, et al. Bone mineral density and vitamin D status in Parkinson's disease patients. J Neurol. 2013;260:754–60.

    Article  CAS  PubMed  Google Scholar 

  52. Evatt ML, Delong MR, Khazai N, Rosen A, Triche S, Tangpricha V. Prevalence of vitamin d insufficiency in patients with Parkinson disease and Alzheimer disease. Arch Neurol. 2008;65:1348–52.

    Article  PubMed  PubMed Central  Google Scholar 

  53. Suzuki M, Yoshioka M, Hashimoto M, Murakami M, Kawasaki K, Noya M, et al. 25-hydroxyvitamin D, vitamin D receptor gene polymorphisms, and severity of Parkinson’s disease. Mov Disord. 2012;27:264–71.

    Article  CAS  PubMed  Google Scholar 

  54. Peterson AL. A review of vitamin D and Parkinson's disease. Maturitas. 2014;78:40–4.

    Article  CAS  PubMed  Google Scholar 

  55. 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;25:666–74.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Liu Y, Zhang BS. Serum 25-hydroxyvitamin D predicts severity in Parkinson’s disease patients. Neurol Sci. 2014;35:67–71.

    Article  PubMed  Google Scholar 

  57. Luo X, Ou R, Dutta R, Tian Y, Xiong H, Shang H. Association between serum vitamin D levels and Parkinson’s disease: a systematic review and meta-analysis. Front Neurol. 2018;9:909.

    Article  PubMed  PubMed Central  Google Scholar 

  58. Mrabet S, Ben Ali N, Achouri A, Dabbeche R, Najjar T, Haouet S, et al. Gastrointestinal dysfunction and neuropathologic correlations in Parkinson disease. J Clin Gastroenterol. 2016;50:e85–90.

    Article  CAS  PubMed  Google Scholar 

  59. Wang L, Evatt ML, Maldonado LG, Perry WR, Ritchie JC, Beecham GW, et al. Vitamin D from different sources is inversely associated with Parkinson disease. Mov Disord. 2015;30:560–6.

    Article  CAS  PubMed  Google Scholar 

  60. Evatt ML, DeLong MR, Kumari M, Auinger P, McDermott MP, Tangpricha V, et al. High prevalence of hypovitaminosis D status in patients with early Parkinson disease. Arch Neurol. 2011;68:314–9.

    Article  PubMed  Google Scholar 

  61. Petersen MS, Bech S, Christiansen DH, Schmedes AV, Halling J. The role of vitamin D levels and vitamin D receptor polymorphism on Parkinson’s disease in the Faroe Islands. Neurosci Lett. 2014;561:74–9.

    Article  CAS  PubMed  Google Scholar 

  62. Knekt P, Kilkkinen A, Rissanen H, Marniemi J, Saaksjarvi K, Heliovaara M. Serum vitamin D and the risk of Parkinson disease. Arch Neurol. 2010;67:808–11.

    Article  PubMed  PubMed Central  Google Scholar 

  63. 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.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  64. Zhu D, Liu GY, Lv Z, Wen SR, Bi S, Wang WZ. Inverse associations of outdoor activity and vitamin D intake with the risk of Parkinson’s disease. J Zhejiang Univ Sci B. 2014;15:923–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Kenborg L, Lassen CF, Ritz B, Schernhammer ES, Hansen J, Gatto NM, et al. Outdoor work and risk for Parkinson’s disease: a population-based case-control study. Occup Environ Med. 2011;68:273–8.

    Article  PubMed  Google Scholar 

  66. Kravietz A, Kab S, Wald L, Dugravot A, Singh-Manoux A, Moisan F, et al. Association of UV radiation with Parkinson disease incidence: a nationwide French ecologic study. Environ Res. 2017;154:50–6.

    Article  CAS  PubMed  Google Scholar 

  67. Shrestha S, Lutsey PL, Alonso A, Huang X, Mosley TH Jr, Chen H. Serum 25-hydroxyvitamin D concentrations in mid-adulthood and Parkinson's disease risk. Mov Disord. 2016;31:972–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Larsson SC, Singleton AB, Nalls MA, Richards JB, International Parkinson’s Disease Genomics C. No clear support for a role for vitamin D in Parkinson’s disease: A Mendelian randomization study. Mov Disord. 2017;32:1249–52.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Scherzer CR, Eklund AC, Morse LJ, Liao Z, Locascio JJ, Fefer D, et al. Molecular markers of early Parkinson’s disease based on gene expression in blood. Proc Natl Acad Sci U S A. 2007;104:955–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Uitterlinden AG, Fang Y, Van Meurs JB, Pols HA, Van Leeuwen JP. Genetics and biology of vitamin D receptor polymorphisms. Gene. 2004;338:143–56.

    Article  CAS  PubMed  Google Scholar 

  71. Uitterlinden AG, Fang Y, van Meurs JB, van Leeuwen H, Pols HA. Vitamin D receptor gene polymorphisms in relation to Vitamin D related disease states. J Steroid Biochem Mol Biol. 2004;89–90:187–93.

    Article  PubMed  CAS  Google Scholar 

  72. Jurutka PW, Whitfield GK, Hsieh JC, Thompson PD, Haussler CA, Haussler MR. Molecular nature of the vitamin D receptor and its role in regulation of gene expression. Rev Endocr Metab Disord. 2001;2:203–16.

    Article  CAS  PubMed  Google Scholar 

  73. Kim JS, Kim YI, Song C, Yoon I, Park JW, Choi YB, et al. Association of vitamin D receptor gene polymorphism and Parkinson’s disease in Koreans. J Korean Med Sci. 2005;20:495–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Torok R, Torok N, Szalardy L, Plangar I, Szolnoki Z, Somogyvari F, et al. Association of vitamin D receptor gene polymorphisms and Parkinson's disease in Hungarians. Neurosci Lett. 2013;551:70–4.

    Article  CAS  PubMed  Google Scholar 

  75. Niu MY, Wang L, Xie AM. ApaI, BsmI, FokI, and TaqI polymorphisms in the vitamin D receptor gene and Parkinson’s disease. Chin Med J (Engl). 2015;128:1809–14.

    Article  CAS  Google Scholar 

  76. 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.

    Article  CAS  PubMed  Google Scholar 

  77. Tanaka K, Miyake Y, Fukushima W, Kiyohara C, Sasaki S, Tsuboi Y, et al. Vitamin D receptor gene polymorphisms, smoking, and risk of sporadic Parkinson’s disease in Japan. Neurosci Lett. 2017;643:97–102.

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  79. Gatto NM, Paul KC, Sinsheimer JS, Bronstein JM, Bordelon Y, Rausch R, et al. Vitamin D receptor gene polymorphisms and cognitive decline in Parkinson's disease. J Neurol Sci. 2016;370:100–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Gatto NM, Sinsheimer JS, Cockburn M, Escobedo LA, 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.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Kang SY, Park S, Oh E, Park J, Youn J, Kim JS, et al. Vitamin D receptor polymorphisms and Parkinson’s disease in a Korean population: revisited. Neurosci Lett. 2016;628:230–5.

    Article  CAS  PubMed  Google Scholar 

  82. Invernizzi M, Carda S, Viscontini GS, Cisari C. Osteoporosis in Parkinson’s disease. Parkinsonism Relat Disord. 2009;15:339–46.

    Article  PubMed  Google Scholar 

  83. Lee CK, Choi SK, Shin DA, Yi S, Kim KN, Kim I, et al. Parkinson’s disease and the risk of osteoporotic vertebral compression fracture: a nationwide population-based study. Osteoporos Int. 2018;29:1117–24.

    Article  CAS  PubMed  Google Scholar 

  84. Vaserman N. Parkinson’s disease and osteoporosis. Joint Bone Spine. 2005;72:484–8.

    Article  PubMed  Google Scholar 

  85. Bezza A, Ouzzif Z, Naji H, Achemlal L, Mounach A, Nouijai M, et al. Prevalence and risk factors of osteoporosis in patients with Parkinson’s disease. Rheumatol Int. 2008;28:1205–9.

    Article  CAS  PubMed  Google Scholar 

  86. Park SB, Chung CK, Lee JY, Lee JY, Kim J. Risk factors for vertebral, hip, and femoral fractures among patients with Parkinson’s disease: a 5-year follow-up in Korea. J Am Med Dir Assoc. 2019;20:617–23.

    Article  PubMed  Google Scholar 

  87. Metta V, Sanchez TC, Padmakumar C. Osteoporosis: a hidden nonmotor face of Parkinson’s disease. Int Rev Neurobiol. 2017;134:877–90.

    Article  PubMed  Google Scholar 

  88. Sato Y, Kaji M, Tsuru T, Oizumi K. Risk factors for hip fracture among elderly patients with Parkinson’s disease. J Neurol Sci. 2001;182:89–93.

    Article  CAS  PubMed  Google Scholar 

  89. van den Bos F, Speelman AD, Samson M, Munneke M, Bloem BR, Verhaar HJ. Parkinson’s disease and osteoporosis. Age Ageing. 2013;42:156–62.

    Article  PubMed  Google Scholar 

  90. Sato Y, Kikuyama M, Oizumi K. High prevalence of vitamin D deficiency and reduced bone mass in Parkinson’s disease. Neurology. 1997;49:1273–8.

    Article  CAS  PubMed  Google Scholar 

  91. Lyell V, Henderson E, Devine M, Gregson C. Assessment and management of fracture risk in patients with Parkinson’s disease. Age Ageing. 2015;44:34–41.

    Article  PubMed  Google Scholar 

  92. Binks S, Dobson R. Risk factors, epidemiology and treatment strategies for metabolic bone disease in patients with neurological disease. Curr Osteoporos Rep. 2016;14:199–210.

    Article  CAS  PubMed  Google Scholar 

  93. Ozturk EA, Gundogdu I, Tonuk B, Kocer BG, Tombak Y, Comoglu S, et al. Bone mass and vitamin D levels in Parkinson’s disease- is there any difference between genders. J Phys Ther Sci. 2016;28:2204–9.

    Article  PubMed  PubMed Central  Google Scholar 

  94. Malochet-Guinamand S, Durif F, Thomas T. Parkinson’s disease: a risk factor for osteoporosis. Joint Bone Spine. 2015;82:406–10.

    Article  PubMed  Google Scholar 

  95. Bischoff-Ferrari HA, Dawson-Hughes B, Staehelin HB, Orav JE, Stuck AE, Theiler R, et al. Fall prevention with supplemental and active forms of vitamin D: a meta-analysis of randomised controlled trials. BMJ. 2009;339:b3692.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Chitsaz A, Maracy M, Basiri K, Izadi Boroujeni M, Tanhaei AP, Rahimi M, et al. 25-hydroxyvitamin d and severity of Parkinson’s disease. Int J Endocrinol. 2013;2013:689149.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  97. Sleeman I, Aspray T, Lawson R, Coleman S, Duncan G, Khoo TK, et al. The role of vitamin D in disease progression in early Parkinson’s disease. J Parkinsons Dis. 2017;7:669–75.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Peterson AL, Mancini M, Horak FB. The relationship between balance control and vitamin D in Parkinson’s disease-a pilot study. Mov Disord. 2013;28:1133–7.

    Article  PubMed  Google Scholar 

  99. Feart C, Helmer C, Merle B, Herrmann FR, Annweiler C, Dartigues JF, et al. Associations of lower vitamin D concentrations with cognitive decline and long-term risk of dementia and Alzheimer’s disease in older adults. Alzheimers Dement. 2017;13:1207–16.

    Article  PubMed  Google Scholar 

  100. Peterson AL, Murchison C, Zabetian C, Leverenz JB, Watson GS, Montine T, et al. Memory, mood, and vitamin D in persons with Parkinson’s disease. J Parkinsons Dis. 2013;3:547–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Hooshmand B, Lokk J, Solomon A, Mangialasche F, Miralbell J, Spulber G, et al. Vitamin D in relation to cognitive impairment, cerebrospinal fluid biomarkers, and brain volumes. J Gerontol A Biol Sci Med Sci. 2014;69:1132–8.

    Article  CAS  PubMed  Google Scholar 

  102. Kalaitzakis ME, Graeber MB, Gentleman SM, Pearce RK. Striatal amyloid deposition in Parkinson disease with dementia. J Neuropathol Exp Neurol. 2008;67:155–61.

    Article  PubMed  Google Scholar 

  103. Grimm MOW, Thiel A, Lauer AA, Winkler J, Lehmann J, Regner L, et al. Vitamin D and its analogues decrease amyloid-beta (Abeta) formation and increase Abeta-degradation. Int J Mol Sci. 2017;18:1-21.

  104. Jia J, Hu J, Huo X, Miao R, Zhang Y, Ma F. Effects of vitamin D supplementation on cognitive function and blood Abeta-related biomarkers in older adults with Alzheimer’s disease: a randomised, double-blind, placebo-controlled trial. J Neurol Neurosurg Psychiatry. 2019;90:1347–52.

    PubMed  Google Scholar 

  105. Irwin DJ, Lee VM, Trojanowski JQ. Parkinson’s disease dementia: convergence of alpha-synuclein, tau and amyloid-beta pathologies. Nat Rev Neurosci. 2013;14:626–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Phillipson OT. Alpha-synuclein, epigenetics, mitochondria, metabolism, calcium traffic, & circadian dysfunction in Parkinson’s disease. An integrated strategy for management. Ageing Res Rev. 2017;40:149–67.

    Article  CAS  PubMed  Google Scholar 

  107. Latimer CS, Brewer LD, Searcy JL, Chen KC, Popovic J, Kraner SD, et al. Vitamin D prevents cognitive decline and enhances hippocampal synaptic function in aging rats. Proc Natl Acad Sci U S A. 2014;111:E4359–66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Lee EY, Eslinger PJ, Du G, Kong L, Lewis MM, Huang X. Olfactory-related cortical atrophy is associated with olfactory dysfunction in Parkinson’s disease. Mov Disord. 2014;29:1205–8.

    Article  PubMed  PubMed Central  Google Scholar 

  109. Baba T, Kikuchi A, Hirayama K, Nishio Y, Hosokai Y, Kanno S, et al. Severe olfactory dysfunction is a prodromal symptom of dementia associated with Parkinson’s disease: a 3 year longitudinal study. Brain. 2012;135:161–9.

    Article  PubMed  Google Scholar 

  110. Takeda A, Baba T, Kikuchi A, Hasegawa T, Sugeno N, Konno M, et al. Olfactory dysfunction and dementia in Parkinson’s disease. J Parkinsons Dis. 2014;4:181–7.

    Article  PubMed  Google Scholar 

  111. Kim JE, Oh E, Park J, Youn J, Kim JS, Jang W. Serum 25-hydroxyvitamin D3 level may be associated with olfactory dysfunction in de novo Parkinson’s disease. J Clin Neurosci. 2018;57:131–5.

    Article  CAS  PubMed  Google Scholar 

  112. Braak H, Del Tredici K, Rüb U, de Vos RA, Jansen Steur EN, Braak E. Staging of brain pathology related to sporadic Parkinson’s disease. Neurobiol Aging. 2003;24:197–211.

    Article  PubMed  Google Scholar 

  113. Kwon KY, Jo KD, Lee MK, Oh M, Kim EN, Park J, et al. Low serum vitamin D levels may contribute to gastric dysmotility in de novo Parkinson’s disease. Neurodegener Dis. 2016;16:199–205.

    Article  CAS  PubMed  Google Scholar 

  114. Jang W, Park J, Kim JS, Youn J, Oh E, Kwon KY, et al. Vitamin D deficiency in Parkinson’s disease patients with orthostatic hypotension. Acta Neurol Scand. 2015;132:242–50.

    Article  CAS  PubMed  Google Scholar 

  115. Pertile RAN, Cui X, Hammond L, Eyles DW. Vitamin D regulation of GDNF/ret signaling in dopaminergic neurons. FASEB J. 2018;32:819–28.

    Article  CAS  PubMed  Google Scholar 

  116. Cui X, Pelekanos M, Burne TH, McGrath JJ, Eyles DW. Maternal vitamin D deficiency alters the expression of genes involved in dopamine specification in the developing rat mesencephalon. Neurosci Lett. 2010;486:220–3.

    Article  CAS  PubMed  Google Scholar 

  117. Vinh Quoc Luong K, Thi Hoang Nguyen L. Vitamin D and Parkinson’s disease. J Neurosci Res. 2012;90:2227–36.

    Article  PubMed  CAS  Google Scholar 

  118. Claro da Silva T, Hiller C, Gai Z, Kullak-Ublick GA. Vitamin D3 transactivates the zinc and manganese transporter SLC30A10 via the vitamin D receptor. J Steroid Biochem Mol Biol. 2016;163:77–87.

    Article  PubMed  CAS  Google Scholar 

  119. Jiang P, Zhang LH, Cai HL, Li HD, Liu YP, Tang MM, et al. Neurochemical effects of chronic administration of calcitriol in rats. Nutrients. 2014;6:6048–59.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  120. Ebert R, Schutze N, Adamski J, Jakob F. Vitamin D signaling is modulated on multiple levels in health and disease. Mol Cell Endocrinol. 2006;248:149–59.

    Article  CAS  PubMed  Google Scholar 

  121. Airavaara M, Voutilainen MH, Wang Y, Hoffer B. Neurorestoration. Park Relat Disord. 2012;18:S143–6.

    Article  Google Scholar 

  122. Weissmiller AM, Wu C. Current advances in using neurotrophic factors to treat neurodegenerative disorders. Transl Neurodegener. 2012;1:14.

    Article  PubMed  PubMed Central  Google Scholar 

  123. Eyles D, Brown J, Mackay-Sim A, McGrath J, Feron F. Vitamin d3 and brain development. Neuroscience. 2003;118:641–53.

    Article  CAS  PubMed  Google Scholar 

  124. McGrath JJ, Feron FP, Burne TH, Mackay-Sim A, Eyles DW. Vitamin D3-implications for brain development. J Steroid Biochem Mol Biol. 2004;89–90:557–60.

    Article  PubMed  CAS  Google Scholar 

  125. Sariola H, Saarma M. Novel functions and signalling pathways for GDNF. J Cell Sci. 2003;116:3855–62.

    Article  CAS  PubMed  Google Scholar 

  126. García-Martínez JM, Pérez-Navarro E, Gavaldà N, Alberch J. Glial cell line-derived neurotrophic factor promotes the arborization of cultured striatal neurons through the p42/p44 mitogen-activated protein kinase pathway. J Neurosci Res. 2006;83:68–79.

    Article  PubMed  CAS  Google Scholar 

  127. Chao CC, Lee EH. Neuroprotective mechanism of glial cell line-derived neurotrophic factor on dopamine neurons: role of antioxidation. Neuropharmacology. 1999;38:913–6.

    Article  CAS  PubMed  Google Scholar 

  128. Garcion E, Sindji L, Leblondel G, Brachet P, Darcy F. 1,25-Dihydroxyvitamin D3 regulates the synthesis of γ-glutamyl transpeptidase and glutathione levels in rat primary astrocytes. J Neurochem. 1999;73:859–66.

    Article  CAS  PubMed  Google Scholar 

  129. Berridge MJ. Vitamin D cell signalling in health and disease. Biochem Biophys Res Commun. 2015;460:53–71.

    Article  CAS  PubMed  Google Scholar 

  130. Wiseman H. Vitamin D is a membrane antioxidant. Ability to inhibit iron-dependent lipid peroxidation in liposomes compared to cholesterol, ergosterol and tamoxifen and relevance to anticancer action. FEBS Lett. 1993;326:285–8.

    Article  CAS  PubMed  Google Scholar 

  131. Lin AM, Chen KB, Chao PL. Antioxidative effect of vitamin D3 on zinc-induced oxidative stress in CNS. Ann N Y Acad Sci. 2005;1053:319–29.

    Article  CAS  PubMed  Google Scholar 

  132. Calvello R, Cianciulli A, Nicolardi G, De Nuccio F, Giannotti L, Salvatore R, et al. 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;12:327–39.

    Article  PubMed  Google Scholar 

  133. Lima LAR, Lopes MJP, Costa RO, Lima FAV, Neves KRT, Calou IBF, et al. Vitamin D protects dopaminergic neurons against neuroinflammation and oxidative stress in hemiparkinsonian rats. J Neuroinflammation. 2018;15:249.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  134. Sanchez B, Relova JL, 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;87:723–32.

    Article  CAS  PubMed  Google Scholar 

  135. Rcom-H’cheo-Gauthier AN, Meedeniya AC, Pountney DL. Calcipotriol inhibits alpha-synuclein aggregation in SH-SY5Y neuroblastoma cells by a Calbindin-D28k-dependent mechanism. J Neurochem. 2017;141:263–74.

    Article  PubMed  CAS  Google Scholar 

  136. McLeary FA, Rcom-H’cheo-Gauthier AN, Goulding M, Radford RAW, Okita Y, Faller P, et al. Switching on endogenous metal binding proteins in Parkinson's disease. Cells. 2019;8:179.

    Article  CAS  PubMed Central  Google Scholar 

  137. Rcom-H'cheo-Gauthier AN, Osborne SL, Meedeniya AC, Pountney DL. Calcium: alpha-synuclein interactions in alpha-synucleinopathies. Front Neurosci. 2016;10:570.

    Article  PubMed  PubMed Central  Google Scholar 

  138. Santner A, Uversky VN. Metalloproteomics and metal toxicology of alpha-synuclein. Metallomics. 2010;2:378–92.

    Article  CAS  PubMed  Google Scholar 

  139. Pertile RA, Cui X, Eyles DW. Vitamin D signaling and the differentiation of developing dopamine systems. Neuroscience. 2016;333:193–203.

    Article  CAS  PubMed  Google Scholar 

  140. Li H, Jang W, Kim HJ, Jo KD, Lee MK, Song SH, et al. Biochemical protective effect of 1,25-dihydroxyvitamin D3 through autophagy induction in the MPTP mouse model of Parkinson’s disease. Neuroreport. 2015;26:669–74.

    Article  CAS  PubMed  Google Scholar 

  141. Cui X, Pertile R, Liu P, Eyles DW. Vitamin D regulates tyrosine hydroxylase expression: N-cadherin a possible mediator. Neuroscience. 2015;304:90–100.

    Article  CAS  PubMed  Google Scholar 

  142. Sequeira VB, Rybchyn MS, Tongkao-On W, Gordon-Thomson C, Malloy PJ, Nemere I, et al. The role of the vitamin D receptor and ERp57 in photoprotection by 1alpha,25-dihydroxyvitamin D3. Mol Endocrinol. 2012;26:574–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  143. Landel V, Stephan D, Cui X, Eyles D, Feron F. Differential expression of vitamin D-associated enzymes and receptors in brain cell subtypes. J Steroid Biochem Mol Biol. 2018;177:129–34.

    Article  CAS  PubMed  Google Scholar 

  144. Aureli C, Cassano T, Masci A, Francioso A, Martire S, Cocciolo A, et al. 5-S-cysteinyldopamine neurotoxicity: influence on the expression of alpha-synuclein and ERp57 in cellular and animal models of Parkinson's disease. J Neurosci Res. 2014;92:347–58.

    Article  CAS  PubMed  Google Scholar 

  145. Kuang XL, Liu F, Chen H, Li Y, Liu Y, Xiao J, et al. Reductions of the components of the calreticulin/calnexin quality-control system by proteasome inhibitors and their relevance in a rodent model of Parkinson’s disease. J Neurosci Res. 2014;92:1319–29.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Kim-Han JS, O’Malley KL. Cell stress induced by the parkinsonian mimetic, 6-hydroxydopamine, is concurrent with oxidation of the chaperone, ERp57, and aggresome formation. Antioxid Redox Signal. 2007;9:2255–64.

    Article  CAS  PubMed  Google Scholar 

  147. Grillo C, D’Ambrosio C, Scaloni A, Maceroni M, Merluzzi S, Turano C, et al. Cooperative activity of Ref-1/APE and ERp57 in reductive activation of transcription factors. Free Radic Biol Med. 2006;41:1113–23.

    Article  CAS  PubMed  Google Scholar 

  148. Hiller AL, Murchison CF, Lobb BM, O’Connor S, O'Connor M, Quinn JF. 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.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  149. Muir SW, Montero-Odasso M. Effect of vitamin D supplementation on muscle strength, gait and balance in older adults: a systematic review and meta-analysis. J Am Geriatr Soc. 2011;59:2291–300.

    Article  PubMed  Google Scholar 

  150. Asaka T, Yahata K, Mani H, Wang Y. Modulations of muscle modes in automatic postural responses induced by external surface translations. J Mot Behav. 2011;43:165–72.

    Article  PubMed  Google Scholar 

  151. Torres-Oviedo G, Ting LH. Muscle synergies characterizing human postural responses. J Neurophysiol. 2007;98:2144–56.

    Article  PubMed  Google Scholar 

  152. Lawton M, Baig F, Toulson G, Morovat A, Evetts SG, Ben-Shlomo Y, et al. Blood biomarkers with Parkinson’s disease clusters and prognosis: the oxford discovery cohort. Mov Disord. 2020;35:279–87.

    Article  CAS  PubMed  Google Scholar 

  153. Mark KA, Dumas KJ, Bhaumik D, Schilling B, Davis S, Oron TR, 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;17:1227–37.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  154. Vieth R. Vitamin D supplementation, 25-hydroxyvitamin D concentrations, and safety. Am J Clin Nutr. 1999;69:842–56.

    Article  CAS  PubMed  Google Scholar 

  155. Dudenkov DV, Yawn BP, Oberhelman SS, Fischer PR, Singh RJ, Cha SS, et al. Changing incidence of serum 25-hydroxyvitamin D values above 50 ng/mL: a 10-year population-based study. Mayo Clin Proc. 2015;90:577–86.

    Article  CAS  PubMed  Google Scholar 

  156. Ross AC, Manson JE, Abrams SA, Aloia JF, Brannon PM, Clinton SK, et al. The 2011 report on dietary reference intakes for calcium and vitamin D from the Institute of Medicine: what clinicians need to know. J Clin Endocrinol Metab. 2011;96:53–8.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We are grateful for the support from the National Natural Science Foundation of China (81430025 and U1801681) and the Key Field Research Development Program of Guangdong Province (2018B030337001).

Funding

The authors of this review were supported by the National Natural Science Foundation of China (81971201) and the National Science Foundation of Hunan Province (2019JJ40450).

Author information

Authors and Affiliations

Authors

Contributions

LLL reviewed the literature, drafted and revised the manuscript; WCY and TJQ gave suggestions for the article. All other authors critically revised the manuscript. All authors approved the final manuscript.

Corresponding author

Correspondence to Chunyu Wang.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lv, L., Tan, X., Peng, X. et al. The relationships of vitamin D, vitamin D receptor gene polymorphisms, and vitamin D supplementation with Parkinson’s disease. Transl Neurodegener 9, 34 (2020). https://doi.org/10.1186/s40035-020-00213-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40035-020-00213-2

Keywords