Sei sulla pagina 1di 19

REVIEWS

New aspects of vitamin D metabolism


and action — addressing the skin as
source and target
Daniel Bikle   1,2* and Sylvia Christakos3
Abstract | Vitamin D has a key role in stimulating calcium absorption from the gut and promoting
skeletal health, as well as many other important physiological functions. Vitamin D is produced in
the skin. It is subsequently metabolized to its hormonally active form, 1,25-dihydroxyvitamin D
(1,25(OH)2D), by the 1-hydroxylase and catabolized by the 24-hydroxylase. In this Review , we pay
special attention to the effect of mutations in these enzymes and their clinical manifestations. We
then discuss the role of vitamin D binding protein in transporting vitamin D and its metabolites from
their source to their targets, the free hormone hypothesis for cell entry and HSP70 for intracellular
transport. This is followed by discussion of the vitamin D receptor (VDR) that mediates the cellular
actions of 1,25(OH)2D. Cell-specific recruitment of co-regulatory complexes by liganded VDR
leads to changes in gene expression that result in distinct physiological actions by 1,25(OH)2D,
which are disrupted by mutations in the VDR . We then discuss the epidermis and hair follicle, to
provide a non-skeletal example of a tissue that expresses VDR that not only makes vitamin D but
also can metabolize it to its hormonally active form. This enables vitamin D to regulate epidermal
differentiation and hair follicle cycling and, in so doing, to promote barrier function, wound healing
and hair growth, while limiting cancer development.

Vitamin D, a fat-soluble secosteroid, comes in two forms: The Kandutsch–Russell pathway is an alternative path-
vitamin D2 and vitamin D3. These forms differ only in the way to the Bloch pathway for cholesterol synthesis that
side chain, where vitamin D2 has a double bond between directs zymosterol to 7-DHC, which can be converted
C22 and C23, and a methyl group at C24, unlike vitamin by UVB radiation to vitamin D, rather than to des-
D3. Vitamin D3 is produced in the skin from 7-dehydro- mosterol, which cannot. Conversion to desmosterol is
cholesterol (7-DHC)1, whereas vitamin D2 is produced in achieved by reversing the order in which 24-dehydro-
plants and fungi from ergosterol2. The biological activi- cholesterol reductase (DHCR24) and DHCR7 act at the
ties are comparable, although there are subtle differences different stages of cholesterol synthesis. The Bloch path-
in pharmacokinetics3; unless stated otherwise in the text, way is primarily utilized by steroidogenic tissues, liver
vitamin D with no subscript refers to both forms. and kidney, whereas the Kandutsch–Russell pathway is
1
Departments of Medicine Vitamin D3 is produced in the skin via a two-step pro- primarily utilized by skin, brain and muscle6. Therefore,
and Dermatology, University cess under the influence of ultraviolet B (UVB) radia- DHCR7 is really the critical first step in the production of
of California San Francisco,
tion (290–315 nm) from the vitamin D substrate 7-DHC. vitamin D, and, as discussed below, its regulation impacts
San Francisco, CA, USA.
The synthesis begins with the conversion of 7-DHC to the amount of 7-DHC available for photoproduction of
2
VA Medical Center,
San Francisco, CA, USA.
pre-vitamin D, which then undergoes thermal conversion vitamin D in the skin.
to vitamin D4. Clothing and deep pigmentation of skin Vitamin D has limited biological activity, although it
3
Departments of
Microbiology, Biochemistry limit epidermal vitamin D production, but, according to does inhibit Smoothened, which is a key component of
and Molecular Genetics, a 2019 review, sunscreens do not limit production very the Hedgehog pathway and might be part of the means
New Jersey Medical School, much5. The conversion of 7-DHC to pre-vitamin D and by which vitamin D signalling protects the skin from the
Rutgers, the State University subsequent conversion of pre-vitamin D to vitamin D carcinogenic effect of UVB radiation7. For most vitamin
of New Jersey, Newark,
NJ, USA.
are non-enzymatic steps. The production 7-DHC, how- D-mediated actions, vitamin D must first be metabolized
ever, is enzymatic. 7-DHC is produced in the skin via the to 25-hydroxyvitamin D (25OHD), a process that occurs
*e-mail: Daniel.bikle@
ucsf.edu Kandutsch–Russell pathway. This pathway results in primarily, but not exclusively, in the liver8. 25OHD is the
https://doi.org/10.1038/ the production of either cholesterol by the enzyme 7-DHC major circulating metabolite of vitamin D9. Numerous
s41574-019-0312-5 reductase (DHCR7) or vitamin D by UVB radiation6. enzymes have 25-hydroxylase activity, but the major

Nature Reviews | Endocrinology


Reviews

Key points total and free 25OHD19. However, with the exception of
cells (such as renal cells) that express the megalin–cubilin
• Mutations in the enzymes involved in vitamin D metabolism cause human diseases complex, most cells are thought to take up 25OHD and
such as rickets. 1,25(OH)2D in the non-protein-bound form, a concept
• Vitamin D binding protein is the major transport protein for vitamin D metabolites, termed the free hormone hypothesis20. This hypothesis
but with the exception of a few tissues such as the kidney, it is the free forms of these has generated a discussion as to whether free 25OHD
metabolites that enter cells. might provide a better measure of vitamin D status than
• The vitamin D receptor (VDR) is the major mediator of vitamin D biological action. total 25OHD. In this regard, studies in mice lacking
• VDR binding sites can be located in a range of locations including introns and at distal DBP and a case report of a patient lacking DBP indicate
intergenic regions of regulated genes. that loss of DBP does not result in vitamin D deficiency,
• Co-regulators of VDR provide cell-specific genomic regulation. even in the presence of undetectable 25OHD levels
• The skin is an excellent non-skeletal model for the study of vitamin D as it is the source unless intake of vitamin D is restricted21,22.
of vitamin D, contains the vitamin D-metabolizing enzymes and expresses the VDR, Within the cells that have been studied, HSP70 has
making it a target tissue as well. been identified as a potential transporter of 25OHD
through the plasma membrane to the mitochon-
enzyme responsible for 25OHD production under phys- dria, where it is converted to 1,25(OH)2D23. From the
iological circumstances is CYP2R1. Knockout of Cyp2r1 mitochondria, HSP70 is then thought to transport
in mice results in a reduction in blood levels of 25OHD by 1,25(OH)2D into the nucleus where it binds to the vita-
>50%, but not to zero10. Moreover, mutations in CYP2R1 min D receptor (VDR)23. HSP70 has a higher affinity
in humans leads to rickets11. Although 25OHD is the for 25OHD than DBP, which facilities 25OHD trans-
major circulating form of vitamin D, its active hormo- port across the cell membrane, but lower affinity for
nal form, which is synthesized by the enzyme CYP27B1, 1,25(OH)2D than VDR, which facilities 1,25(OH)2D
is 1,25(OH)2D. offloading onto VDR in the nucleus23.
Renal cells are the major source of circulating VDR also mediates the genomic, and some of the
1,25(OH)2D, but many other cells, including keratino- non-genomic actions, of 1,25(OH)2D24,25. In addition,
cytes and immune cells, express CYP27B1. The regula- VDR has actions that are independent of 1,25(OH)2D,
tion of the production of 1,25(OH)2D in non-renal cells including the regulation of hair follicle cycling 26.
differs from that in renal cells, leading to the concept that Furthermore, each cell has its own repertoire of responses
non-renal cells are capable of producing 1,25(OH)2D to 1,25(OH)2D, indicating that within different cells are
for their own needs in an autocrine and/or paracrine factors, such as co-regulators and other transcription fac-
fashion12. If the activity of CYP27B1 is unchecked, the tors, that influence VDR function in a cell-specific man-
result can be the over-production of 1,25(OH)2D, which ner. One excellent example of the cell-specific actions of
can result in hypercalcaemia and/or hypercalciuria12. 1,25(OH)2D and/or VDR is the regulation of Cyp27b1 in
In the kidney, CYP27B1 activity is stimulated by para- renal cells in contrast to the lack of regualtion of Cyp27b1
thyroid hormone (PTH) and inhibited by the action of activity by 1,25(OH)2D/VDR in macrophages27. In par-
fibroblast growth factor 23 (FGF23) and 1,25(OH)2D. ticular, the regulatory region of Cyp27b1 in mouse renal
In other tissues, the regulation of CYP27B1 occurs cells that responds to 1,25(OH)2D is not responsive to
primarily by cytokines such as tumour necrosis factor 1,25(OH)2D in macrophages and most other non-renal
and interferon-γ12. 1,25(OH)2D indirectly regulates its cells27. Over the past 5 years, genetic and epigenetic
own levels in these cells via CYP24A1, a 24-hydroxy- mechanisms have been discovered that underlie the
lase. CYP24A1, which is found in most tissues, is key tissue-specific actions of VDR and its ligand. The differ-
to the catabolism of both 25OHD and 1,25(OH)2D. As ent regulation of CYP27B1 and CYP24A1 by hormones
the production of CYP24A1 is induced by 1,25(OH)2D, and cytokines between renal cells and non-renal cells
CYP24A1 serves as the brake on levels of 1,25(OH)2D. is a good example of this that is relevant to vitamin D
Not surprisingly, mutations in CYP27B1 and CYP24A1 metabolism27,28, as is the different regulation of receptor
cause profound disturbances. Inactivating mutations in activator of nuclear factor-κΒ ligand (RANKL) by hor-
CYP2R1 and CYP27B1 lead to rickets13,14, and inactivat- mones and cytokines between osteoblasts and T cells29.
ing mutations in CYP24A1 are one cause of idiopathic Furthermore, structural analyses have examined how
infantile hypercalcaemia (IIH) in children and recurrent mutations in the VDR can affect its function, including
kidney stone formation in adults15. its interactions with the various co-regulators, with some
Vitamin D and its metabolites are generally trans- mutations in the zinc finger and hinge region interfering
ported from the site of production to their target tis- with retinoid X receptor (RXR) binding and others in
sues. Vitamin D binding protein (DBP), which binds the AP2 domain interfering with co-activator binding30.
to ~85% of the circulating 25OHD and 1,25(OH)2D, The skin is the only organ that has the capacity to
is primarily responsible for such transport16,17. DBP is make vitamin D, metabolize it to 1,25(OH)2D, and res­
a highly polymorphic protein, and polymorphisms pond to it in a cell-specific fashion31. Psoriasis, a prolif-
can affect its functions, including its ability to bind to erative inflammatory disease of the skin, is a pro­minent
25OHD18. The finding that polymorphisms in DBP can example of a non-skeletal disease in which 1,25(OH)2D
alter its function has sparked new interest in defining and its analogues have proved to be effective ther-
the poly­morphisms, the effect they have on circulating apy32. This finding illustrates both the antiproliferative
levels of vitamin D and the response of individuals to and anti-inflammatory actions of vitamin D. The cells
vitamin D supplementation with respect to increases in responsible for these responses to 1,25(OH)2D in the

www.nature.com/nrendo
Reviews

skin are keratinocytes, the dominant cell type of both impact of mutations in these enzymes. Figure 1 shows the
the epidermis and hair follicles1. As such, the skin pro- overall pathway that is discussed, and Box 1 summarizes
vides an excellent model tissue in which to study both the mutations and their clinical features.
vitamin D metabolism and its mechanisms of action,
as the regulation of metabolism in keratinocytes differs DHCR7. As noted in the introduction, DHCR7 regu-
from that in the kidney31. For these reasons, the free lates the levels of 7-DHC in the skin. Its activity con-
hormone hypothesis can be directly tested in the skin. verts 7-DHC to cholesterol. When DHCR7 is inhibited
Furthermore, in the skin the cell-specific actions of VDR (or inactivated by mutations), more 7-DHC is available
and/or 1,25(OH)2D are clearly illustrated by the differ- for vitamin D production. In addition to the skin,
ent molecular events that are regulated by VDR and/or this enzyme is found in the brain, muscle and heart33.
1,25(OH)2D during the temporal and sequential events Cholesterol and vitamin D (but not 1,25(OH)2D) incre­
of epidermal differentiation and hair follicle cycling1. ase proteasomal degradation of DHCR7, leading to
In this Review, we discuss what we have learned from increased vitamin D production34. AMP-activated pro-
the new discoveries in vitamin D metabolism. We focus tein kinase, a key sensor and regulator of cellular energy
on mutations in the enzymes that metabolize vitamin D, homeostasis, and protein kinase A, are potent inhibitors
DBP and its role in regulating tissue access to the of DHCR7 (refs6,35).
vitamin D metabolites it transports, and the mechanisms Inactivating mutations of DHCR7 result in the
that underlie the tissue-specific actions of VDR. In the Smith–Lemli–Opitz syndrome, a developmental dis­
final section we use the skin to illustrate a number of order in which patients can present with mental
these points, with a focus on the diversity of vitamin D retardation, growth deficiency and congenital malfor-
actions within the vitamin D endocrine system, and mations36,37. These patients are affected primarily by the
how this diversity is utilized in the regulation of three consequences of too little cholesterol, steroids or bile
important functions of the skin; namely, protection acids. However, they seem to be more sensitive to UVB
from the environment, promotion of wound healing and radiation, and can present with increased serum 25OHD
suppression of cancer development. concentrations 38, although not with toxic levels 38.
Polymorphisms in DHCR7 have been associated with
Vitamin D production and metabolism either reduced39,40 or increased41 25OHD levels. Whether
In this section, we discuss the key enzymes involved in these polymorphisms, all of which are in introns, affect
vitamin D production and metabolism including the enzymatic function is not clear.

↑ P and Ca OH
OH
↑ 1,25(OH)2D
↑ FGF23
↓ PTH

CH2
OH CYP24A1
HO
24,25(OH)2D3
Liver Kidney

CYP2R1
CH2 CH2
CYP27B1
OH
HO HO
D3 25OHD3
↓ P and Ca
↑ PTH
↓ FGF23 CH2

HO OH
1,25(OH)2D3

Fig. 1 | Vitamin d production and metabolism. 7-Dehydrocholesterol reductase (DHCR7) is involved in the last step in
cholesterol synthesis, which in the case of the skin competes with vitamin D production as each requires the substrate
7-dehydrocholesterol (7-DHC). Thus DHCR7 activity is relevant to vitamin D production as it controls the level of the
substrate for vitamin D production, namely 7-DHC. When exposed to UVB radiation, the path to vitamin D in the skin is
initiated with breaking of the B ring of 7-DHC to form previtamin D3, which undergoes isomerization to vitamin D3. Vitamin D3
then undergoes two hydroxylations first in the 25 position principally in the liver by CYP2R1 and then in the kidney either
to 24(OH)2D3 by CYP24A1 or to the active metabolite 1,25(OH)2D3. In the kidney, these latter hydroxylations are tightly
controlled by parathyroid hormone (PTH) that stimulates CYP27B1 and inhibits CYP24A1 and fibroblast growth factor 23
(FGF23) that does the reverse. 1,25(OH)2D3 induces CYP24A1 while suppressing CYP27B1. Calcium and phosphate directly
or via their regulation of PTH and FGF23 production likewise participate in this regulation. However, in extrarenal tissues,
regulation of CYP27B1 is different, and depends primarily on various cytokines functioning as autocrine or paracrine factors.

Nature Reviews | Endocrinology


Reviews

Cerebrotendinous
CYP2R1 and other 25-hydroxylases. As mentioned in the To date, five functional mutations in CYP2R1 have
xanthomatosis Introduction, the liver is the major, if not the sole, source of been described. The first was reported in a young man
The disease that results from 25OHD production from vitamin D8. 25-Hydroxylation from Nigeria who had severe bone disease and bio-
inactivating mutations in activity in the liver has been detected in both the mito- chemical evidence of rickets, including a low 25OHD
CYP27A1. Cerebrotendinous
chondrial and microsomal fractions 8. Numerous level but a normal 1,25(OH)2D level. This individual
xanthomatosis manifests as a
reduction in bile and enzymes, including CYP3A4 and CYP27A1, have sub- had a Leu99Pro mutation in CYP2R1, the first mutation
cholesterol metabolism. sequently been found to have 25-hydroxylase activity42. identified11. In vitro testing confirmed that this muta-
Initial studies suggested that CYP27A1, a mitochondrial tion decreased the activity of the mutated CYP2R1 to
enzyme with substantial homology to both CYP27B1 and produce 25OHD11. Subsequently, an additional muta-
CYP24A1 (the 1α and 24-hydroxylases, respectively), tion, Lys242Asn, was found in an individual carrying
was the major 25-hydroxylase; however, several prob- compound heterozygote mutations in CYP2R1, when
lems subsequently emerged with this identification. For additional Nigerian families with familial rickets were
example, CYP27A1 was originally identified as a sterol screened13. Two more mutations were found in siblings
27-hydroxy­lase involved in bile acid synthesis, and inacti­ of an Arabian family carrying compound heterozygote
vating mutations result in cerebrotendinous xanthomatosis mutations involving a splice site (367+1, Gly to Ala) and
with abnormal bile and cholesterol metabolism, but not an insert (768, iThr)45. Finally a mutation involving a
rickets43. Moreover, CYP27A1 does not 25-hydroxylate deletion — Gly42_Leu46 with an inserted Arg — was
vitamin D2 (refs2,3). Furthermore, in mouse studies, found in a French family46. These mutations, when
deletion of Cyp27a1 resulted in increased blood levels of tested in vitro, showed little or no 25-hydroxylase activ-
25OHD rather than a decrease, indicating that expres- ity13,45,46. In addition, the identified individuals main-
sion of CYP27A1 is not required for the production of tained normal or even high 1,25(OH)2D levels, and some
25OHD, and might even be inhibitory10. responded to treatment with vitamin D and 1αOHD,
Current data support CYP2R1 as the major with a further increase in 1,25(OH)2D following treat-
25-hydroxylase in the liver. This enzyme 25-hydroxylates ment13,46. Moreover, as the individuals reach adulthood,
both vitamin D 2 and vitamin D 3 equally, unlike they generally maintained normal calcium homeosta-
CYP27A1, which does not hydroxylate vitamin D2 as sis without vitamin D supplementation46. Thus, other
noted above. In humans, CYP2R1 is expressed primar- enzymes with 25-hydroxylase activity might be upregu­
ily in the microsomal fraction of the liver and testes42,44. lated to compensate for the reduced CYP2R1 activity.
As noted in the Introduction, deletion of Cyp2r1 in Nevertheless, the affected individuals developed classic
mice reduces but does not eliminate 25OHD produc- rickets, albeit sometimes not diagnosed until early teens,
tion and has little effect on blood levels of calcium and along with high PTH and alkaline phosphatase levels as
phosphate10. These findings indicate the presence of well as very low 25OHD levels13,45,46. Affected individu-
a compensatory mechanism by other enzymes with als responded very well to physiological doses of calci­
25-hydroxylase activity. Even the double deletion of fediol (25OHD)13. This form of rickets, comparable to
Cyp2r1 and Cyp27a1 in mice does not reduce the blood vitamin D deficiency rickets with disrupted endochon-
level of 25OHD to zero10. In humans, however, mutations dral bone formation and impaired mineralization of oste-
in CYP2R1 have a greater inhibitory effect on 25OHD oid affecting growth and skeletal integrity, is currently
levels and result in rickets, but the severity of the rickets called vitamin D-dependent rickets type 1B (VDDR1B).
is not as great as seen with mutations in CYP27B1 (ref.13). One of the aforementioned compensating enzymes
could be CYP3A4, the major drug-metabolizing enzyme
Box 1 | Mutations in the vitamin d metabolism pathway that is preferentially located in the liver and intestine47.
This enzyme has 25-hydroxylase activity47, however, it
• 7-dehydrocholesterol reductase (dHcr7). This enzyme converts the substrate also serves to degrade both 25OHD and 1,25(OH)2D
for vitamin D formation (7-dehydrocholesterol) to cholesterol. Polymorphisms are with hydroxylations in the 23 and 24 positions as well,
associated with reduced 25-hydroxyvitamin D (25OHD) levels, and inactivating
in a similar manner to CYP24A1 (ref.48). An activating
mutations have been associated with increased vitamin D production38.
mutation in CYP3A was described in 2018 that increases
• cyP2r1. This enzyme is considered the major vitamin D 25-hydroxylase. Inactivating the catabolism of 1,25(OH)2D and presumably 25OHD
mutations result in a form of rickets currently called vitamin D-dependent rickets,
leading to rickets with reductions in both 25OHD and
type 1B (VDDR1B)46. Affected individuals have low 25OHD levels but normal
1,25-dihydroxyvitamin D (1,25(OH)2D) levels. They can be treated with 25OHD 1,25(OH)2D, decreased serum calcium and phosphate
(calcidiol) and may no longer require treatment as adults. with elevated PTH and alkaline phosphatase. The
• cyP27B1. This enzyme is the 25OHD-1α hydroxylase responsible for producing the authors of this study called this type of rickets vitamin
hormonal form of vitamin D — 1,25(OH)2D. Mutations in this enzyme result in a form D-dependent rickets type 3 (ref.49).
of rickets called pseudovitamin D deficiency rickets or vitamin D-dependent rickets,
type 1A (VDDR1A)52. Affected individuals present with low 1,25(OH)2D but normal CYP27B1. To date, CYP27B1 is the only 25OHD-1α
levels of 25OHD. They respond to 1,25(OH)2D (calcitriol) but not to vitamin D. hydroxylase to have been described. Within cells, this
• cyP24A1. This enzyme is the 24-hydroxylase with both 25OHD and 1,25(OH)2D as enzyme is found in mitochondria, and while the kidney
substrates. Its primary role is that of catabolizing its substrates to inactive forms. might be the main contributor to circulating 1,25(OH)2D,
Mutations result in a form of idiopathic infantile hypercalcaemia (IIH), although with a number of other tissues express this enzyme includ-
the discovery of these mutations, this form of IIH is no longer idiopathic and, as the ing the epidermis and other epithelial tissues, bone, pla-
disease can manifest in adults, is no longer just infantile15,63. In children, hypercalcaemia centa and cells of the immune system12. The regulation
is the major concern. Presentation in adults is usually with a history of kidney stones of CYP27B1 expression in these sites outside the kid-
associated with hypercalcaemia and elevated 1,25(OH)2D levels.
ney differs from that in the kidney, where regulation by

www.nature.com/nrendo
Reviews

PTH and FGF23 dominate12. In non-renal tissues, deficiency present with low 25OHD levels and often
CYP27B1 is regulated by a number of cytokines including normal 1,25(OH)2D levels, except those with extreme
interferon-γ and tumour necrosis factor, but not by cal- vitamin D deficiency59. Patients with hypophosphatae-
cium, PTH or 1,25(OH)2D50. This accounts for the hyper- mic rickets (such as X-linked or autosomal dominant
calcaemia with inappropriate elevations of 1,25(OH)2D hypophosphataemic rickets) will also present with low
that complicate a number of granulomatous diseases 1,25(OH)2D and low phosphate levels, but these forms
and lymphomas, demonstrating stimulation by these of rickets are associated with elevated FGF23 levels,
cytokines overexpressed in these diseases with lack of which is not found in those with CYP27B1 mutations60.
feedback by the elevation of calcium and 1,25(OH)2D and As noted above, calcitriol is the treatment of choice, but
suppressed PTH12,51. In this section, we focus specifically needs to be accompanied by an adequate intake of cal-
on the mutations in CYP27B1. cium59. Monitoring urine calcium excretion is necessary
Mutations in CYP27B1 cause a disease previously to avoid overdosing that can result in nephrocalcinosis
referred to as pseudovitamin D deficiency rickets, and stone formation59.
which is now known as vitamin D-dependent rickets
type 1A (VDDR1A). A few years after the discovery CYP24A1 and other 23- and 24-hydroxylases. In most tis-
of 1,25(OH)2D as the active metabolite of vitamin D sues, CYP24A1 is the dominant 24-hydroxylase involved
and the development of assays to measure it, Fraser and in vitamin D metabolism. The liver and intestine seem
colleagues52 and Scriver and colleagues53 were the first to be exceptions in that CYP3A4 is highly expressed
to describe patients with VDDR1A. The investiga- in these tissues, although CYP3A4 lacks the specificity
tors observed that 1,25(OH)2D levels were very low in for vitamin D metabolites shown by CYP24A1 (ref.48).
patients who presented with rickets despite normal lev- Nevertheless, its induction by drugs such as rifampin can
els of 25OHD, distinguishing them from other patients result in reduced levels of 25OHD and 1,25(OH)2D, which
with vitamin D-deficient rickets. They called the dis- leads to osteomalacia61. Both CYP24A1 and CYP3A4 have
ease pseudovitamin D deficiency rickets because they 24-hydroxylase and 23-hydroxylase activity, although,
could not cure the rickets with vitamin D. Several years at least for CYP24A1, the ratio is species-specific in that in
after the first descriptions of VDDR1A, Delvin and col- humans 24-hydroxylase activity predominates, whereas
leagues54 found that VDDR1A can be treated effectively in the opossum 23-hydroxylase predominates62,63. Both
with 1,25(OH)2D (calcitriol), but not with vitamin D. enzymes are induced by 1,25(OH)2D, and in the intestine
It was not until 1997, however, that the cloning and the induction of CYP3A4 seems to be at least as great as
sequencing CYP27B1 was accomplished14,55–57. In one that of CYP24A1 (ref.64). That said, CYP24A1 is the domi-
of these studies, Portale and colleagues sequenced the nant vitamin D catabolic enzyme in most tissues and thus
human gene for the first time and identified the first mut­ is the focus of the remainder of this section.
ation in a patient with VDDR1A14. Subsequently, muta- The 24-hydroxylase pathway terminates with the bio-
tions have been found throughout the gene58. Both the logically inactive calcitroic acid, whereas the 23-hydroxy­
renal and extrarenal CYP27B1 have the same sequence, lase pathway produces the biologically active 1,25,26,23
but their differences in regulation occur as a result lactone. CYP24A1 catalyses all steps in these pathways65.
of tissue-specific multicomponent control modules, Both 1,25(OH)2D and 25OHD are metabolized by
as noted in a 2017 article describing loss of function CYP24A1, although 1,25(OH)2D is the preferred sub-
enhancer deletion studies that demonstrated different strate63. CYP24A1 is generally considered a catabolic
regions of the CYP27B1 promoter region responding to enzyme in vitamin D metabolism, but this is only par-
different regulators in different cells27. tially true as some of its products have biological activ-
The clinical symptoms of rickets generally occur in ity. For example 1,24,25(OH)3D has substantial affinity
the first year of life — embryonic development is nor- for the VDR with biological activity being ~10% that
mal52. In addition to the skeletal defects of classic rickets of 1,25(OH)2D66. A specific receptor for 24,25(OH)2D
(flared metaphyses, bowing of the legs, rachitic rosary has been identified in bone and other tissues, such as
and generalized osteopenia), patients develop tooth the skin. This receptor has been shown to be involved
dysplasia, muscle weakness and growth retardation, in fracture repair67 and might have a role in mediating
and can have convulsions and/or tetany due to the very the effects of 24,25(OH)2D on endochondral bone for-
low calcium levels59. Pseudofractures can develop in the mation68. The deletion of CYP24A1 in mice, however,
long bones59. Laboratory assessment in addition to low resulted in defective mineralization of intramembranous
1,25(OH)2D include hypocalcaemia, hypophosphatae- (not endochondral) bone69, although this seemed to be
mia and secondary hyperparathyroidism. In patients due to the elevated levels of 1,25(OH)2D in the knockout
with rickets, 25OHD levels are generally normal or even mouse and not to the loss of 24,25(OH)2D per se.
elevated when dietary insufficiency is excluded59. Thorax Inactivating mutations in CYP24A1 have been estab-
deformities can lead to poor air exchange with predis- lished as a major cause of idiopathic infantile hypercalce-
position to pulmonary infections59. As mentioned previ- mia (IIH), a syndrome that is associated with severe
ously, physiological doses of calcitriol (1,25(OH)2D) are hypercalcaemia, hypercalciuria and nephrocalcinosis,
curative59. Distinguishing rickets from vitamin D defi- decreased PTH, low 24,25(OH)2D and inappropriately
ciency, especially in countries where low calcium and normal-to-high 1,25(OH)2D15. In a 2017 survey, 21 mis-
vitamin D intake are common, can be difficult59. As such, sense mutations were reported in the CYP24A1 gene70.
a careful family and dietary history is necessary; how- With regards to the clinical presentation of patients,
ever, it is important to note that patients with vitamin D not all patients with mutations in CYP24A1 present

Nature Reviews | Endocrinology


Reviews

a Free hormone hypothesis b Megalin–cubilin-mediated uptake this hypophosphataemic syndrome can also present with
Free 25OHD hypercalcaemia with increased 1,25(OH)2D levels78,79.
DBP DBP However, 24,25(OH)2D levels are normal in patients
with this condition79. CYP24A1 mutations can be identi-
25OHD
Megalin Cubilin fied by determining the ratio of 25OHD to 24,25(OH)2D
by tandem mass spectrometry80 with confirmation of
the findings by gene sequencing80. Vitamin D-sufficient
subjects have a 25OHD:24,25(OH)2D ratio between 5
and 25, whereas patients with CYP24A1 mutations have
Mitochondrion a ratio above 80 (ref.80) (Box 1). Treatment involves ces-
sation of all vitamin D supplements and if necessary a
HSP70 DBP
low-calcium diet.

Cellular transport of D metabolites


HSP70 VDR
In this section, we describe the main transport of vita-
1,25(OH)2D Nucleus min D and its metabolites in blood via DBP, the role
of DBP in regulating the transport of these metabolites
Intracellular HSP70 shuttle into cells including the free hormone hypothesis, and
the intracellular transport of the vitamin D metabolites
Fig. 2 | Transport of 25oHd between and within cells. Vitamin D binding protein (DBP)
is the major transport protein for the vitamin D metabolites within the bloodstream.
by HSP70 (Fig. 2).
For most cells, however, it is the unbound or free form of 25-hydroxyvitamin D (25OHD) that
enters cells, whereas in a few tissues such as the kidney and probably also the parathyroid DBP. DBP (originally known as Gc-globulin) was dis-
gland and placenta, DBP with its bound 25OHD enters the cell via the megalin–cubilin covered in 1959 (ref.81), but it was not until 1975 that
complex. Once in the cell, HSP70 functions as an intracellular shuttle first offloading its function as a carrier of vitamin D metabolites was
25OHD from DBP, moving it to mitochondria for metabolism to 1,25(OH)2D, and subsequently described82. In healthy non-pregnant individuals, the
facilitating the movement of 1,25(OH)2D into the nucleus for binding to the vitamin D binding of 25OHD and 1,25(OH)2D to DBP accounts
receptor (VDR). The generality of this intracellular shuttle mechanism has not been for ~85% of the total levels of these vitamin D metabo-
established for all target cells. lites, whereas albumin with its lower affinity binds ~15%
of the total levels of the vitamin D metabolites. The
in childhood. For example, several patients presented non-protein-bound or free 25OHD fraction comprises
during pregnancy, presumably because of the increased about 0.03% of the total 25OHD, with free 1,25(OH)2D
production of 1,25(OH)2D by the placenta71,72. These slightly higher at 0.4% of the total 1,25(OH)2D. The
patients had a history of renal stones. Other adults affinity of DBP for the vitamin D2 metabolites is less than
with this syndrome have been described who were not the affinity of DBP for 25OHD and 1,25(OH)2D83, such
pregnant73,74. Such adults generally present with early that the free fraction would be expected to be higher, but
onset nephrolithiasis and/or nephrocalcinosis, and the this has not been directly determined.
hypercalcaemia is generally less severe than in those For lipophilic hormones such as steroid hormones,
presenting as infants. Although CYP24A1 mutations thyroid hormone and vitamin D metabolites, it is the free
are generally recessive, adults and several infants hetero­ fraction that is thought to enter cells — a concept known
zygous for CYP24A1 mutations who presented with IIH as the free hormone hypothesis84. However, at least for
or nephrocalcinosis have been described74,75. These indi- some tissues, a transport system has been identified that
viduals typically came from families with a history of takes up the 25OHD (and presumably other vitamin
renal stones, indicating that screening such families for D metabolites) attached to DBP. This system involves
CYP24A1 mutations may be warranted. megalin and/or cubilin85. This complex is found pri-
IIH was first described in the UK in the 1950s marily in the kidney, parathyroid gland, placenta and a
when infants were routinely treated with high doses of few other tissues19, most of which have not been tested
vitamin D to prevent rickets76. The exact cause of IIH, as to the role of this complex in taking up DBP-bound
however, was not known at that time. We now know that vitamin D metabolites. However, mice lacking the
IIH can be caused by at least two other conditions from megalin and/or cubilin complex have poor survival
which patients with CYP24A1 mutations should be dis- with evidence of osteomalacia, which indicates its role
tinguished. The first is the Williams–Beuren syndrome. in vitamin D transport into critical cells involved with
Children with Williams–Beuren syndrome also present vitamin D signalling85. Knockout of Lrp2 (encoding
with hypercalcaemia during infancy, but have a number of megalin) or Dbp genes in mice provides a useful tool for
phenotypic features that distinguish them from children determining the respective roles of DBP and megalin
with CYP24A1 mutations who have IIH77. For example, and/or cubilin in transporting vitamin D metabolites
children with Williams–Beuren syndrome have a deletion into cells. In Dbp-knockout mice the vitamin D metab-
in chromosome 7 that includes loss of a transcription fac- olites are presumably all free and/or bioavailable21,86.
tor, the Williams syndrome transcription factor (WSTF), Unlike the Lrp2-knockout mice, mice lacking Dbp do
which can cause hypercalcaemia due to unchecked VDR not show evidence of vitamin D deficiency unless placed
activity. The second is a hypophosphataemic syndrome on a vitamin D-deficient diet despite having very low
resulting from mutations in SLA34A1, which encodes the levels of serum 25OHD and 1,25(OH)2D21. In subse-
renal sodium phosphate transporter A1. Children with quent studies, tissue levels of 1,25(OH)2D were found to

www.nature.com/nrendo
Reviews

be normal in the Dbp-knockout mice as were markers of factor (MAF), as only the glycosylated vari­ants can be
vitamin D action such as expression of intestinal TRPV6, converted to DBP–MAF100. The different variants have
calbindin 9k (also known as protein S100G), PMCA1b been shown to have different affinities for 25OHD101,
and renal TRPV5 (ref.86). These studies support the although the results and biological significances of dif-
concept that DBP serves as a critical reservoir for ferences in affinities have not been consistent among
the vitamin D metabolites reducing the risk of vitamin D different laboratories102–105. That said, the affinities of
deficiency when epidermal production or intake is low, the DBP variants seem to be altered in different clini­
but entry of vitamin D metabolites into cells involves the cal conditions. For example, in the third trimester of
free fraction in most cells. The more detrimental effect pregnancy, directly measured free 25OHD tends to be
of deleting the LRP2 gene, however, might be explained increased and free 1,25(OH)2D is substantially higher
by the inability of the kidney to reclaim the DBP-bound in pregnant women than would be expected by meas-
metabolites that are then lost in the urine85. uring total levels in comparison to non-pregnant
In further support of the mouse studies, a 2019 case women106,107. This finding suggests that the affinity of
report reported a large deletion in the coding portion DBP for the vitamin D metabolites is decreased during
of the DBP gene (and adjacent NPFFR2 gene) in a sin- pregnancy. Similarly, directly measured free 25OHD
gle family22. The proband had normal calcium, phos- and 1,25(OH)2D tend to be higher in outpatients with
phate and PTH levels with vitamin D supplementation cirrhosis than in other groups107,108 despite the fact that
despite very low levels of 25OHD, 24,25(OH)2D and individuals with cirrhosis have reduced total vitamin D
1,25(OH)2D that were not responsive to massive doses metabolite concentrations108. This changed relationship
of vitamin D (oral or parenteral). The free 25OHD between total and free 25OHD levels can be appreciated
was nearly normal. The carrier sibling had vitamin D by the steeper but more variable slope in the correlation
metabolite levels between those of the proband and the between free 25OHD and total 25OHD in patients with
healthy sibling. Thus, studies in both Dbp-null mice and liver disease compared with that in healthy people. These
humans with mutations in DBP support the free hor- data indicate that there is a decreased affinity of DBP for
mone hypothesis while also supporting the role of DBP 25OHD in these patients. A similar observation has been
as a circulating reservoir for vitamin D metabolites. made in patients in nursing homes109,110.
Expression of DBP is widespread, but by far the great- Allelic variation might also affect the affinity of
est proportion originates in the liver87. The expression of DBP for the vitamin D metabolites of DBP. Most assays
DBP is increased by oestrogen88, which is illustrated by have shown a modest reduction in circulating levels
the rise in DBP during pregnancy89,90 and with oral con- of 25OHD levels in individuals with the Gc2 variant of
traceptive use91. Androgens, on the other hand, do not DBP compared with those with the Gc1 variants111–114,
seem to affect DBP expression88. Dexamethasone, and but the differences in directly measured free 25OHD lev-
certain cytokines such as IL-6 also increase DBP pro- els are nearly the same115. A genome-wide association
duction, whereas transforming growth factor-β (TGFβ) study reported in 2010 demonstrated that rs2282679, an
is inhibitory92. These drugs and cytokines contribute intronic polymorphism in the DBP gene, was likewise
to the changes in DBP production following trauma93 associated with lower 25OHD and DBP levels40. On the
and acute liver failure94. Primary hyperparathyroidism, other hand differences in these alleles were not found
on the other hand, is associated with a reduction in to contribute to a difference in fracture rate in a study
DBP levels, that is likely to contribute to the reduced in 12,781 African American and white individuals116
25OHD levels in these patients as the free 25OHD is not or in patients with other calcaemic or cardiometabolic
reduced95. Vitamin D does not regulate DBP production, diseases in the Canadian Multicentre Osteoporosis
nor do any of its metabolites92,96. study117. Nevertheless, a large number of chronic dis-
Human DBP is 58 kDa, although differences in glyco- eases including type 1 and type 2 diabetes mellitus118–120,
sylation of the protein can alter its size87. DBP is the most osteoporosis121–123, chronic obstructive lung disease124,
polymorphic gene known. Over 120 variants have been endometriosis125, inflammatory bowel disease126, some
described based on electrophoretic properties97 with cancers127–130 and tuberculosis131 have been associated
1,242 polymorphisms currently listed in the NCBI data- with various DBP alleles.
base98. Of these variants, the Gc1F and Gc1S (rs7041)
and Gc2 (rs4588) are the most common. The Gc1F and Intracellular transport of vitamin D metabolites. The
Gc1S involve two polymorphisms, one at amino acid 432 primary role of DBP is to transport the vitamin D
(416 in the mature DBP) and one at amino acid 436 (420 metabolites between cells, but an intracellular transport
in the mature DBP). The 1F allele encodes the sequence network also exists. What was originally called intracel-
of amino acids between 432 and 436 as DATPT, the lular DBP, now known as HSP70 (ref.132), promotes the
1S allele encodes the sequence EATPT. The Gc2 allele cellular uptake and distribution of vitamin D metabolites
encodes DATPK. within the cell23. Overexpression of this protein in cells
In addition to the polymorphisms, the level of glyco­ from Old World monkeys (MLA-144 cells) increased
sylation further distinguishes the Gc1 variants from the the amount of 25OHD in these cells, promoted synthe-
Gc2 variant. The threonine (T) in Gc1 binds N-acetyl­ sis of 1,25(OH)2D and increased the genomic actions
galactosamine to which galactose and sialic acid bind of 1,25(OH)2D as assessed by increased expression of
in tandem. The lysine (K) in comparable position in Gc2 CYP24A1 (ref. 133) . Antisense constructs to HSP70
does not bind these residues18,99. These differences affect blocked these actions. The concept is that HSP70 func-
the conversion of DBP to DBP macrophage activation tions to transport 25OHD to the mitochondria where

Nature Reviews | Endocrinology


Reviews

CYP27B1 (the 1-hydroxylase) converts the 25OHD to VDR structure. The structure of VDR is shown in Fig. 3
1,25(OH)2D, which is then transported to the nucleus (ref.138). The human VDR contains 427 amino acid res-
where it binds the VDR for its genomic actions. HSP70 idues. The core domains of VDR are a highly variable
has higher affinity for 25OHD than DBP facilitating C terminal ligand-binding domain (LBD), a highly
25OHD transport across the cell membrane, but lower conserved DNA binding domain (DBD) and a hinge
affinity for 1,25(OH)2D than VDR, which facilitates region connecting the DBD to the LBD. As indicated
1,25(OH)2D offloading onto VDR in the nucleus133. by X-ray crystallography data, the DBD contains two
zinc fingers. Each zinc atom is held in a tetrahedral
VDR configuration by four cysteine residues139. The LBD
Tissue distribution. VDR, a member of the nuclear is composed of 12 α-helices (H1–12)140. The terminal
receptor superfamily, was initially identified in the H12 in the non-liganded VDR is open, but with ligand
intestine, the site which at the time was the best place to binding H12 moves over the ligand and encloses it in
study the action of vitamin D134. Although the highest the binding pocket140. Along with H3 and H4, H12 (the
concentration of VDR is in tissues involved in the main- ligand-dependent activation function (AF2) corre-
tenance of calcium homeostasis (intestine, kidney and sponds to H12) provides the interface for co-activator
bone), VDR has been found in many other tissues binding140. The LBD of VDR is large and can accom-
and cells not involved in calcium homeostasis including modate a number of ligands including lithocholic acid
pancreas, skin, pituitary, breast, colon and prostate cancer in addition to 1,25(OH)2D, although 1,25(OH)2D is the
cells, and immune cells, suggesting that vitamin D affects preferred ligand as it has the highest affinity for VDR141.
many cellular processes and a role for vitamin D beyond Moreover, a model has been proposed in which the
regulation of calcium homeostasis135,136. The functions of VDR can accommodate two types of ligand configura-
VDR in these other tissues and cells is a topic of ongoing tions — the genomic pocket for analogues with genomic
investigation, as the expression levels of VDR in different activity and the alternative pocket for analogues trigger-
cells are quite variable and can change with differentiation ing rapid and non-genomic responses142. This model,
of the cell135,137. however, remains controversial.
In addition to X-ray crystallography data, which
Side view Top view
provide a static view of VDR structure, complementary
RXRα LBD data from NMR has provided insights into the confor-
mational changes that occur when VDR is bound to
different ligands. These conformational changes result
in differing biological activity143. Solution methodology
(such as X-ray scattering and cryo-electron microscopy,
RXRα LBD as well as hydrogen–deuterium exchange mass spec-
trometry) has also yielded new information concerning
VDR LBD VDR LBD the structure and function of the VDR–RXR hetero­
dimer144–146. The VDR hinge region has been found to
RXRα be essential for stabilizing the VDR–RXR complex and
DBD
VDR DBD VDR maintaining the integrity of the functional structure144,145.
RXRα DBD hinge
Conformational dynamics indicate that the binding of
the heterodimer to DNA results in significant alterations
VDR in the conformation of the LBD within regions that are
DBD critical for interaction with co-activators146. Sequences
of the DNA response element and ligand structure dif-
ferentially affect the conformation of the heterodimer,
which can lead to selectivity of co-activator binding and
Fig. 3 | Vdr–rXrα vitamin d receptor binding site (dr3) model. The vitamin D receptor gene-specific effects of VDR138,146. Further studies using
(VDR)–retinoid X receptor-α (RXRα) heterodimer was created in the ‘open’ conformation,
solution structure methods for VDR complexes that are
and a partial cryo-electron microscopy model was used to guide the creation of the model.
The VDR–RXRα heterodimer forms an extended L-shaped organization with RXRα DNA difficult to crystallize will provide an increased under-
binding domain (DBD) occupying one half-site of DR3 (a synthetic oligonucleotide standing of the cooperative influence of 1,25(OH)2D,
representing an idealized direct repeat VDR binding site; shown in black in the figure) DNA, VDR–RXR heterodimer and co-activator binding
at the 5′-end and VDR DBD occupying another at the 3′-end145. The VDR hinge domain, on the mechanism of action of vitamin D.
a connection linker between the DBD and the ligand binding domain (LBD), plays a key
role in determining structural orientation. In this model, the VDR hinge domain adopts an Mutations of the VDR. When VDR is defective due to
α-helical structure residing close to DR3 and makes extensive contacts with phosphate mutations in VDR, the result is early onset of rickets
and the backbone of DNA145. This is consistent with observations made with hydrogen– with resistance to 1,25(OH)2D. The resulting form of
deuterium exchange mass spectrometry showing that the VDR hinge undergoes different rickets is termed hereditary vitamin D-resistant rickets
extents of protection upon binding to various VDR binding sequences138. Unlike VDR ,
(HVDRR), which is also known as vitamin D-dependent
DNA binding does not perturb RXRα hinge dynamics as it forms a flexible linker that allows
enhanced adaptability of the heterodimer to diverse response elements138. This model rickets type II. Patients have low serum calcium and
suggests that the hinge could possess intrinsic properties to orient the LBD dimer in a phosphate, high circulating levels of PTH and alkaline
precise way and provide a structural basis for crosstalk between LBDs and DBDs. VDR , phosphatase and very high levels of 1,25(OH)2D30. To
RXRα and DNA are coloured in blue, grey and black, respectively. Image courtesy of date, at least 49 different mutations in over 100 patients
P. Griffin, Scripps Research Institute, USA. have been described30,147. Mutations have been identified

www.nature.com/nrendo
Reviews

in the DBD and LBD, which inhibit VDR–RXR hetero­ and the mediator (MED) complex, which functions to
dimerization30. These mutations result in bone defects recruit RNA polymerase II136,153,154. VDR binding sites
and alopecia areata30. Mutations in the LBD that alter are associated with sites for other transcription factors
binding affinity for 1,25(OH)2D (for example, Arg­ such as C-EBPα, C-EBPβ, RUNX2 and PU.1, which can
274Leu, Arg274His and His305Gln) demonstrate the cooperate with VDR and VDR co-regulators to influence
important amino acids for contact points of 1,25(OH)2D. 1,25(OH)2D responses in target cells155–157. Genome-wide
The mutation in H12 (Glu420Lys) has been found not studies have shown that VDR binding sites are located
to affect ligand binding, DNA binding or VDR–RXR within introns, in intergenic regions and many kilobases
heterodimerization but prevents co-activator binding, upstream or downstream of regulated genes as well as in
indicating the critical role of co-activator recruitment promoters adjacent to transcription start sites153.
by VDR to enable its function in preventing rickets30,147. Consistent with the ability of VDR to recruit co-
Mutations that disrupt ligand binding or co-activator regulatory complexes that regulate chromatin structure,
binding result in rickets but do not cause alopecia areata, evidence provided over the past several years shows that
indicating that control of hair follicle cycling by VDR levels of acetylation at H4K5, H3K9 and H3K27 define
is ligand-independent30,147. Mouse models of HVDRR sites of action of 1,25(OH)2D and occur within enhanc-
have been created that express human VDR mutants ers that regulate 1,25(OH)2D target genes158. Specific
in the LBD, in which 1,25(OH)2D binding is impaired in enhancer deletions in vivo have provided an important
the absence of endogenous mouse Vdr148–150. In these advance in the determination of tissue-specific determi­
VDR-mutant mice with impaired 1,25(OH)2D binding, nants of VDR actions. These studies using loss-of-
the alopecia areata but not the biochemical and skele- function enhancer deletion studies in mice have identi­fied
tal aspects of the Vdr-null mice was rescued, confirm- a kidney-specific multicomponent control module distal
ing a role for unliganded VDR in hair follicle cycling. to Cyp27b1 that controls basal and hormone-regulated
Interestingly, skeletal defects in the mice expressing expression of Cyp27b1 in the kidney, but that is not active
ligand-binding-deficient VDR were more severe than in other cells such as immune cells which use another
in the Vdr-null mice148,149. In addition, in a subset of module conferring cytokine regulation. These data
VDR target genes in the intestine (Trpv6, Cyp24a1 and provide an insight, for the first time, into cell-specific
Slc30a10), gene downregulation in the Vdr-mutant mecha­nisms involved in the regulation of CYP27B1 that
mice was stronger than in the Vdr-null mice, suggest- had remained undefined27.
ing suppression by VDR in the absence of 1,25(OH)2D3 Studies involving genome-scale definition of VDR
(ref. 149) . Although VDR binding to DNA is mostly action have also shown the translational potential of
ligand-dependent, these in vivo findings are notable VDR as a drug target. The VDR ligand calcipotriol
since they suggest that signalling pathways mediated (a less-hypercalcaemic analogue of 1,25(OH)2D) was
by unliganded VDR as well as ligand-dependent VDR found to suppress profibrotic gene expression in liver
signalling can have a substantial role in calcium and stellate cells through antagonism of SMAD3 recruit-
bone homeostasis. ment, thus defining a role for VDR in the regulation
Patients with HVDRR are treated with pharmacolog- of liver fibrogenesis159. VDR was also found to act as a
ical doses of vitamin D. Treatment with 25(OH)D3 or master transcriptional repressor of the pancreatic stel-
1,25(OH)2D3 has been shown to benefit individuals with lar cell activation state, which suggests that reprogram-
mutations in the LBD that reduce binding affinity for ming by vitamin D could be an adjunct in pancreatic
1,25(OH)2D147. However, if individuals fail to respond ductal adenocarcinoma therapy160. In addition, VDR
to vitamin D or vitamin D metabolites, oral calcium was found to have a role in β-cell survival by reversing
therapy or intravenous calcium infusion reverses all inflammation-driven transcriptional changes161. VDR
aspects of HVDRR except alopecia areata151, which ligand was shown to promote VDR association with the
indicates that the most important role of 1,25(OH)2D PBAF chromatin remodelling complex, which enhanced
and/or VDR in calcium homeostasis is regulation of chromatin accessibility at vitamin D response elements
intestinal calcium absorption. As patients grow older, resulting in inhibition of inflammatory response genes161.
the need for calcium supplementation lessens, suggest- These findings indicate that VDR transcriptional
ing VDR-independent pathways for intestinal calcium mechanisms are involved in β-cell survival. Further stud-
absorption in patients with HVDRR later in life147,152. ies involving genome-scale definition will reveal addi-
tional novel mechanisms of 1,25(OH)2D action that could
Regulation of VDR. The multiple actions of result in new approaches not only to sustain calcium bal-
1,25(OH)2D that are mediated by VDR involve bind- ance, but also to enhance the suggested antitumour and
ing of 1,25(OH)2D and VDR–RXR to specific DNA anti-inflammatory actions of 1,25(OH)2D (Box 2).
sequences in the regulatory regions of target genes
and the recruitment of co-regulatory complexes, The role of vitamin D in skin
which results in the modulation of gene expression153. The epidermis and hair follicle are models of vitamin D
Co-regulatory factors involved in VDR-mediated tran- signalling that demonstrate the role of vitamin D beyond
scription include factors with histone acetylase activity, that of regulating bone mineralization and intestinal
including steroid receptor co-activator 1 (SRC1), SRC2 calcium transport. Aside from producing vitamin D,
and SRC3, and CREB-binding protein p300. Additional epidermal cells (keratinocytes) convert the vitamin D to
co-regulatory factors include SWI–SNF ATP-dependent its biologically active metabolite 1,25(OH)2D162. More­
chromatin-remodelling complex, methyltransferases over, keratinocytes contain VDR163,164, and respond to

Nature Reviews | Endocrinology


Reviews

1,25(OH)2D with changes in proliferation and differenti- Lamellar bodies, which contain both long-chain fatty
ation31,165,166. Vitamin D signalling in keratinocytes works acids and antimicrobial peptides from the stratum gran­
in partnership with calcium, acting in part through ulosum, line the membrane separating the stratum
the calcium-sensing receptor (CaSR) and β-catenin1. granulosum and the stratum corneum184. The lamellar
Deletion of VDR and/or CaSR in epidermal keratino- bodies are positioned to offload their contents into the
cytes (epiVdr-null and/or epiCasr-null) disrupts epidermal extracellular space where the lipids and antimicrobial
differentiation167,168, impairs wound healing169,170 and pre- peptides contribute to the permeability barrier and
disposes to cancer171. Although VDR is generally thought defence against microorganisms185,186.
to function in combination with its ligand 1,25(OH)2D, 1,25(OH)2D with its receptor VDR regulate all steps
hair follicle cycling is one notable exception. Vdr-null in the differentiation process. For example, 1,25(OH)2D
animals develop loss of hair follicle cycling172,173, whereas induces keratin 1, involucrin and transglutaminase
those lacking Cyp27b1 or Casr, or those placed on a activity in the stratum spinosum, and filaggrin, loricrin,
vitamin D-deficient diet, do not1. On the other hand, antimicrobial peptide and long-chain fatty acid synthesis
β-catenin signalling is an important partner with VDR and cornified envelope formation in the stratum granu-
in regulating hair follicle cycling26,174 and stem cell acti- losum165,166,187–190. In addition to its role in epidermal dif-
vation as during the response to wounding169. Within ferentiation, 1,25(OH)2D promotes the innate immune
the epidermis, β-catenin has a dual role in enabling function of keratinocytes by inducing the Toll-like
proliferation as a transcription factor but facilitating receptor 2 (TLR2) and its co-receptor CD14 (ref.191).
differentiation as a key component of the membrane The induction of TLR2 and CD14 in turn has a feedback
E-cadherin–catenin complex (Box 3). element in that activation of TLR2 and CD14 induces
CYP27B1, which increases production of 1,25(OH)2D
Regulation of keratinocyte differentiation. The epider- that then induces cathelicidin, a potent antimicrobial
mis is an excellent example of differentiation (Fig. 4). The peptide191,192. Deletion of either Vdr or Cyp27b1 in mice
basal layer (stratum basale), which rests on the basal results in defective epidermal differentiation, a defective
lamina (the layer that separates the dermis and the epi- permeability barrier193, and a defective response of the
dermis), is where the interfollicular epidermal stem cells innate immune system to infection and wounding191,194.
reside175. These stem cells give rise to transient amplify- The process of epidermal differentiation and its
ing cells that begin the differentiation process as they regulation by 1,25(OH)2D and VDR is sequential in
move upward into the next layer, the stratum spinosum1. that different genes and pathways are induced in the
Cells in this layer produce keratins K1 and K10 (kera­ keratinocytes at their different states of differentia-
tins in the basal layer are K5 and K14)176. Moreover, one tion. VDR and CYP27B1 are expressed throughout
of the important components of the cornified envelope, the epidermis, although their highest expression is in the
involucrin177, is produced in the stratum spinosum along stratum basale195,196. As noted previously, the transcrip-
with the enzyme transglutaminase K, which is respon- tional activity of VDR is tightly regulated by a number
sible for the ε-(γ-glutamyl)lysine crosslinking of involu- of co-regulators that can show cell specificity136,153. The
crin and other substrates (for example, loricrin) into the major co-activator complexes in the epidermis regulat-
insoluble cornified envelope178. ing VDR function are MED (formerly called DRIP or
Above the stratum spinosum is the stratum granu­ VDR interacting proteins) of which MED1 is the major
losum, which as its name implies is characterized by VDR-binding component, and the SRC complex, of
electron-dense keratohyalin granules179. The larger gran- which SRC2 and SRC3 are the major VDR-binding com-
ules contain profilaggrin180. Profilaggrin is the precursor of ponents197. The functions of VDR in undifferentiated
filaggrin, which is thought to facilitate the aggregation keratinocytes are controlled primarily by MED, whereas
of keratin filaments181, but when hydrolysed is thought the actions of VDR in the more differentiated cells are
also to provide a moisturizing factor for the skin182. The regulated by SRC2 and SRC3 (refs197,198). MED1 and
smaller granules contain loricrin, which like involucrin SRC3 are reciprocally expressed in the epidermis, with
is an important component of the cornified envelope183. MED1 primarily found in basal keratinocytes and SRC3
in the more differentiated layers, that is, the stratum
granulosum197. Deletion of Med1 but not Src3 in mice
Box 2 | Key concepts in the regulation of tissue-specific actions of Vdr
results in increased keratinocyte proliferation198 and
• Conformational dynamics have shown that the vitamin D receptor (VDR)–retinoid X a profound disruption of hair follicle cycling199, and has a
receptor (RXR) heterodimer adopts different conformations depending on sequences greater effect on K1, K10 and involucrin expression than
in the DNA response element, ligand structure and co-regulatory protein binding138. does the deletion of Src. On the other hand, deletion of
• Genome-wide studies have noted that actions of 1,25(OH)2D involve regulation of Src3 reduces glucosylceramide production and lamellar
gene activity at a range of locations. VDR binding may be many kilobases upstream body formation, in a similar way to Vdr deletion193, and
or downstream of regulated genes154.
prevents 1,25(OH)2D-induced cathelicidin expression200.
• Sites of action of 1,25(OH)2D generally occur within enhancers that regulate As alluded to earlier, calcium partners with
1,25(OH)2D target genes154. 1,25(OH)2D and/or VDR in their effects on prolifera-
• Loss of function enhancer deletion studies in mice have identified a multicomponent tion and differentiation. Increases in the extracellular
control module that controls basal and hormone regulated expression of Cyp27b1. calcium concentration above 0.1 mM change the cells
• Genome-scale definition of VDR action has noted the translational potential of from a proliferative state to one of differentiation1. This
VDR as a drug target in liver fibrogenesis, pancreatic ductal adenocarcinoma and change in state includes the development of intercellu-
type 2 diabetes154.
lar contacts, the formation of the E-cadherin–catenin

www.nature.com/nrendo
Reviews

Box 3 | The skin as a model for the study of vitamin d we consider the roles of vitamin D signalling in hair
follicle cycling, skin cancer and wound healing.
• The skin is the major source of vitamin D as it contains a large store of The ΔNp63α and other mutant forms of p53 have
7-dehydrocholesterol for conversion to vitamin D upon UVB radiation exposure6. been implicated in vitamin D regulation of prolifera-
• All the enzymes required for vitamin D metabolism are present in the keratinocytes tion218,219, a role that may be important in the regulation
of the epidermis and hair follicle. Regulation of at least CYP27B1, but most likely by VDR in cancer. Similarly, VDR has been implicated
CYP24A1 as well, differs from that in the kidney1. as a master regulator of the MYC/MXD1 network220,
• Keratinocytes express vitamin D receptor (VDR), and so respond to its ligand playing a possible role in keratinocyte proliferation and
1,25(OH)2D with changes in proliferation, differentiation, barrier formation and cancer. The E3 ubiquitin ligase MDM2 has been found
immune function. However, at different stages in the differentiation process different
to bind to and decrease the function of VDR, potentially
co-regulators of VDR are required1,197.
contributing to uncontrolled proliferation and cancer221.
• Hair follicle cycling is the best example known of a VDR-regulated process not
Further investigation of their roles in vitamin D signal-
requiring 1,25(OH)2D173.
ling in the skin is required to fully understand their level
• Lack of VDR in the epidermis and hair follicle predisposes to deficient wound healing
of participation.
and cancer169,255.

Hair follicle cycling. The hair follicle cycle is composed


complex containing phosphoinositide 3-kinase (PI3K), of three distinct phases. The phase of hair follicle growth
α-catenin and β-catenin, and phosphatidyl inositol 4- is called anagen. Anagen is initiated following contact
phosphate 5-kinase 1α (PIP5K1α)1. All of these compo- between the hair follicle and the dermal papilla, a spe-
nents have critical roles in the differentiation process. cialized collection of mesenchymal cells in the dermis222.
For example, PI3K and PIP5K1α are involved in the for- At a given point following the growth phase, the follicle
mation of PIP3 in the membrane, which among other regresses — this is called catagen222. The trigger for cata­
functions, activates phospholipase Cγ1 (PLCγ1), a key gen, however, is not clear. Following catagen, the hair
enzyme in maintaining the increased intracellular cal- follicle enters a resting phase called telogen222. Following
cium required for differentiation201. In addition to these the developmental cycle, which leads to the initial coat
functions, the calcium switch leads to the membrane of hair, the hair follicle undergoes repetitive cycling
localization of the CaSR, and SRC kinases, which also until senescence222. Although signals from the dermal
have important roles in the calcium-induced differen- papilla are clearly involved in the reinitiation of anagen,
tiation process by helping form the E-cadherin–catenin the length of telogen and the timing of the new cycle
complex202–207. α-Catenin links the E-cadherin–catenin are not well understood and vary between the sexes
complex to the actin cytoskeleton enabling cell migra- and different parts of the body222. Only the proximal or
tion. CaSR is essential for the keratinocyte response to dermal portion of the hair follicle cycles; the distal
calcium204,208. Deletion of Casr from keratinocytes blocks or epidermal portion does not222.
their response to extracellular calcium by reducing their VDR regulation of the hair follicle cycle begins only
stores of calcium and preventing the formation of the after the developmental hair follicle cycle is completed,
E-caderin–catenin complex204,209. VDR and/or CYP27B1 so that the initial coat of hair initiated during embryo­
and CaSR are mutually required for each other’s full genesis proceeds normally in Vdr-null mice172. Hair fol-
expression168,210, and participate in the joint regulation licle cycling is dependent on stem cells in the bulge, a
of the expression of a number of genes by calcium and specialized region of the outer root sheath (ORS) below
1,25(OH)2D204,211–213. the sebaceous gland223. These cells have a higher level
The precise interaction between VDR and β-catenin of VDR expression than other cells in the hair follicle.
is dependent on which cells are involved. Both VDR and The ORS is a direct extension of the stratum basale, and
β-catenin are required for hair follicle cycling, but their separates the hair follicle from the surrounding connec-
interactions in the interfollicular epidermis are more tive tissue sheath224. The ORS is home to other stem cell
complex214. As noted previously, although 1,25(OH)2D is niches, including cells supporting the sebaceous gland
not required for VDR regulation of hair follicle cycling, and the epidermis itself 225. The stem cells that support
β-catenin is. Deletion of Ctnnb1 (encoding β-catenin) the sebaceous gland and the epidermis are crucial in the
from keratinocytes in mice results in a nearly identical epidermal response to wounding225.
phenotype as Vdr deletion with loss of hair follicle cycling A marked characteristic of many patients with muta-
combined with sebocyte hypertrophy, dermal cysts lined tions in VDR226,227 is alopecia, as is the case in Vdr-null
with epithelia expressing epidermal markers, and epi- mice173. The block in hair follicle cycling occurs at the
dermal hyperplasia172,215. These results indicate different point of reinitiation of anagen following the develop-
roles for VDR and β-catenin in the stem cells of the hair ment cycle228. This loss of hair follicle cycling is specific
follicle, sebaceous gland and interfollicular epidermis216. for deletion of Vdr in the keratinocyte229, as demon-
β-Catenin interacts with the AF2 domain of VDR in a strated in a conditional mouse knockout model that
ligand-dependent manner (like other co-activators) but had none of the metabolic changes seen in the global
lymphoid enhancer binding factor 1 (LEF1; an impor- Vdr-knockout models.
tant transcription factor in β-catenin signalling) binds to There are, however, mutations in other signal-
the N terminus of VDR in a ligand-independent man- ling mechanisms that are phenocopies of the alope-
ner217 possibly essential for hair follicle cycling by these cia in Vdr-null mice (without the metabolic changes).
transcription factors. The significance of these inter­ These mutations provide insight into the role of VDR
actions are discussed further later in the Review where in its regulation of hair follicle cycling. Mutations in

Nature Reviews | Endocrinology


Reviews

Lamellar body transcription factors GLI1 and GLI2, whereas β-catenin


Keratohyalin signalling involves LEF1. The expression of GLI1 and
granule GLI2 is reduced during the latter stages of catagen in
both the Hr- and VDR-null hair follicle171. On the other
Stratum corneum hand, expression of Hedgehog signalling components
Cornified envelope
is increased in the interfollicular epidermis and utricles
SRC3 of the Vdr-null mouse240, but their expression is blocked
Stratum granulosum
Profilaggrin and/or by 1,25(OH)2D229. These observations emphasize the
loricrin, cathelicidin differences in signalling between the cycling portion of
and lipid production the hair follicle and that in the non-cycling portion
of the hair follicle and epidermis. As noted previously, in
the Vdr-null mouse as well as in Hr-mutant and Ctnnb1-
Stratum spinosum
INV and/or TG or mutant animals222,233,238,241, large lipid-filled dermal cysts
K1 and/or K10 develop that contain markers of the differentiated inter-
follicular epidermis along with hypertrophy of the seba-
MED1 ceous glands215,238,242. These findings suggest that changes
in cell fate of the stem cells within these niches differ
Stratum basale from those in the bulge.
• K5 and/or K14
• VDR, CYP27B1 Alopecia areata, as opposed to the alopecia totalis
• Stem cells found in Vdr-null mice (and humans), is often accom-
panied by vitamin D deficiency, and data show that it can
be improved by vitamin D supplementation243,244. This
Fig. 4 | The four layers of the epidermis, their distinct functions and location of condition, however, is thought to be due to immune
major Vdr co-regulators. The epidermis is composed of four layers. The stem cells are destruction of the hair follicle, with 1,25(OH)2D serv-
found in the stratum basale, where the keratins K5 and K14 are expressed. Here are also ing to suppress this process similar to the mechanism
the highest concentrations of vitamin D receptor (VDR) and CYP27B1. As the progeny in other inflammatory diseases of the skin such as pso-
of the stem cells move up from the basal layer they become part of the stratum spinosum. riasis245. A transient disruption of hair follicle cycling
Here, an important component of the cornified envelope, involucrin (INV), is produced was observed in S100g-knockout mice raised to wean-
along with transglutaminase (TG) that crosslinks INV and other proteins into the insoluble ing by mothers on a vitamin D-deficient low-calcium
cornified envelope. The main keratins expressed in the stratum spinosum are K5 and K10.
diet. Normal hair follicle cycling was restored after
The stratum granulosum lies just under the stratum corneum and expresses a number of
proteins such as loricrin (in the keratohyalin vesicles) that also form part of the cornified weaning246. The explanation for this phenomenon is not
envelope, and profilaggrin, that originally was thought to bundle the keratins but also clear, but might point to a role of calcium in hair folli-
provides peptides for incorporation into the lipid envelope lying beneath the cornified cle cycling not observed in mice lacking Casr in their
envelope. The stratum granulosum also expresses the enzymes for processing long-chain keratinocytes. Med1 deletion leads to an acceleration of
fatty acids that are incorporated into lamellar bodies along with defensins such as hair follicle cycling during the initial cycles, opposite to
cathelicidin that are secreted into the lipid envelope and cornified envelope to both the effect of Vdr deletion, but hair follicle differentiation
‘waterproof’ the cornified envelope and provide defence against invading organisms. is disrupted and, as is the case with Vdr deletion, Med1
1,25(OH)2D/VDR is a key regulator of proliferation of the stem cells and their subsequent deletion results in a gradual loss of bulge stem cells199.
differentiation as the cell moves from the stratum basale to the stratum corneum. This
change in 1,25(OH)2D/VDR-regulated gene expression during the differentiation process
Epidermal cancer development
is regulated by a shift in co-regulators of which the mediator complex is critical for
the early stages and steroid receptor co-activator 3 (SRC3) for the later stages of the Skin cancer is the most common form of cancer247. Most
differentiation process. MED1, mediator complex subunit 1. skin cancers are basal cell carcinomas (BCC; 80%) and
squamous cell carcinomas (SCC; 16%), with the more
deadly melanomas comprising about 4%. In individuals
Hairless (Hr) in both mice230 and humans231,232, as well with these cancers, circulating levels of 25OHD are not
as transcriptionally inactivating Ctnnb1 mutations in deficient and might even be increased248–250. BCC and
mice222,233, are two such examples. Hairless (Hr) binds SCC are currently referred to as keratinocyte carcinomas.
to VDR and in epidermal keratinocytes suppresses its Moreover, these cancer cells can produce 1,25(OH)2D
1,25(OH)2D-stimulated transcriptional activity234–236. The and 24,25(OH)2D, with rates of production that can be
role of Hr in ligand-independent VDR regulation of hair comparable to those of normal keratinocytes in some
follicle cycling is less clear. β-Catenin can enhance VDR cases188. However, when tested in vitro, keratinocyte car-
signalling especially in the hair follicle, whereas VDR can cinoma cells lack a normal response to either calcium or
inhibit β-catenin transcriptional activity in the epider- 1,25(OH)2D188,250,251, suggesting a mechanism other than
mis174,214,237. In mice with these different mutations that loss of VDR or CYP27B1 that is disrupted. As discussed
cause aborted hair follicle cycling, the pathological pro- previously, MED and SRC co-activator complexes regu­
cess begins during catagen at the end of the developmen- late the sequential actions of VDR and/or 1,25(OH)2D
tal cycle, the dermal papilla dissociates from the hair on keratinocyte proliferation and/or differentiation with
follicle, and anagen is not reinitiated172,230,238. Signalling the shift from MED to SRC being required for terminal
via the Hedgehog and β-catenin pathways is required for differentiation of the cell. The keratinocyte carcinoma
these mesenchymal and/or epithelial interactions, and cell lines we have studied overexpress MED1, with little
they interact with each other and with VDR to enable hair or no expression of SRC3, which suggests that this blocks
follicle cycling239. Hedgehog signalling is mediated by the the ability of VDR to promote the differentiation process

www.nature.com/nrendo
Reviews

by blocking access to SRC3 required for induction of the adult Vdr-null mice240, unlike the cycling part of the hair
more differentiated functions of the keratinocyte197. follicle in which the Hedgehog pathway is suppressed in
A second explanation stems from the reduction in Vdr-null mice. 1,25(OH)2D and/or VDR suppresses the
formation of the E-cadherin–catenin complex in kera­ expression of components of the Hedgehog pathway240,
tinocyte carcinomas. As noted previously, formation presumably through the VDR response elements found
of the E-cadherin–catenin complex is a key step in in their promoters174,217. Surprisingly, vitamin D by itself
keratinocyte differentiation. Moreover, failure to form seems to directly inhibit Smo, a key member of the
this complex, which would maintain β-catenin in the Hedgehog pathway7.
membrane, enables more β-catenin to translocate to
the nucleus, thereby promoting proliferation and Regulation of β-catenin signalling. Although the
bloc­king differentiation. Additionally, there is a shift tumours (pilomatricomas or trichofolliculomas) that
in the mode of activation of PLCγ1 from PIP3 (via the result from overexpression or activating mutations in the
E-cadherin–catenin complex), a mode of activation that β-catenin pathway are not malignant271, when combined
promotes the ability of PLCγ1 to maintain elevated intra- with the deletion of VDR, BCC develops174. This finding
cellular calcium levels required for differentiation, to its suggests that VDR restrains the potentially oncogenic
activation by growth factors such as epidermal growth actions of β-catenin via the mechanisms discussed above
factor, shifting its role to promoting proliferation252. in the section describing the interactions between VDR
Although these mechanisms may help explain the and β-catenin in keratinocytes. Moreover, in a number
lack of response to 1,25(OH)2D in cancers that express of genes response elements for both β-catenin and VDR
the VDR, lack of VDR predisposes to the development of have been found in close proximity174, which suggests
keratinocyte carcinomas in otherwise normal keratino­ dual regulation of these genes that in the absence of VDR
cytes. Mice null for Vdr develop tumours (papillomas might be oncogenic.
and keratinocyte carcinomas) at a much higher rate than
controls when exposed to carcinogens (7,12-dimethyl­ Regulation of long non-coding RNAs. Long non-coding
benzanthracene)253 or UVB radiation229,254. Moreover, RNAs (lncRNAs) are endogenous cellular RNAs larger
mice lacking both Vdr and Casr were found to develop than 200 bases that account for 80% of the transcriptome272.
tumours spontaneously255. That said, linking vitamin D They regulate many aspects of cellular physiology273,
signalling to keratinocyte carcinomas in humans has including cancer development. We found that Vdr dele-
been difficult, and the results of epidemiological studies tion in cultured keratinocytes and in the epidermis in vivo
have been mixed256,257. That said, several mechanisms, led to increased expression of oncogenic lncRNAs with
not mutually exclusive, underlying this predisposition to decreased expression of tumour suppressor lncRNAs274.
keratinocyte carcinoma formation at least in mice have
been proposed258, and are covered briefly here. Regulation of the hyaluronan–CD44 pathway.
Hyaluronan, a major glycosaminoglycan within the
Regulation of DNA damage response. UVB radiation extracellular matrix, binds to CD44 (ref.275), a function-
(280–320 nm) induces the formation of cyclobutane ally important membrane receptor found in most cells
pyrimidine dimers and pyrimidine primidone photo- including keratinocytes. CD44 is encoded by 19 exons,
products in the epidermis that if not repaired result in C but is expressed in numerous forms due to alternative
to T or CC to TT mutations — which are the UVB radi- splicing, in particular of exons 6–14 (ref.276). Differen­
ation signature lesions259. UVA radiation (320–400 nm) tiated keratinocytes express a more complete form of
damages DNA primarily by oxidative processes (such as CD44 (known as epican), whereas undifferentiated
8-hydroxy-2′-deoxyguanosine production). Reversing keratinocytes, mouse skin after chronic UVB irradiation,
these lesions is the role of DNA damage response oper- and SCCs express a variety of shorter CD44s that signal
ating primarily by nucleotide excision repair (NER)260,261. differently277.
This process is defective in the epidermis of Vdr-null Hyaluronic acid is synthesized by different hyal­
mice 229,262. In normal mice, topical application of uronic acid synthases and is degraded by different hyal­
1,25(OH)2D seems to be effective in preventing these uronidases. UVB radiation alters both hyaluronic acid
lesions and/or expediting their repair263. 1,25(OH)2D synthesis and its degradation 278. Large hyaluronic
induces two genes important for NER — XPC and DDB2 acid predominates in normal mouse skin, whereas small
(ref.264) — but in both cases VDR is required. hyaluronic acid predominates in cancer tissue279. Large
hyaluronic acid promotes transcriptional activation and
Regulation of the Hedgehog pathway. Overexpression of differentiation, whereas small hyaluronic acid induces
Hedgehog signalling due to mutations in PTCH1 or other the expression of pro-inflammatory cytokines and/or
genes of the Hedgehog pathway is a well-documented chemokines as well as cell proliferation and migration280.
cause of nearly all BCC265–267. The Hedgehog pathway We have proposed that the large hyaluronic acid acting
leads to the activation of the GLI family of transcription on epican promotes keratinocyte differentiation and
factors, which in turn increase the expression of each DNA repair through RAC–PKNγ and p38 MAPK sig-
other as well as PTCH1, and a variety of antiapoptotic nalling, whereas small hyaluronic acid acting on shorter
proproliferative factors while suppressing expression CD44 variants promotes proliferation and inflamma-
of proteins associated with keratinocyte differentia- tion through RhoA–ROK-dependent NF-κB–STAT3
tion268–270. This pathway is overexpressed in the epider- signalling281. 1,25(OH)2D blocks small hyaluronic acid–
mis and epidermal portion (utricles) of the hair follicle of CD44-mediated RhoA–ROK activation and NF-κB–p65

Nature Reviews | Endocrinology


Reviews

signalling as well as inflammatory gene expression and interfollicular epidermis and infundibulum of the hair
proliferation in transformed keratinocytes and SCC281. follicle rather than the bulge.
Oda and colleagues 169,297 observed delayed re-
Regulation of VDR levels and function. Tumours that epithelialization of wounds in keratinocyte-specific
are unresponsive to vitamin D have typically lost their Vdr-null mice on a low calcium diet or mice with com-
ability to produce 1,25(OH)2D (caused by decreased lev- bined Vdr and/or Casr deletion. Similar results were
els of CYP27B1)282,283. Furthermore, these tumours may seen in mice lacking Casr in their keratinocytes170.
have an increase in the metabolism of 1,25(OH)2D via These results were associated with reduced numbers
upregulation of CYP24A1 (ref.284), lost VDR transcrip- of stem cells, a reduction in their activation and migra-
tional activity through post-translational alterations tion over the newly formed matrix, and a subsequent
in RXR285, or a decrease in VDR expression284 — the failure of these cells to differentiate to re-form the epi-
last of which might be secondary to increased activity dermis. Others have noted a reduction in the number
in inhibitors of VDR expression such as SNAIL286 and and/or activation of bulge stem cells in global Vdr-null
SLUG287, increased methylation of the VDR promoter288, mice237,298, although not in the context of wounding.
or increased expression of miRNA125b (an inhibitor of In the studies conducted by Oda and colleagues169,297,
VDR expression)289. β-catenin signalling was reduced as was the formation
p53 cooperates with VDR in its antitumour func- of the E-cadherin–catenin complex in the migrating cells
tions290. The nuclear levels of p53 are increased by with failure of Vdr-null keratinocytes to participate in
1,25(OH)2D by non-genomic mechanisms291. Never­ the re-formation of the epidermis. Stem cell activation
theless, p53 mutations in tumours can also contribute requires β-catenin signalling216 and, probably, calcium
to resistance to vitamin D in tumours. The E3 ubiqui- signalling, whereas migration and/or differentiation
tin ligase and transcriptional regulator MDM2 is also requires the E-cadherin–catenin complex201. In studies
regulated by 1,25(OH)2D292, and unlike p53 MDM2 of global Vdr-null mice on a high-calcium rescue diet,
inhibits VDR and transcriptional activity293, as well as Demay and colleagues299 observed that TGF signalling
p53. Moreover, not all actions of VDR are genomic. in the dermis was reduced, associated with a reduction in
As noted previously, the 1,25(OH)2D stimulation of p53 macrophage recruitment and granulation tissue forma-
is non-genomic. Analogues of 1,25(OH)2D that do not tion. Thus vitamin D and calcium signalling have major
have genomic activity protect the skin from the UVB roles in all aspects of wound healing.
radiation-induced formation of cyclobutane pyrimidine
dimers294. These rapid effects of 1,25(OH)2D and its Conclusion
non-genomic analogues are mediated by either or both In this Review we provide an up-to-date discussion of
VDR acting in the membrane and an unrelated receptor vitamin D metabolism, with a focus on lessons that can
known initially as MARRS, but also known as ERp57, be learned by studying mutations in the genes encod-
GRp58, ERp60 or PSIA3. ing enzymes that enable vitamin D production, acti-
The role of vitamin D in skin cancer is likely to be vation and catabolism. We also discuss the means by
multifaceted295. Given the definitive role that Hedgehog which vitamin D and its metabolites are transported
signalling has in BCC, the potential role that vitamin D from sites of production to sites of metabolism and
plays in suppressing this pathway is a compelling ration- biological action. This portion of the Review focuses
ale for its use as a therapeutic agent. Moreover, the on DBP with its numerous polymorphisms that bear on
potential role of vitamin D in promoting DNA damage its function. Within the cell HSP70 acts as a shuttle to
repair might provide a rationale for its use in the preven- first help unload the vitamin D metabolites (25OHD in
tion of UVB radiation-induced skin cancer (both SCC particular) from DBP and shuttle the metabolite to the
and BCC). That said, at this stage of our knowledge it is mitochondria for further processing, then to transport
premature to judge which mechanisms are likely to be the active metabolite 1,25(OH)2D into the nucleus for
most important in the antitumour actions of vitamin D. binding to VDR133.
It is VDR that has the critical biological role in medi-
Epidermal wound repair ating the actions of 1,25(OH)2D. These actions, how-
Wound healing is a multistage process. Inflammation ever, are cell- and tissue-specific, indicating that these
and activation of the innate immune system are actions of 1,25(OH)2D/VDR are tightly regulated by a
important early responses. Keratinocytes null for Vdr, number of factors including co-regulators and epige-
Cyp27b1 or Casr have blunted inflammatory and innate netic modifications of the DNA that dictate which genes
immune responses to wounding168,191. Wounding also can be induced (or suppressed).
leads to a rapid increase in intracellular calcium170 and Following a description of newer insights into vita-
β-catenin signalling169. In response to the cytokines min D metabolism, transport and molecular mecha-
and growth factors expressed during this initial stage, nisms of action, we then discuss an excellent model by
stem cells in the three niches of the hair follicle and which these actions can be illustrated in a non-skeletal
interfollicular epidermis are activated and proliferate, tissue — the skin, in particular the epidermis and hair
and their progeny migrate to re-epithelialize the wound follicle. These are regenerative organs in which the key
over the matrix re-formed by dermal fibroblasts296. cell, the keratinocyte, has all the apparatus needed to
Although stem cells from all three niches contribute produce, metabo­lize and respond to vitamin D and
to re-epithelialization, the majority of cells perma- its metabolites1. We examine the important roles of
nently repopulating the epidermis originate from the 1,25(OH)2D and VDR in epidermal differentiation, the

www.nature.com/nrendo
Reviews

mechanism by which VDR without its ligand controls remains just that — speculation. We conclude our dis-
hair follicle cycling, and the key co-regulators that dif- cussion by demonstrating two important outcomes of
ferentially regulate VDR action in the hair follicle and vitamin D signalling in the skin: cancer suppression and
epidermis. Although it is tempting to speculate that vita- wound healing.
min D produced locally in the skin has a greater effect
on these functions than vitamin D ingested in food, this Published online xx xx xxxx

1. Bikle, D. D. Vitamin D and the skin: physiology and 22. Henderson, C. M. et al. Vitamin D-binding protein 44. Cheng, J. B., Motola, D. L., Mangelsdorf, D. J. &
pathophysiology. Rev. Endocr. Metab. Disord. 13, deficiency and homozygous deletion of the GC gene. Russell, D. W. De-orphanization of cytochrome
3–19 (2012). N. Engl. J. Med. 380, 1150–1157 (2019). P450 2R1: a microsomal vitamin D 25-hydroxylase.
2. Jäpelt, R. B. & Jakob sen, J. Vitamin D in plants: 23. Wu, S. et al. Intracellular vitamin D binding proteins: J. Biol. Chem. 278, 38084–38093 (2003).
a review of occurrence, analysis, and biosynthesis. novel facilitators of vitamin D-directed transactivation. 45. Al Mutair, A. N., Nasrat, G. H. & Russell, D. W.
Front. Plant. Sci. 4, 136 (2013). Mol. Endocrinol. 14, 1387–1397 (2000). Mutation of the CYP2R1 vitamin D 25-hydroxylase in a
3. Tripkovic, L. et al. Comparison of vitamin D2 and 24. Khanal, R. & Nemere, I. Membrane receptors for Saudi Arabian family with severe vitamin D deficiency.
vitamin D3 supplementation in raising serum vitamin D metabolites. Crit. Rev. Eukaryot. Gene Expr. J. Clin. Endocrinol. Metab. 97, E2022–E2025 (2012).
25-hydroxyvitamin D status: a systematic review and 17, 31–47 (2007). 46. Molin, A. et al. Vitamin D-dependent rickets type 1B
meta-analysis. Am. J. Clin. Nutr. 95, 1357–1364 25. Pike, J. W., Meyer, M. B. & Bishop, K. A. Regulation of (25-hydroxylase deficiency): a rare condition or a
(2012). target gene expression by the vitamin D receptor - an misdiagnosed condition? J. Bone Miner. Res. 32,
4. Tian, X. Q. & Holick, M. F. Catalyzed thermal update on mechanisms. Rev. Endocr. Metab. Disord. 1893–1899 (2017).
isomerization between previtamin D3 and vitamin D3 13, 45–55 (2012). 47. Gupta, R. P., Hollis, B. W., Patel, S. B., Patrick, K. S. &
via β-cyclodextrin complexation. J. Biol. Chem. 270, 26. Demay, M. B. The hair cycle and vitamin D receptor. Bell, N. H. CYP3A4 is a human microsomal vitamin D
8706–8711 (1995). Arch. Biochem. Biophys. 523, 19–21 (2012). 25-hydroxylase. J. Bone Miner. Res. 19, 680–688
5. Passeron, T. et al. Sunscreen photoprotection and 27. Meyer, M. B. et al. A kidney-specific genetic control (2004).
vitamin D status. Br. J. Dermatol. 181, 916–931 module in mice governs endocrine regulation of 48. Gupta, R. P., He, Y. A., Patrick, K. S., Halpert, J. R.
(2019). the cytochrome P450 gene Cyp27b1 essential for & Bell, N. H. CYP3A4 is a vitamin D-24- and
6. Prabhu, A. V., Luu, W., Li, D., Sharpe, L. J. & vitamin D3 activation. J. Biol. Chem. 292, 25-hydroxylase: analysis of structure function by
Brown, A. J. DHCR7: a vital enzyme switch between 17541–17558 (2017). site-directed mutagenesis. J. Clin. Endocrinol. Metab.
cholesterol and vitamin D production. Prog. Lipid Res. 28. Meyer, M. B. et al. A chromatin-based mechanism 90, 1210–1219 (2005).
64, 138–151 (2016). controls differential regulation of the cytochrome 49. Roizen, J. D. et al. CYP3A4 mutation causes vitamin
7. Bijlsma, M. F. et al. Repression of smoothened P450 gene Cyp24a1 in renal and non-renal tissues. D-dependent rickets type 3. J. Clin. Invest. 128,
by patched-dependent (pro-)vitamin D3 secretion. J. Biol. Chem. 294, 14467–14481 (2019). 1913–1918 (2018).
PLOS Biol. 4, e232 (2006). 29. Meyer, M. B., Benkusky, N. A., Lee, C. H. & Pike, J. W. 50. Adams, J. S. et al. Regulation of the extrarenal
8. Bhattacharyya, M. H. & DeLuca, H. F. Subcellular Genomic determinants of gene regulation by CYP27B1-hydroxylase. J. Steroid Biochem. Mol. Biol.
location of rat liver calciferol-25-hydroxylase. 1,25-dihydroxyvitamin D3 during osteoblast- 144, 22–27 (2014).
Arch. Biochem. Biophys. 160, 58–62 (1974). lineage cell differentiation. J. Biol. Chem. 289, 51. Tebben, P. J., Singh, R. J. & Kumar, R. Vitamin
9. Horst, R. L., Littledike, E. T., Riley, J. L. & Napoli, J. L. 19539–19554 (2014). D-mediated hypercalcemia: mechanisms, diagnosis,
Quantitation of vitamin D and its metabolites and 30. Malloy, P. J. et al. Vitamin D receptor mutations in and treatment. Endocr. Rev. 37, 521–547 (2016).
their plasma concentrations in five species of animals. patients with hereditary 1,25-dihydroxyvitamin 52. Fraser, D. et al. Pathogenesis of hereditary vitamin-D-
Anal. Biochem. 116, 189–203 (1981). D-resistant rickets. Mol. Genet. Metab. 111, 33–40 dependent rickets — an inborn error of vitamin D
10. Zhu, J. G., Ochalek, J. T., Kaufmann, M., Jones, G. & (2014). metabolism involving defective conversion of
Deluca, H. F. CYP2R1 is a major, but not exclusive, 31. Bikle, D. D. Vitamin D metabolism and function in the 25-hydroxyvitamin D to 1α,25-dihydroxyvitamin D.
contributor to 25-hydroxyvitamin D production skin. Mol. Cell. Endocrinol. 347, 80–89 (2011). N. Engl. J. Med. 289, 817–822 (1973).
in vivo. Proc. Natl Acad. Sci. USA 110, 15650–15655 32. Barrea, L. et al. Vitamin D and its role in psoriasis: 53. Scriver, C. R., Reade, T. M., DeLuca, H. F. &
(2013). an overview of the dermatologist and nutritionist. Hamstra, A. J. Serum 1,25-dihydroxyvitamin D levels
11. Cheng, J. B., Levine, M. A., Bell, N. H., Rev. Endocr. Metab. Disord. 18, 195–205 (2017). in normal subjects and in patients with hereditary
Mangelsdorf, D. J. & Russell, D. W. Genetic evidence 33. Mitsche, M. A., McDonald, J. G., Hobbs, H. H. & rickets or bone disease. N. Engl. J. Med. 299,
that the human CYP2R1 enzyme is a key vitamin D Cohen, J. C. Flux analysis of cholesterol biosynthesis 976–979 (1978).
25-hydroxylase. Proc. Natl Acad. Sci. USA 101, in vivo reveals multiple tissue and cell-type specific 54. Delvin, E. E., Glorieux, F. H., Marie, P. J. & Pettifor, J. M.
7711–7715 (2004). pathways. eLife 4, e07999 (2015). Vitamin D dependency: replacement therapy with
12. Bikle, D. D., Patzek, S. & Wang, Y. Physiologic and 34. Prabhu, A. V., Luu, W., Sharpe, L. J. & Brown, A. J. calcitriol? J. Pediatr. 99, 26–34 (1981).
pathophysiologic roles of extra renal CYP27b1: case Cholesterol-mediated degradation of 55. St-Arnaud, R., Messerlian, S., Moir, J. M., Omdahl, J. L.
report and review. Bone Rep. 8, 255–267 (2018). 7-dehydrocholesterol reductase switches the balance & Glorieux, F. H. The 25-hydroxyvitamin D
13. Thacher, T. D., Fischer, P. R., Singh, R. J., Roizen, J. & from cholesterol to vitamin D synthesis. J. Biol. Chem. 1-alpha-hydroxylase gene maps to the pseudovitamin
Levine, M. A. CYP2R1 mutations impair generation of 291, 8363–8373 (2016). D-deficiency rickets (PDDR) disease locus. J. Bone
25-hydroxyvitamin D and cause an atypical form of 35. Prabhu, A. V., Luu, W., Sharpe, L. J. & Brown, A. J. Miner. Res. 12, 1552–1559 (1997).
vitamin D deficiency. J. Clin. Endocrinol. Metab. 100, Phosphorylation regulates activity of 56. Takeyama, K. et al. 25-Hydroxyvitamin D3
E1005–E1013 (2015). 7-dehydrocholesterol reductase (DHCR7), a terminal 1alpha-hydroxylase and vitamin D synthesis.
14. Fu, G. K. et al. Cloning of human 25-hydroxyvitamin enzyme of cholesterol synthesis. J. Steroid Biochem. Science 277, 1827–1830 (1997).
D-1α-hydroxylase and mutations causing vitamin Mol. Biol. 165, 363–368 (2017). 57. Monkawa, T. et al. Molecular cloning of cDNA and
D-dependent rickets type 1. Mol. Endocrinol. 11, 36. Tint, G. S. et al. Defective cholesterol biosynthesis genomic DNA for human 25-hydroxyvitamin
1961–1970 (1997). associated with the Smith-Lemli-Opitz syndrome. D31α-hydroxylase. Biochem. Biophys. Res. Commun.
15. Schlingmann, K. P. et al. Mutations in CYP24A1 and N. Engl. J. Med. 330, 107–113 (1994). 239, 527–533 (1997).
idiopathic infantile hypercalcemia. N. Engl. J. Med. 37. Xu, G. et al. Reproducing abnormal cholesterol 58. Kim, C. J. et al. Vitamin D 1α-hydroxylase gene
365, 410–421 (2011). biosynthesis as seen in the Smith-Lemli-Opitz mutations in patients with 1α-hydroxylase deficiency.
16. Bikle, D. D., Siiteri, P. K., Ryzen, E. & Haddad, J. G. syndrome by inhibiting the conversion of J. Clin. Endocrinol. Metab. 92, 3177–3182 (2007).
Serum protein binding of 1,25-dihydroxyvitamin D: 7-dehydrocholesterol to cholesterol in rats. 59. Glorieux, F. H. & Pettifor, J. M. Vitamin D/dietary
a reevaluation by direct measurement of free J. Clin. Invest. 95, 76–81 (1995). calcium deficiency rickets and pseudo-vitamin D
metabolite levels. J. Clin. Endocrinol. Metab. 61, 38. Movassaghi, M., Bianconi, S., Feinn, R., Wassif, C. A. deficiency rickets. Bonekey Rep. 3, 524 (2014).
969–975 (1985). & Porter, F. D. Vitamin D levels in Smith-Lemli-Opitz 60. Bitzan, M. & Goodyer, P. R. Hypophosphatemic
17. Bikle, D. D. et al. Assessment of the free fraction syndrome. Am. J. Med. Genet. A 173, 2577–2583 rickets. Pediatr. Clin. North Am. 66, 179–207
of 25-hydroxyvitamin D in serum and its regulation (2017). (2019).
by albumin and the vitamin D-binding protein. 39. Ahn, J. et al. Genome-wide association study of 61. Xu, Y. et al. Intestinal and hepatic CYP3A4 catalyze
J. Clin. Endocrinol. Metab. 63, 954–959 (1986). circulating vitamin D levels. Hum. Mol. Genet. 19, hydroxylation of 1α,25-dihydroxyvitamin D3:
18. Malik, S. et al. Common variants of the vitamin D 2739–2745 (2010). implications for drug-induced osteomalacia.
binding protein gene and adverse health outcomes. 40. Wang, T. J. et al. Common genetic determinants of Mol. Pharmacol. 69, 56–65 (2006).
Crit. Rev. Clin. Lab. Sci. 50, 1–22 (2013). vitamin D insufficiency: a genome-wide association 62. Prosser, D. E., Kaufmann, M., O'Leary, B., Byford, V.
19. Bikle, D. D. & Schwartz, J. Vitamin D binding protein, study. Lancet 376, 180–188 (2010). & Jones, G. Single A326G mutation converts human
total and free vitamin D levels in different physiological 41. Kuan, V., Martineau, A. R., Griffiths, C. J., Hypponen, E. CYP24A1 from 25-OH-D3-24-hydroxylase into
and pathophysiological conditions. Front. Endocrinol. & Walton, R. DHCR7 mutations linked to higher -23-hydroxylase, generating 1α,25-(OH)2D3-26,
10, 317 (2019). vitamin D status allowed early human migration to 23-lactone. Proc. Natl Acad. Sci. USA 104,
20. Bikle, D. D., Malmstroem, S. & Schwartz, J. Current northern latitudes. BMC Evol. Biol. 13, 144 (2013). 12673–12678 (2007).
Controversies: are free vitamin metabolite levels a 42. Zhu, J. & DeLuca, H. F. Vitamin D 25-hydroxylase - 63. Jones, G., Prosser, D. E. & Kaufmann, M.
more accurate assessment of vitamin D status than four decades of searching, are we there yet? 25-Hydroxyvitamin D-24-hydroxylase (CYP24A1):
total levels? Endocrinol. Metab. Clin. North Am. 46, Arch. Biochem. Biophys. 523, 30–36 (2012). its important role in the degradation of vitamin D.
901–918 (2017). 43. Moghadasian, M. H. Cerebrotendinous Arch. Biochem. Biophys. 523, 9–18 (2012).
21. Safadi, F. F. et al. Osteopathy and resistance to xanthomatosis: clinical course, genotypes and 64. Brodie, M. J. et al. Effect of rifampicin and isoniazid
vitamin D toxicity in mice null for vitamin D binding metabolic backgrounds. Clin. Invest. Med. 27, 42–50 on vitamin D metabolism. Clin. Pharmacol. Ther. 32,
protein. J. Clin. Invest. 103, 239–251 (1999). (2004). 525–530 (1982).

Nature Reviews | Endocrinology


Reviews

65. Sakaki, T. et al. Dual metabolic pathway of cancer. J. Steroid Biochem. Mol. Biol. 39, 155–159 (Gc globulin) is related to the Gc phenotype in women.
25-hydroxyvitamin D3 catalyzed by human CYP24. (1991). Clin. Chem. 47, 753–756 (2001).
Eur. J. Biochem. 267, 6158–6165 (2000). 89. Moller, U. K., Streym, S., Heickendorff, L., Mosekilde, L. 112. Hoofnagle, A. N., Eckfeldt, J. H. & Lutsey, P. L.
66. Harant, H., Spinner, D., Reddy, G. S. & Lindley, I. J. & Rejnmark, L. Effects of 25OHD concentrations on Vitamin D-binding protein concentrations quantified
Natural metabolites of 1α,25-dihydroxyvitamin D3 chances of pregnancy and pregnancy outcomes: a by mass spectrometry. N. Engl. J. Med. 373,
retain biologic activity mediated through the vitamin D cohort study in healthy Danish women. Eur. J. Clin. 1480–1482 (2015).
receptor. J. Cell. Biochem. 78, 112–120 (2000). Nutr. 66, 862–868 (2012). 113. Carpenter, T. O. et al. Vitamin D binding protein is a
67. Martineau, C. et al. Optimal bone fracture repair 90. Zhang, J. Y., Lucey, A. J., Horgan, R., Kenny, L. C. & key determinant of 25-hydroxyvitamin D levels in
requires 24R,25-dihydroxyvitamin D3 and its effector Kiely, M. Impact of pregnancy on vitamin D status: infants and toddlers. J. Bone Miner. 28, 213–221
molecule FAM57B2. J. Clin. Invest. 128, 3546–3557 a longitudinal study. Br. J. Nutr. 112, 1081–1087 (2013).
(2018). (2014). 114. Santos, B. R., Mascarenhas, L. P., Boguszewski, M. C. &
68. Plachot, J. J. et al. In vitro action of 91. Moller, U. K. et al. Increased plasma concentrations Spritzer, P. M. Variations in the vitamin D-binding protein
1,25-dihydroxycholecalciferol and 24,25- of vitamin D metabolites and vitamin D binding (DBP) gene are related to lower 25-hydroxyvitamin D
dihydroxycholecalciferol on matrix organization and protein in women using hormonal contraceptives: levels in healthy girls: a cross-sectional study. Horm. Res.
mineral distribution in rabbit growth plate. a cross-sectional study. Nutrients 5, 3470–3480 Paediatr. 79, 162–168 (2013).
Metab. Bone Dis. Relat. Res. 4, 135–142 (1982). (2013). 115. Sollid, S. T. et al. Effects of vitamin D binding protein
69. St-Arnaud, R. et al. Deficient mineralization of 92. Guha, C., Osawa, M., Werner, P. A., Galbraith, R. M. & phenotypes and vitamin D supplementation on serum
intramembranous bone in vitamin D-24-hydroxylase- Paddock, G. V. Regulation of human Gc (vitamin D– total 25(OH)D and directly measured free 25(OH)D.
ablated mice is due to elevated 1,25-dihydroxyvitamin D binding) protein levels: hormonal and cytokine control of Eur. J. Endocrinol 174, 445–452 (2016).
and not to the absence of 24,25-dihydroxyvitamin D. gene expression in vitro. Hepatology 21, 1675–1681 116. Takiar, R. et al. The associations of 25-hydroxyvitamin D
Endocrinology 141, 2658–2666 (2000). (1995). levels, vitamin D binding protein gene polymorphisms,
70. Jones, G., Kottler, M. L. & Schlingmann, K. P. Genetic 93. Dahl, B. et al. Trauma stimulates the synthesis of and race with risk of incident fracture-related
diseases of vitamin D metabolizing enzymes. Gc-globulin. Intensive Care Med. 27, 394–399 hospitalization: twenty-year follow-up in a bi-ethnic
Endocrinol. Metab. Clin. North Am. 46, 1095–1117 (2001). cohort (the ARIC study). Bone 78, 94–101 (2015).
(2017). 94. Schiodt, F. V. Gc-globulin in liver disease. Dan. Med. 117. Leong, A. et al. The causal effect of vitamin D binding
71. Dinour, D. et al. Maternal and infantile hypercalcemia Bull. 55, 131–146 (2008). protein (DBP) levels on calcemic and cardiometabolic
caused by vitamin-D-hydroxylase mutations and 95. Wang, X., Shapses, S. A. & Al-Hraishawi, H. Free and diseases: a Mendelian randomization study.
vitamin D intake. Pediatr. Nephrol. 30, 145–152 bioavailable 25-hydroxyvitamin D levels in patients PLOS Med. 11, e1001751 (2014).
(2015). with primary hyperparathyroidism. Endocr. Pract. 23, 118. Hirai, M. et al. Group specific component protein
72. Shah, A. D. et al. Maternal hypercalcemia due to 66–71 (2017). genotype is associated with NIDDM in Japan.
failure of 1,25-dihydroxyvitamin-D3 catabolism in a 96. Song, Y. H., Ray, K., Liebhaber, S. A. & Cooke, N. E. Diabetologia 41, 742–743 (1998).
patient with CYP24A1 mutations. J. Clin. Endocrinol. Vitamin D-binding protein gene transcription is 119. Baier, L. J., Dobberfuhl, A. M., Pratley, R. E.,
Metab. 100, 2832–2836 (2015). regulated by the relative abundance of hepatocyte Hanson, R. L. & Bogardus, C. Variations in the vitamin
73. Tebben, P. J. et al. Hypercalcemia, hypercalciuria, and nuclear factors 1α and 1β. J. Biol. Chem. 273, D-binding protein (Gc locus) are associated with oral
elevated calcitriol concentrations with autosomal 28408–28418 (1998). glucose tolerance in nondiabetic Pima Indians. J. Clin.
dominant transmission due to CYP24A1 mutations: 97. Cleve, H. & Constans, J. The mutants of the vitamin-D- Endocrinol. Metab. 83, 2993–2996 (1998).
effects of ketoconazole therapy. J. Clin. Endocrinol. binding protein: more than 120 variants of the GC/ 120. Ye, W. Z., Dubois-Laforgue, D., Bellanne-Chantelot, C.,
Metab. 97, E423–E427 (2012). DBP system. Vox Sang. 54, 215–225 (1988). Timsit, J. & Velho, G. Variations in the vitamin
74. Cools, M. et al. Calcium and bone homeostasis in 98. Chun, R. F. New perspectives on the vitamin D binding D-binding protein (Gc locus) and risk of type 2
heterozygous carriers of CYP24A1 mutations: protein. Cell Biochem. Funct. 30, 445–456 (2012). diabetes mellitus in French Caucasians. Metabolism
a cross-sectional study. Bone 81, 89–96 (2015). 99. Nagasawa, H. et al. Gc protein (vitamin D-binding 50, 366–369 (2001).
75. Molin, A. et al. CYP24A1 mutations in a cohort of protein): Gc genotyping and GcMAF precursor activity. 121. Lauridsen, A. L. et al. Female premenopausal fracture
hypercalcemic patients: evidence for a recessive trait. Anticancer Res. 25, 3689–3695 (2005). risk is associated with Gc phenotype. J. Bone Miner.
J. Clin. Endocrinol. Metab. 100, E1343–E1352 (2015). 100. Uto, Y. et al. β-Galactosidase treatment is a common 19, 875–881 (2004).
76. Lightwood, R. & Stapleton, T. Idiopathic hypercalcaemia first-stage modification of the three major subtypes 122. Papiha, S. S., Allcroft, L. C., Kanan, R. M., Francis, R. M.
in infants. Lancet 265, 255–256 (1953). of Gc protein to GcMAF. Anticancer Res. 32, & Datta, H. K. Vitamin D binding protein gene in male
77. Pober, B. R. Williams-Beuren syndrome. N. Engl. 2359–2364 (2012). osteoporosis: association of plasma DBP and bone
J. Med. 362, 239–252 (2010). 101. Arnaud, J. & Constans, J. Affinity differences for mineral density with (TAAA)(n)-Alu polymorphism in
78. Chen, A. et al. Description of 5 novel SLC34A3/NPT2c vitamin D metabolites associated with the genetic DBP. Calcif. Tissue Int. 65, 262–266 (1999).
mutations causing hereditary hypophosphatemic isoforms of the human serum carrier protein (DBP). 123. Ezura, Y. et al. Association of molecular variants,
rickets with hypercalciuria. Kidney Int. Rep. 4, Hum. Genet. 92, 183–188 (1993). haplotypes, and linkage disequilibrium within the
1179–1186 (2019). 102. Bouillon, R., van Baelen, H. & de Moor, P. Comparative human vitamin D-binding protein (DBP) gene with
79. Bergwitz, C. & Miyamoto, K. I. Hereditary study of the affinity of the serum vitamin D-binding postmenopausal bone mineral density. J. Bone Miner.
hypophosphatemic rickets with hypercalciuria: protein. J. Steroid Biochem. 13, 1029–1034 (1980). Res. 18, 1642–1649 (2003).
pathophysiology, clinical presentation, diagnosis and 103. Boutin, B., Galbraith, R. M. & Arnaud, P. Comparative 124. Chishimba, L., Thickett, D. R., Stockley, R. A. &
therapy. Pflugers Arch. 471, 149–163 (2019). affinity of the major genetic variants of human Wood, A. M. The vitamin D axis in the lung: a key role
80. Kaufmann, M. et al. Clinical utility of simultaneous group-specific component (vitamin D-binding protein) for vitamin D-binding protein. Thorax 65, 456–462
quantitation of 25-hydroxyvitamin D and for 25-(OH) vitamin D. J. Steroid Biochem. 32, 59–63 (2010).
24,25-dihydroxyvitamin D by LC-MS/MS involving (1989). 125. Faserl, K. et al. Polymorphism in vitamin D-binding
derivatization with DMEQ-TAD. J. Clin. Endocrinol. 104. Jones, K. S. et al. 25(OH)D2 half-life is shorter than protein as a genetic risk factor in the pathogenesis
Metab. 99, 2567–2574 (2014). 25(OH)D3 half-life and is influenced by DBP of endometriosis. J. Clin. Endocrinol. Metab. 96,
81. Hirschfeld, J. Immune-electrophoretic demonstration concentration and genotype. J. Clin. Endocrinol. Metab. E233–E241 (2011).
of qualitative differences in human sera and their 99, 3373–3381 (2014). 126. Eloranta, J. J. et al. Association of a common vitamin
relation to the haptoglobins. Acta Pathol. Microbiol. 105. Chun, R. F. et al. Vitamin D-binding protein directs D-binding protein polymorphism with inflammatory
Scand. 47, 160–168 (1959). monocyte responses to 25-hydroxy- and bowel disease. Pharmacogenet Genomics 21, 559–564
82. Daiger, S. P., Schanfield, M. S. & Cavalli-Sforza, L. L. 1,25-dihydroxyvitamin D. J. Clin. Endocrinol. Metab. (2011).
Group-specific component (Gc) proteins bind vitamin D 95, 3368–3376 (2010). 127. Abbas, S. et al. The Gc2 allele of the vitamin D binding
and 25-hydroxyvitamin D. Proc. Natl Acad. Sci. USA 106. Bikle, D. D., Gee, E., Halloran, B. & Haddad, J. G. protein is associated with a decreased postmenopausal
72, 2076–2080 (1975). Free 1,25-dihydroxyvitamin D levels in serum from breast cancer risk, independent of the vitamin D status.
83. Armas, L. A., Hollis, B. W. & Heaney, R. P. Vitamin D2 normal subjects, pregnant subjects, and subjects Cancer Epidemiol. Biomarkers Prev. 17, 1339–1343
is much less effective than vitamin D3 in humans. with liver disease. J. Clin. Invest. 74, 1966–1971 (2008).
J. Clin. Endocrinol. Metab. 89, 5387–5391 (2004). (1984). 128. Dimopoulos, M. A. et al. Genetic markers in carcinoma
84. Mendel, C. M. The free hormone hypothesis: 107. Schwartz, J. B. et al. A comparison of measured and of the prostate. Eur. Urol. 10, 315–316 (1984).
a physiologically based mathematical model. calculated free 25(OH) vitamin D levels in clinical 129. Zhou, L. et al. GC Glu416Asp and Thr420Lys
Endocr. Rev. 10, 232–274 (1989). populations. J. Clin. Endocrinol. Metab. 99, polymorphisms contribute to gastrointestinal cancer
85. Nykjaer, A. et al. An endocytic pathway essential for 1631–1637 (2014). susceptibility in a Chinese population. Int. J. Clin.
renal uptake and activation of the steroid 25-(OH) 108. Bikle, D. D., Halloran, B. P., Gee, E., Ryzen, E. & Exp. Med. 5, 72–79 (2012).
vitamin D3. Cell 96, 507–515 (1999). Haddad, J. G. Free 25-hydroxyvitamin D levels are 130. Poynter, J. N. et al. Genetic variation in the vitamin D
86. Zella, L. A., Shevde, N. K., Hollis, B. W., Cooke, N. E. normal in subjects with liver disease and reduced receptor (VDR) and the vitamin D-binding protein (GC)
& Pike, J. W. Vitamin D-binding protein influences total 25-hydroxyvitamin D levels. J. Clin. Invest. 78, and risk for colorectal cancer: results from the Colon
total circulating levels of 1,25-dihydroxyvitamin D3 748–752 (1986). Cancer Family Registry. Cancer Epidemiol. Biomarkers
but does not directly modulate the bioactive levels of 109. Schwartz, J. B. et al. Determination of free 25(OH)D Prev. 19, 525–536 (2010).
the hormone in vivo. Endocrinology 149, 3656–3667 concentrations and their relationships to total 25(OH) 131. Martineau, A. R. et al. Association between Gc
(2008). D in multiple clinical populations. J. Clin. Endocrinol. genotype and susceptibility to TB is dependent on
87. Cooke, N. E., McLeod, J. F., Wang, X. K. & Ray, K. Metab. 103, 3278–3288 (2018). vitamin D status. Eur. Respir. J. 35, 1106–1112 (2010).
Vitamin D binding protein: genomic structure, 110. Schwartz, J. B., Kane, L. & Bikle, D. Response of 132. Gacad, M. A., Chen, H., Arbelle, J. E., LeBon, T.
functional domains, and mRNA expression in tissues. vitamin D concentration to vitamin D3 administration & Adams, J. S. Functional characterization and
J. Steroid Biochem. Mol. Biol. 40, 787–793 (1991). in older adults without sun exposure: a randomized purification of an intracellular vitamin D-binding
88. Hagenfeldt, Y., Carlstrom, K., Berlin, T. & Stege, R. double-blind trial. J. Am. Geriatr. Soc. 64, 65–72 protein in vitamin D-resistant New World primate
Effects of orchidectomy and different modes of high (2016). cells. Amino acid sequence homology with proteins
dose estrogen treatment on circulating “free” and total 111. Lauridsen, A. L., Vestergaard, P. & Nexo, E. Mean in the hsp-70 family. J. Biol. Chem. 272, 8433–8440
1,25-dihydroxyvitamin D in patients with prostatic serum concentration of vitamin D-binding protein (1997).

www.nature.com/nrendo
Reviews

133. Adams, J. S. et al. Response element binding proteins Transcription: epigenetic modification involving 178. Thacher, S. M. & Rice, R. H. Keratinocyte-specific
and intracellular vitamin D binding proteins: novel cross-talk between protein-arginine methyltransferase transglutaminase of cultured human epidermal cells:
regulators of vitamin D trafficking, action and 5 and the SWI/SNF complex. J. Biol. Chem. 289, relation to cross-linked envelope formation and
metabolism. J. Steroid Biochem. Mol. Biol. 89-90, 33958–33970 (2014). terminal differentiation. Cell 40, 685–695 (1985).
461–465 (2004). 157. Meyer, M. B., Benkusky, N. A. & Pike, J. W. 179. Steven, A. C., Bisher, M. E., Roop, D. R. & Steinert, P. M.
134. Brumbaugh, P. F. & Haussler, M. R. 1α,25- The RUNX2 cistrome in osteoblasts: characterization, Biosynthetic pathways of filaggrin and loricrin–two
Dihydroxycholecalciferol receptors in intestine. down-regulation following differentiation, and major proteins expressed by terminally differentiated
II. Temperature-dependent transfer of the hormone to relationship to gene expression. J. Biol. Chem. 289, epidermal keratinocytes. J. Struct. Biol. 104, 150–162
chromatin via a specific cytosol receptor. J. Biol. Chem. 16016–16031 (2014). (1990).
249, 1258–1262 (1974). 158. Pike, J. W., Meyer, M. B., St John, H. C. & 180. Freeman, S. C. & Sonthalia, S. Histology, keratohyalin
135. Wang, Y., Zhu, J. & DeLuca, H. F. Where is the vitamin D Benkusky, N. A. Epigenetic histone modifications granules. StatPearls https://www.ncbi.nlm.nih.gov/
receptor? Arch. Biochem. Biophys. 523, 123–133 and master regulators as determinants of context books/NBK537049/ (2019).
(2012). dependent nuclear receptor activity in bone cells. 181. Dale, B. A., Resing, K. A. & Lonsdale-Eccles, J. D.
136. Christakos, S., Dhawan, P., Verstuyf, A., Verlinden, L. Bone 81, 757–764 (2015). Filaggrin: a keratin filament associated protein.
& Carmeliet, G. Vitamin D: metabolism, molecular 159. Ding, N. et al. A vitamin D receptor/SMAD genomic Ann. NY Acad. Sci. 455, 330–342 (1985).
mechanism of action, and pleiotropic effects. circuit gates hepatic fibrotic response. Cell 153, 182. Levin, J., Friedlander, S. F. & Del Rosso, J. Q. Atopic
Physiol. Rev. 96, 365–408 (2016). 601–613 (2013). dermatitis and the stratum corneum: part 2: other
137. Gunton, J. E. & Girgis, C. M. Vitamin D and muscle. 160. Sherman, M. H. et al. Vitamin D receptor-mediated structural and functional characteristics of the stratum
Bone Rep. 8, 163–167 (2018). stromal reprogramming suppresses pancreatitis and corneum barrier in atopic skin. J. Clin. Aesthet.
138. Zheng, J. et al. HDX reveals the conformational enhances pancreatic cancer therapy. Cell 159, 80–93 Dermatol. 6, 49–54 (2013).
dynamics of DNA sequence specific VDR co-activator (2014). 183. Mehrel, T. et al. Identification of a major keratinocyte
interactions. Nat. Commun. 8, 923 (2017). 161. Wei, Z. et al. Vitamin D switches BAF complexes to cell envelope protein, loricrin. Cell 61, 1103–1112
139. Shaffer, P. L. & Gewirth, D. T. Structural basis of protect β cells. Cell 173, 1135–1149.e15 (2018). (1990).
VDR-DNA interactions on direct repeat response 162. Bikle, D. D., Nemanic, M. K., Whitney, J. O. & Elias, P. W. 184. Menon, G. K., Lee, S. E. & Lee, S. H. An overview of
elements. EMBO J. 21, 2242–2252 (2002). Neonatal human foreskin keratinocytes produce epidermal lamellar bodies: novel roles in biological
140. Rochel, N., Wurtz, J. M., Mitschler, A., Klaholz, B. & 1,25-dihydroxyvitamin D3. Biochemistry 25, adaptations and secondary barriers. J. Dermatol. Sci.
Moras, D. The crystal structure of the nuclear receptor 1545–1548 (1986). 92, 10–17 (2018).
for vitamin D bound to its natural ligand. Mol. Cell 5, 163. Stumpf, W. E., Sar, M., Reid, F. A., Tanaka, Y. & 185. Feingold, K. R. & Elias, P. M. Role of lipids in the
173–179 (2000). DeLuca, H. F. Target cells for 1,25-dihydroxyvitamin D3 formation and maintenance of the cutaneous
141. Carlberg, C., Molnár, F. & Mouriño, A. Vitamin D in intestinal tract, stomach, kidney, skin, pituitary, and permeability barrier. Biochim. Biophys. Acta 1841,
receptor ligands: the impact of crystal structures. parathyroid. Science 206, 1188–1190 (1979). 280–294 (2014).
Expert. Opin. Ther. Pat. 22, 417–435 (2012). 164. Pillai, S., Bikle, D. D. & Elias, P. M. 1,25- 186. Aberg, K. M. et al. Co-regulation and interdependence
142. Mizwicki, M. T. & Norman, A. W. The vitamin D Dihydroxyvitamin D production and receptor binding of the mammalian epidermal permeability and
sterol-vitamin D receptor ensemble model offers in human keratinocytes varies with differentiation. antimicrobial barriers. J. Invest. Dermatol. 128,
unique insights into both genomic and rapid-response J. Biol. Chem. 263, 5390–5395 (1988). 917–925 (2008).
signaling. Sci. Signal. 2, re4 (2009). 165. Hosomi, J., Hosoi, J., Abe, E., Suda, T. & Kuroki, T. 187. Bikle, D. D. & Pillai, S. Vitamin D, calcium, and epidermal
143. Singarapu, K. K. et al. Ligand-specific structural Regulation of terminal differentiation of cultured differentiation. Endocr. Rev. 14, 3–19 (1993).
changes in the vitamin D receptor in solution. mouse epidermal cells by 1α,25-dihydroxyvitamin D3. 188. Bikle, D. D., Pillai, S. & Gee, E. Squamous carcinoma
Biochemistry 50, 11025–11033 (2011). Endocrinology 113, 1950–1957 (1983). cell lines produce 1,25 dihydroxyvitamin D, but fail to
144. Rochel, N. et al. Common architecture of nuclear 166. Smith, E. L., Walworth, N. C. & Holick, M. F. Effect of respond to its prodifferentiating effect. J. Invest.
receptor heterodimers on DNA direct repeat elements 1α,25-dihydroxyvitamin D3 on the morphologic and Dermatol. 97, 435–441 (1991).
with different spacings. Nat. Struct. Mol. Biol. 18, biochemical differentiation of cultured human 189. McLane, J. A., Katz, M. & Abdelkader, N. Effect of
564–570 (2011). epidermal keratinocytes grown in serum-free 1,25-dihydroxyvitamin D3 on human keratinocytes
145. Orlov, I., Rochel, N., Moras, D. & Klaholz, B. P. conditions. J. Invest. Dermatol. 86, 709–714 (1986). grown under different culture conditions. In Vitro Cell.
Structure of the full human RXR/VDR nuclear receptor 167. Oda, Y. et al. Vitamin D receptor and coactivators Dev. Biol. 26, 379–387 (1990).
heterodimer complex with its DR3 target DNA. SRC2 and 3 regulate epidermis-specific sphingolipid 190. Hawker, N. P., Pennypacker, S. D., Chang, S. M.
EMBO J. 31, 291–300 (2012). production and permeability barrier formation. & Bikle, D. D. Regulation of human epidermal
146. Zhang, J. et al. DNA binding alters coactivator J. Invest. Dermatol. 129, 1367–1378 (2009). keratinocyte differentiation by the vitamin D receptor
interaction surfaces of the intact VDR-RXR complex. 168. Tu, C. L. et al. Ablation of the calcium-sensing receptor and its coactivators DRIP205, SRC2, and SRC3.
Nat. Struct. Mol. Biol. 18, 556–563 (2011). in keratinocytes impairs epidermal differentiation and J. Invest. Dermatol. 127, 874–880 (2007).
147. Feldman, D. & Malloy, P. J. Mutations in the vitamin D barrier function. J. Invest. Dermatol. 132, 2350–2359 191. Schauber, J. et al. Injury enhances TLR2 function and
receptor and hereditary vitamin D-resistant rickets. (2012). antimicrobial peptide expression through a vitamin
Bonekey Rep. 3, 510 (2014). 169. Oda, Y. et al. Vitamin D receptor is required for D-dependent mechanism. J. Clin. Invest. 117, 803–811
148. Lee, S. M., Goellner, J. J., O'Brien, C. A. & Pike, J. W. proliferation, migration, and differentiation of (2007).
A humanized mouse model of hereditary epidermal stem cells and progeny during cutaneous 192. Schauber, J., Dorschner, R. A., Yamasaki, K., Brouha, B.
1,25-dihydroxyvitamin D-resistant rickets without wound repair. J. Invest. Dermatol. 138, 2423–2431 & Gallo, R. L. Control of the innate epithelial
alopecia. Endocrinology 155, 4137–4148 (2014). (2018). antimicrobial response is cell-type specific and
149. Huet, T. et al. A vitamin D receptor selectively activated 170. Tu, C. L., Celli, A., Mauro, T. & Chang, W. The dependent on relevant microenvironmental stimuli.
by gemini analogs reveals ligand dependent and calcium-sensing receptor regulates epidermal Immunology 118, 509–519 (2006).
independent effects. Cell Rep. 10, 516–526 (2015). intracellular Ca(2+) signaling and re-epithelialization 193. Oda, Y. et al. Vitamin D receptor and coactivators
150. Lee, S. M. & Pike, J. W. The vitamin D receptor after wounding. J. Invest. Dermatol. 139, 919–929 SRC 2 and 3 regulate epidermis-specific sphingolipid
functions as a transcription regulator in the absence of (2019). production and permeability barrier formation.
1,25-dihydroxyvitamin D3. J. Steroid Biochem. Mol. 171. Teichert, A., Elalieh, H. & Bikle, D. Disruption of the J. Invest. Dermatol. 129, 1367–1378 (2009).
Biol. 164, 265–270 (2016). hedgehog signaling pathway contributes to the 194. Muehleisen, B. et al. PTH/PTHrP and vitamin D
151. Balsan, S. et al. Long-term nocturnal calcium infusions hair follicle cycling deficiency in Vdr knockout mice. control antimicrobial peptide expression and
can cure rickets and promote normal mineralization J. Cell. Physiol. 225, 482–489 (2010). susceptibility to bacterial skin infection. Sci. Transl
in hereditary resistance to 1,25-dihydroxyvitamin D. 172. Bikle, D. D., Elalieh, H., Chang, S., Xie, Z. & Med. 4, 135ra66 (2012).
J. Clin. Invest. 77, 1661–1667 (1986). Sundberg, J. P. Development and progression of 195. Zehnder, D. et al. Extrarenal expression of
152. Chaturvedi, D., Garabedian, M., Carel, J. C. & Leger, J. alopecia in the vitamin D receptor null mouse. J. Cell. 25-hydroxyvitamin D3-1α-hydroxylase. J. Clin.
Different mechanisms of intestinal calcium absorption Physiol. 207, 340–353 (2006). Endocrinol. Metab. 86, 888–894 (2001).
at different life stages: therapeutic implications and 173. Li, Y. C. et al. Targeted ablation of the vitamin D 196. Stumpf, W. E., Clark, S. A., Sar, M. & DeLuca, H. F.
long-term responses to treatment in patients with receptor: an animal model of vitamin D-dependent Topographical and developmental studies on target
hereditary vitamin D-resistant rickets. Horm. Res. rickets type II with alopecia. Proc. Natl Acad. Sci. USA sites of 1,25 (OH)2 vitamin D3 in skin. Cell Tissue Res.
Paediatr. 78, 326–331 (2012). 94, 9831–9835 (1997). 238, 489–496 (1984).
153. Pike, J. W. & Christakos, S. Biology and mechanisms of 174. Palmer, H. G., Anjos-Afonso, F., Carmeliet, G., 197. Oda, Y. et al. Two distinct coactivators, DRIP/mediator
action of the vitamin D hormone. Endocrinol. Metab. Takeda, H. & Watt, F. M. The vitamin D receptor is a and SRC/p160, are differentially involved in vitamin D
Clin. North Am. 46, 815–843 (2017). Wnt effector that controls hair follicle differentiation receptor transactivation during keratinocyte
154. Pike, J. W., Meyer, M. B., Lee, S. M., Onal, M. & and specifies tumor type in adult epidermis. differentiation. Mol. Endocrinol. 17, 2329–2339
Benkusky, N. A. The vitamin D receptor: contemporary PLOS ONE 3, e1483 (2008). (2003).
genomic approaches reveal new basic and 175. Kretzschmar, K. & Watt, F. M. Markers of epidermal 198. Oda, Y., Ishikawa, M. H., Hawker, N. P., Yun, Q. C. &
translational insights. J. Clin. Invest. 127, 1146–1154 stem cell subpopulations in adult mammalian skin. Bikle, D. D. Differential role of two VDR coactivators,
(2017). Cold Spring Harb. Perspect. Med. 4, a013631 DRIP205 and SRC-3, in keratinocyte proliferation and
155. Wei, R., Dhawan, P., Baiocchi, R. A., Kim, K. Y. & (2014). differentiation. J. Steroid Biochem. Mol. Biol. 103,
Christakos, S. PU.1 and epigenetic signals modulate 176. Eichner, R., Sun, T. T. & Aebi, U. The role of keratin 776–780 (2007).
1,25-dihydroxyvitamin D3 and C/EBPα regulation subfamilies and keratin pairs in the formation of 199. Oda, Y. et al. Coactivator MED1 ablation in
of the human cathelicidin antimicrobial peptide human epidermal intermediate filaments. J. Cell Biol. keratinocytes results in hair-cycling defects and
gene in lung epithelial cells. J. Cell Physiol. 234, 102, 1767–1777 (1986). epidermal alterations. J. Invest. Dermatol. 132,
10345–10359 (2019). 177. Warhol, M. J., Roth, J., Lucocq, J. M., Pinkus, G. S. 1075–1083 (2012).
156. Seth-Vollenweider, T., Joshi, S., Dhawan, P., Sif, S. & Rice, R. H. Immuno-ultrastructural localization of 200. Schauber, J. et al. Histone acetylation in keratinocytes
& Christakos, S. Novel mechanism of negative involucrin in squamous epithelium and cultured enables control of the expression of cathelicidin
regulation of 1,25-dihydroxyvitamin D3-induced keratinocytes. J. Histochem. Cytochem. 33, 141–149 and CD14 by 1,25-dihydroxyvitamin D3. J. Invest.
25-hydroxyvitamin D3 24-hydroxylase (Cyp24a1) (1985). Dermatol. 128, 816–824 (2008).

Nature Reviews | Endocrinology


Reviews

201. Bikle, D. D., Xie, Z. & Tu, C. L. Calcium regulation of 224. Schneider, M. R., Schmidt-Ullrich, R. & Paus, R. The 248. Reichrath, J. et al. Analysis of 1,25-dihydroxyvitamin D(3)
keratinocyte differentiation. Expert. Rev. Endocrinol. hair follicle as a dynamic miniorgan. Curr. Biol. 19, receptors (VDR) in basal cell carcinomas. Am. J. Pathol.
Metab. 7, 461–472 (2012). R132–R142 (2009). 155, 583–589 (1999).
202. Tu, C. L., Chang, W. & Bikle, D. D. The extracellular 225. Gonzales, K. A. U. & Fuchs, E. Skin and its 249. Reichrath, J. et al. Analysis of the vitamin D system in
calcium-sensing receptor is required for regenerative powers: an alliance between stem cells cutaneous squamous cell carcinomas. J. Cutan. Pathol.
calcium-induced differentiation in human and their niche. Dev. Cell 43, 387–401 (2017). 31, 224–231 (2004).
keratinocytes. J. Biol. Chem. 276, 41079–41085 226. Hochberg, Z. et al. Calcitriol-resistant rickets with 250. Ratnam, A. V., Bikle, D. D., Su, M. J. & Pillai, S.
(2001). alopecia. Arch. Dermatol. 121, 646–647 (1985). Squamous carcinoma cell lines fail to respond to
203. Tu, C. L., Chang, W. & Bikle, D. D. Phospholipase Cγ1 227. Marx, S. J., Bliziotes, M. M. & Nanes, M. Analysis of 1,25-dihydroxyvitamin D despite normal levels
is required for activation of store-operated channels the relation between alopecia and resistance to of the vitamin D receptor. J. Invest. Dermatol. 106,
in human keratinocytes. J. Invest. Dermatol. 124, 1,25-dihydroxyvitamin D. Clin. Endocrinol. 25, 522–525 (1996).
187–197 (2005). 373–381 (1986). 251. Rheinwald, J. G. & Beckett, M. A. Defective terminal
204. Tu, C., Chang, W., Xie, Z. & Bikle, D. Inactivation of the 228. Sakai, Y. & Demay, M. B. Evaluation of keratinocyte differentiation in culture as a consistent and selectable
calcium sensing receptor inhibits E-cadherin-mediated proliferation and differentiation in vitamin D receptor character of malignant human keratinocytes. Cell 22,
cell-cell adhesion and calcium-induced differentiation knockout mice. Endocrinology 141, 2043–2049 629–632 (1980).
in human epidermal keratinocytes. J. Biol. Chem. 283, (2000). 252. Xie, Z. et al. Phospholipase C-γ1 is required for the
3519–3528 (2008). 229. Teichert, A. E., Elalieh, H., Elias, P. M., Welsh, J. & epidermal growth factor receptor-induced squamous
205. Xie, Z., Singleton, P. A., Bourguignon, L. Y. & Bikle, D. D. Bikle, D. D. Overexpression of hedgehog signaling is cell carcinoma cell mitogenesis. Biochem. Biophys.
Calcium-induced human keratinocyte differentiation associated with epidermal tumor formation in vitamin Res. Commun. 397, 296–300 (2010).
requires src- and fyn-mediated phosphatidylinositol D receptor-null mice. J. Invest. Dermatol. 131, 253. Zinser, G. M., Sundberg, J. P. & Welsh, J. Vitamin D(3)
3-kinase-dependent activation of phospholipase C-γ1. 2289–2297 (2011). receptor ablation sensitizes skin to chemically induced
Mol. Biol. Cell 16, 3236–3246 (2005). 230. Panteleyev, A. A., Botchkareva, N. V., Sundberg, J. P., tumorigenesis. Carcinogenesis 23, 2103–2109 (2002).
206. Xie, Z. & Bikle, D. D. The recruitment of Christiano, A. M. & Paus, R. The role of the hairless 254. Ellison, T. I., Smith, M. K., Gilliam, A. C. &
phosphatidylinositol 3-kinase to the E-cadherin- (hr) gene in the regulation of hair follicle catagen Macdonald, P. N. Inactivation of the vitamin D
catenin complex at the plasma membrane is required transformation. Am. J. Pathol. 155, 159–171 (1999). receptor enhances susceptibility of murine skin to
for calcium-induced phospholipase C-γ1 activation and 231. Miller, J. et al. Atrichia caused by mutations in the UV-induced tumorigenesis. J. Invest. Dermatol. 128,
human keratinocyte differentiation. J. Biol. Chem. vitamin D receptor gene is a phenocopy of generalized 2508–2517 (2008).
282, 8695–8703 (2007). atrichia caused by mutations in the hairless gene. 255. Bikle, D. D., Oda, Y., Tu, C. L. & Jiang, Y. Novel
207. Xie, Z., Chang, S. M., Pennypacker, S. D., Liao, E. Y. & J. Invest. Dermatol. 117, 612–617 (2001). mechanisms for the vitamin D receptor (VDR) in the
Bikle, D. D. Phosphatidylinositol-4-phosphate 5-kinase 232. Ahmad, W. et al. Alopecia universalis associated with skin and in skin cancer. J. Steroid Biochem. Mol. Biol.
1α mediates extracellular calcium-induced keratinocyte a mutation in the human hairless gene. Science 279, 148, 47–51 (2015).
differentiation. Mol. Biol. Cell 20, 1695–1704 (2009). 720–724 (1998). 256. Tang, J. Y. et al. Vitamin D in cutaneous carcinogenesis:
208. Tu, C. L., Oda, Y., Komuves, L. & Bikle, D. D. The role 233. Millar, S. E. Molecular mechanisms regulating hair part I. J. Am. Acad. Dermatol. 67, 803.e1–803.e12
of the calcium-sensing receptor in epidermal follicle development. J. Invest. Dermatol. 118, (2012).
differentiation. Cell Calcium 35, 265–273 (2004). 216–225 (2002). 257. Tang, J. Y. et al. Vitamin D in cutaneous carcinogenesis:
209. Tu, C. L., Chang, W. & Bikle, D. D. The role of 234. Xie, Z., Chang, S., Oda, Y. & Bikle, D. D. Hairless part II. J. Am. Acad. Dermatol. 67, 817.e1–817.e11
the calcium sensing receptor in regulating intracellular suppresses vitamin D receptor transactivation in (2012).
calcium handling in human epidermal keratinocytes. human keratinocytes. Endocrinology 147, 314–323 258. Bikle, D. D., Jiang, Y., Nguyen, T., Oda, Y. & Tu, C. L.
J. Invest. Dermatol. 127, 1074–1083 (2007). (2006). Disruption of vitamin D and calcium signaling in
210. Canaff, L. & Hendy, G. N. Human calcium-sensing 235. Hsieh, J. C. et al. Analysis of hairless corepressor keratinocytes predisposes to skin cancer. Front.
receptor gene. Vitamin D response elements in mutants to characterize molecular cooperation with Physiol. 7, 296 (2016).
promoters P1 and P2 confer transcriptional the vitamin D receptor in promoting the mammalian 259. Hussein, M. R. Ultraviolet radiation and skin cancer:
responsiveness to 1,25-dihydroxyvitamin D. J. Biol. hair cycle. J. Cell. Biochem. 110, 671–686 (2010). molecular mechanisms. J. Cutan. Pathol. 32,
Chem. 277, 30337–30350 (2002). 236. Hsieh, J. C. et al. Physical and functional interaction 191–205 (2005).
211. Xie, Z. & Bikle, D. D. Cloning of the human between the vitamin D receptor and hairless 260. Chen, R. H., Maher, V. M. & McCormick, J. J. Effect of
phospholipase C-γ1 promoter and identification of a corepressor, two proteins required for hair cycling. excision repair by diploid human fibroblasts on the
DR6-type vitamin D-responsive element. J. Biol. Chem. J. Biol. Chem. 278, 38665–38674 (2003). kinds and locations of mutations induced by (+/-)-7
272, 6573–6577 (1997). 237. Cianferotti, L., Cox, M., Skorija, K. & Demay, M. B. beta,8 alpha-dihydroxy-9 alpha,10 alpha-epoxy-
212. Su, M. J., Bikle, D. D., Mancianti, M. L. & Pillai, S. Vitamin D receptor is essential for normal keratinocyte 7,8,9,10- tetrahydrobenzo[a]pyrene in the coding
1,25-Dihydroxyvitamin D3 potentiates the keratinocyte stem cell function. Proc. Natl Acad. Sci. USA 104, region of the HPRT gene. Proc. Natl Acad. Sci. USA
response to calcium. J. Biol. Chem. 269, 14723–14729 9428–9433 (2007). 87, 8680–8684 (1990).
(1994). 238. Huelsken, J., Vogel, R., Erdmann, B., Cotsarelis, G. 261. Wood, R. D. DNA damage recognition during
213. Gniadecki, R., Gajkowska, B. & Hansen, M. & Birchmeier, W. β-Catenin controls hair follicle nucleotide excision repair in mammalian cells.
1,25-Dihydroxyvitamin D3 stimulates the assembly of morphogenesis and stem cell differentiation in the Biochimie 81, 39–44 (1999).
adherens junctions in keratinocytes: involvement of skin. Cell 105, 533–545 (2001). 262. Demetriou, S. K. et al. Vitamin D receptor mediates
protein kinase C. Endocrinology 138, 2241–2248 239. Lisse, T. S. et al. The vitamin D receptor is required DNA repair and is UV inducible in intact epidermis but
(1997). for activation of cWnt and hedgehog signaling in not in cultured keratinocytes. J. Invest. Dermatol.
214. Hu, L., Bikle, D. D. & Oda, Y. Reciprocal role of keratinocytes. Mol. Endocrinol. 28, 1698–1706 132, 2097–2100 (2012).
vitamin D receptor on β-catenin regulated keratinocyte (2014). 263. Gupta, R. et al. Photoprotection by 1,25
proliferation and differentiation. J. Steroid Biochem. 240. Teichert, A., Elalieh, H., Elias, P., Welsh, J. & Bikle, D. dihydroxyvitamin D3 is associated with an increase in
Mol. Biol. 144, 237–241 (2014). Over-expression of hedgehog signaling is associated p53 and a decrease in nitric oxide products. J. Invest.
215. Xie, Z. et al. Lack of the vitamin D receptor is with epidermal tumor formation in vitamin D receptor Dermatol. 127, 707–715 (2007).
associated with reduced epidermal differentiation and null mice. J. Invest. Dermatol. 131, 2289–2297 (2011). 264. Moll, P. R., Sander, V., Frischauf, A. M. & Richter, K.
hair follicle growth. J. Invest. Dermatol. 118, 11–16 241. Zarach, J. M., Beaudoin, G. M. 3rd, Coulombe, P. A. & Expression profiling of vitamin D treated primary
(2002). Thompson, C. C. The co-repressor hairless has a role human keratinocytes. J. Cell. Biochem. 100, 574–592
216. Watt, F. M. & Collins, C. A. Role of β-catenin in in epithelial cell differentiation in the skin. (2007).
epidermal stem cell expansion, lineage selection, Development 131, 4189–4200 (2004). 265. Ping, X. L. et al. PTCH mutations in squamous cell
and cancer. Cold Spring Harb. Symp. Quant. Biol. 73, 242. DasGupta, R., Rhee, H. & Fuchs, E. A developmental carcinoma of the skin. J. Invest. Dermatol. 116,
503–512 (2008). conundrum: a stabilized form of β-catenin lacking the 614–616 (2001).
217. Luderer, H. F., Gori, F. & Demay, M. B. Lymphoid transcriptional activation domain triggers features 266. Aszterbaum, M. et al. Identification of mutations in
enhancer-binding factor-1 (LEF1) interacts with of hair cell fate in epidermal cells and epidermal cell the human PATCHED gene in sporadic basal cell
the DNA-binding domain of the vitamin D receptor. fate in hair follicle cells. J. Cell Biol. 158, 331–344 carcinomas and in patients with the basal cell nevus
J. Biol. Chem. 286, 18444–18451 (2011). (2002). syndrome. J. Invest. Dermatol. 110, 885–888 (1998).
218. Hill, N. T. et al. 1α, 25-Dihydroxyvitamin D(3) and the 243. Daroach, M., Narang, T., Saikia, U. N., Sachdeva, N. 267. Hahn, H. et al. Mutations of the human homolog of
vitamin D receptor regulates ΔNp63α levels and & Sendhil Kumaran, M. Correlation of vitamin D and DROSOPHILA patched in the nevoid basal cell
keratinocyte proliferation. Cell Death Dis. 6, e1781 vitamin D receptor expression in patients with carcinoma syndrome. Cell 85, 841–851 (1996).
(2015). alopecia areata: a clinical paradigm. Int. J. Dermatol. 268. Regl, G. et al. The zinc-finger transcription factor GLI2
219. Stambolsky, P. et al. Modulation of the vitamin D3 57, 217–222 (2018). antagonizes contact inhibition and differentiation of
response by cancer-associated mutant p53. Cancer 244. Lee, S., Kim, B. J., Lee, C. H. & Lee, W. S. Increased human epidermal cells. Oncogene 23, 1263–1274
Cell 17, 273–285 (2010). prevalence of vitamin D deficiency in patients with (2004).
220. Salehi-Tabar, R. et al. Vitamin D receptor as a master alopecia areata: a systematic review and meta-analysis. 269. Regl, G. et al. Human GLI2 and GLI1 are part of a
regulator of the c-MYC/MXD1 network. Proc. Natl J. Eur. Acad. Dermatol. Venereol. 32, 1214–1221 positive feedback mechanism in basal cell carcinoma.
Acad. Sci. USA 109, 18827–18832 (2012). (2018). Oncogene 21, 5529–5539 (2002).
221. Reichrath, J., Saternus, R. & Vogt, T. Endocrine actions 245. Umar, M. et al. Vitamin D and the pathophysiology of 270. Grachtchouk, M. et al. Basal cell carcinomas in mice
of vitamin D in skin: relevance for photocarcinogenesis of inflammatory skin diseases. Skin Pharmacol. Physiol. overexpressing Gli2 in skin. Nat. Genet. 24, 216–217
non-melanoma skin cancer, and beyond. Mol. Cell. 31, 74–86 (2018). (2000).
Endocrinol. 453, 96–102 (2017). 246. Mady, L. J. et al. The transient role for calcium and 271. Chan, E. F., Gat, U., McNiff, J. M. & Fuchs, E. A common
222. Stenn, K. S. & Paus, R. Controls of hair follicle cycling. vitamin D during the developmental hair follicle cycle. human skin tumour is caused by activating mutations in
Physiol. Rev. 81, 449–494 (2001). J. Invest. Dermatol. 136, 1337–1345 (2016). β-catenin. Nat. Genet. 21, 410–413 (1999).
223. Morris, R. J. et al. Capturing and profiling adult hair 247. Greenlee, R. T., Hill-Harmon, M. B., Murray, T. & Thun, M. 272. Mercer, T. R., Dinger, M. E. & Mattick, J. S. Long
follicle stem cells. Nat. Biotechnol. 22, 411–417 Cancer statistics, 2001. CA Cancer J. Clin. 51, 15–36 non-coding RNAs: insights into functions. Nat. Rev.
(2004). (2001). Genet. 10, 155–159 (2009).

www.nature.com/nrendo
Reviews

273. Mattick, J. S. Long noncoding RNAs in cell and enzyme 1α-hydroxylase (CYP27B1) decreases during 294. Dixon, K. M. et al. 1α,25(OH)(2)-vitamin D and a
developmental biology. Semin. Cell Dev. Biol. 22, 327 melanoma progression. Hum. Pathol. 44, 374–387 nongenomic vitamin D analogue inhibit ultraviolet
(2011). (2013). radiation-induced skin carcinogenesis. Cancer Prev.
274. Jiang, Y. J. & Bikle, D. D. LncRNA: a new player in 1α, 284. Anderson, M. G., Nakane, M., Ruan, X., Kroeger, P. E. & Res. 4, 1485–1494 (2011).
25(OH)(2) vitamin D(3) /VDR protection against skin Wu-Wong, J. R. Expression of VDR and CYP24A1 mRNA 295. Skrajnowska, D. & Bobrowska-Korczak, B. Potential
cancer formation. Exp. Dermatol. 23, 147–150 in human tumors. Cancer Chemother. Pharmacol. 57, molecular mechanisms of the anti-cancer activity of
(2014). 234–240 (2006). vitamin D. Anticancer Res. 39, 3353–3363 (2019).
275. Underhill, C. CD44: the hyaluronan receptor. J. Cell 285. Solomon, C., White, J. H. & Kremer, R. 296. Plikus, M. V. et al. Epithelial stem cells and
Sci. 103, 293–298 (1992). Mitogen-activated protein kinase inhibits implications for wound repair. Semin. Cell Dev. Biol.
276. Screaton, G. R. et al. Genomic structure of DNA 1,25-dihydroxyvitamin D3-dependent signal 23, 946–953 (2012).
encoding the lymphocyte homing receptor CD44 transduction by phosphorylating human retinoid 297. Oda, Y. et al. Combined deletion of the vitamin D
reveals at least 12 alternatively spliced exons. Proc. X receptor α. J. Clin. Invest. 103, 1729–1735 receptor and calcium-sensing receptor delays wound
Natl Acad. Sci. USA 89, 12160–12164 (1992). (1999). re-epithelialization. Endocrinology 158, 1929–1938
277. Haggerty, J. G., Bretton, R. H. & Milstone, L. M. 286. Larriba, M. J. et al. Snail2 cooperates with Snail1 in (2017).
Identification and characterization of a cell surface the repression of vitamin D receptor in colon cancer. 298. Palmer, H. G., Martinez, D., Carmeliet, G. & Watt, F. M.
proteoglycan on keratinocytes. J. Invest. Dermatol. Carcinogenesis 30, 1459–1468 (2009). The vitamin D receptor is required for mouse hair cycle
99, 374–380 (1992). 287. Mittal, M. K., Myers, J. N., Misra, S., Bailey, C. K. progression but not for maintenance of the epidermal
278. Averbeck, M. et al. Differential regulation of & Chaudhuri, G. In vivo binding to and functional stem cell compartment. J. Invest. Dermatol. 128,
hyaluronan metabolism in the epidermal and dermal repression of the VDR gene promoter by SLUG in 2113–2117 (2008).
compartments of human skin by UVB irradiation. human breast cells. Biochem. Biophys. Res. Commun. 299. Luderer, H. F., Nazarian, R. M., Zhu, E. D. & Demay, M. B.
J. Invest. Dermatol. 127, 687–697 (2007). 372, 30–34 (2008). Ligand-dependent actions of the vitamin D receptor
279. Stern, R. Complicated hyaluronan patterns in skin: 288. Marik, R. et al. DNA methylation-related vitamin D are required for activation of TGF-β signaling during
enlightenment by UVB? J. Invest. Dermatol. 127, receptor insensitivity in breast cancer. Cancer Biol. the inflammatory response to cutaneous injury.
512–513 (2007). Ther. 10, 44–53 (2010). Endocrinology 154, 16–24 (2013).
280. Bourguignon, L. Y. et al. Selective matrix (hyaluronan) 289. Mohri, T., Nakajima, M., Takagi, S., Komagata, S.
interaction with CD44 and RhoGTPase signaling & Yokoi, T. MicroRNA regulates human vitamin D Author contributions
promotes keratinocyte functions and overcomes receptor. Int. J. Cancer 125, 1328–1333 (2009). The authors contributed equally to all aspects of the article.
age-related epidermal dysfunction. J. Dermatol. Sci. 290. Reichrath, J., Reichrath, S., Heyne, K., Vogt, T. &
72, 32–44 (2013). Roemer, K. Tumor suppression in skin and other Competing interests
281. Bourguignon, L. Y. & Bikle, D. Selective hyaluronan– tissues via cross-talk between vitamin D- and The authors declare no competing interests.
CD44 signaling promotes miRNA-21 expression and p53-signaling. Front. Physiol. 5, 166 (2014).
interacts with vitamin D function during cutaneous 291. Sequeira, V. B. et al. The role of the vitamin D Peer review information
squamous cell carcinomas progression following UV receptor and ERp57 in photoprotection by 1α,25- Nature Reviews Endocrinology thanks M. Holick, G. Jones,
irradiation. Front. Immunol. 6, 224 (2015). dihydroxyvitamin D3. Mol. Endocrinol. 26, 574–582 M. Hewison and the other, anonymous, reviewer(s) for their
282. Hsu, J. Y., Feldman, D., McNeal, J. E. & Peehl, D. M. (2012). contribution to the peer review of this work.
Reduced 1α-hydroxylase activity in human prostate 292. Chen, H. et al. Vitamin D directly regulates Mdm2
cancer cells correlates with decreased susceptibility to gene expression in osteoblasts. Biochem. Biophys. Publisher’s note
25-hydroxyvitamin D3-induced growth inhibition. Res. Commun. 430, 370–374 (2013). Springer Nature remains neutral with regard to jurisdictional
Cancer Res. 61, 2852–2856 (2001). 293. Heyne, K., Heil, T. C., Bette, B., Reichrath, J. & claims in published maps and institutional affiliations.
283. Brożyna, A. A., Jóźwicki, W., Janjetovic, Z. & Roemer, K. MDM2 binds and inhibits vitamin D
Slominski, A. T. Expression of the vitamin D-activating receptor. Cell Cycle 14, 2003–2010 (2015). © Springer Nature Limited 2020

Nature Reviews | Endocrinology

Potrebbero piacerti anche