ARTICLE

Current Understanding of the Molecular Actions of Vitamin D

Abstract

Jones, Glenville, Stephen A. Strugnell, and Hector F. DeLuca. Current Understanding of the Molecular Actions of Vitamin D. Physiol. Rev. 78: 1193–1231, 1998. — The important reactions that occur to the vitamin D molecule and the important reactions involved in the expression of the final active form of vitamin D are reviewed in a critical manner. After an overview of the metabolism of vitamin D to its active form and to its metabolic degradation products, the molecular understanding of the 1α-hydroxylation reaction and the 24-hydroxylation reaction of the vitamin D hormone is presented. Furthermore, the role of vitamin D in maintenance of serum calcium is reviewed at the physiological level and at the molecular level whenever possible. Of particular importance is the regulation of the parathyroid gland by the vitamin D hormone. A third section describes the known molecular events involved in the action of 1α,25-dihydroxyvitamin D3 on its target cells. This includes reviewing what is now known concerning the overall mechanism of transcriptional regulation by vitamin D. It describes the vitamin D receptors that have been cloned and identified and describes the coactivators and retinoid X receptors required for the function of vitamin D in its genomic actions. The presence of receptor in previously uncharted target organs of vitamin D action has led to a study of the possible function of vitamin D in these organs. A good example of a new function described for 1α,25-dihydroxyvitamin D3 is that found in the parathyroid gland. This is also true for the role of vitamin D hormone in skin, the immune system, a possible role in the pancreas, i.e., in the islet cells, and a possible role in female reproduction. This review also raises the intriguing question of whether vitamin D plays an important role in embryonic development, since vitamin D deficiency does not prohibit development, nor does vitamin D receptor knockout. The final section reviews some interesting analogs of the vitamin D hormone and their possible uses. The review ends with possible ideas with regard to future directions of vitamin D drug design.

I. METABOLIC ACTIVATION OF VITAMIN D

A. Introduction

An appreciation that vitamin D3 represents only a precursor to its functionally active form, 1α,25-dihydroxyvitamin D3 [1,25-(OH)2D3], is arguably one of the most important developments in vitamin research during the latter half of the 20th century. The discovery of the two activation steps involved in the metabolism of vitamin D3 to the hormone 1,25-(OH)2D3 (sect. i) set the stage for the elucidation of the role of vitamin D in the physiological events involved in calcium and phosphate homeostasis (sect. ii). The realization that it was the metabolites of vitamin D that were important led to an intense focus on the molecular events surrounding the mechanism of action of 1,25-(OH)2D3 , which resulted in the discovery of the vitamin D receptor (VDR) and its interaction with the transcriptional machinery inside vitamin D target cells (sect. iii). Subsequently, this led to the demonstration of new biological actions of 1,25-(OH)2D3 , in particular, its effects on the regulation of growth and differentiation of certain specialized cell types (sect. iv), which represent involvements of vitamin D not even envisioned when 1,25-(OH)2D3 was first discovered. Furthermore, the knowledge of vitamin D metabolism also provided the impetus to study the regulation of cytochrome P-450-containing enzymes involved in the process as well as stimulating the chemical synthesis of a wide range of vitamin D analogs (sect. v). This review seeks to summarize our current understanding of the molecular events surrounding the physiological action of vitamin D in these many varied areas. For a more detailed account of the subjects described in this review, particularly those of a clinical nature, the reader is directed to a recently published text (96).

Fig. 1.

Fig. 1. A: nutritional forms of vitamin D. B: steps involved in activation of vitamin D3 molecule. Note that names of cytochrome P-450 isoforms currently thought to be responsible for enzyme steps are also provided.


B. Overview of Metabolism

Vitamin D, in the form of vitamin D3 , is made from 7-dehydrocholesterol in the skin by exposure to ultraviolet light (270–300 nm range). Alternatively, vitamin D, in the form of either vitamin D2 or vitamin D3 , can be derived from dietary sources (Fig. 1 A). Both vitamin D3 and vitamin D2 undergo the same activation process, involving first 25-hydroxylation in the liver, followed by 1α-hydroxylation in the kidney, to make the biologically active compounds 1,25-(OH)2D3 and 1,25-(OH)2D2 , respectively (Fig. 1 B). There is little evidence that these two active forms differ in their mode of action, and because most is known about the synthesis and action of 1,25-(OH)2D3 , this review focuses on the natural D3 compound. The metabolic activations of vitamin D3 are carried out by specific cytochrome P-450-containing enzymes, the vitamin D3-25-hydroxylase (CYP27) and possibly another P-450 in the hepatocyte and the 25-hydroxyvitamin D-1α-hydroxylase (CYP1α) in the renal proximal tubular cell. Both of the known hydroxylases are located in the inner mitochondrial membrane of these cells (Fig. 2). The synthesis of 25-hydroxyvitamin D3 (25-OH-D3) by the liver appears to be only loosely regulated, whereas the synthesis of 1,25-(OH)2D3 by the renal 1α-hydroxylase is tightly regulated by the levels of plasma 1,25-(OH)2D3 and calcium. The renal enzyme is strongly upregulated by the hormone parathyroid hormone (PTH), a point that is discussed further in section ii. A third vitamin D-related mitochondrial cytochrome P-450-containing enzyme, the 25-hydroxyvitamin D-24-hydroxylase (CYP24), was originally believed to be exclusively located in the kidney and to be involved only in the metabolism of 25-OH-D3 to 24,25-dihydroxyvitamin D3 [24,25-(OH)2D3]. The 24-hydroxylation of 1,25-(OH)2D3 was first realized with the isolation of 1,24,25-(OH)3D3 and subsequently shown to occur in all vitamin D target tissues including enterocytes, osteoblasts, keratinocytes, and parathyroid cells. Thus it is now known that CYP24 will use 1,25-(OH)2D3 as substrate also. Because CYP24 is widely distributed around the body, is strongly induced in target cells by 1,25-(OH)2D3 , and prefers 1,25-(OH)2D3 as substrate to 25-OH-D3 , its role appears to be catabolic. The enzyme CYP24 catalyzes several steps of 1,25-(OH)2D3 degradation, collectively known as the C-24 oxidation pathway which starts with 24-hydroxylation and culminates in the formation of the biliary excretory form, calcitroic acid (Fig. 3). Thus our current view is that both the synthesis and degradation of 1,25-(OH)2D3 are tightly regulated events, attesting to the fact that the concentration of this potent hormone requires fine control at the cellular level, and hence a set of highly specific and finely tuned cytochrome P-450 exist for the purpose.

Fig. 2.

Fig. 2.Electron transport chain for mitochondrial steroid hydroxylases. Concept for 3-dimensional arrangement of components of mitochondrial cytochrome P-450-containing hydroxylases is shown.


Fig. 3.

Fig. 3.C-24 oxidation pathway.


C. Hepatic 25-Hydroxylation

Vitamin D does not circulate for long in the bloodstream but, instead, is immediately taken up by adipose tissue for storage or liver for further metabolism. In humans, tissue storage of vitamin D can last for months or even years. Ultimately, vitamin D3 undergoes its first step of activation, namely, 25-hydroxylation, in the liver (28) (Fig. 1 B). Early data suggested that the liver is the only significant site of 25-hydroxylation in vivo, although there were occasional reports of intestinal and renal extracts containing this activity (346). Research, therefore, focused on purification of the major hepatic enzyme activity. Over the years, there has been some controversy over whether 25-hydroxylation is carried out by one enzyme or two and whether this cytochrome P-450-based enzyme is found in the mitochondrial or microsomal fractions of liver. Madhok and DeLuca (207) reported that a rat liver microsomal system requiring NADPH, molecular oxygen, a flavoprotein, and a cytochrome P-450 was capable of 25-hydroxylation of vitamin D3 , but the cytochrome P-450 responsible has never been cloned. There was some speculation that the microsomal enzyme might be CYP2C11, but this cytochrome is male specific (129) and other data have also been presented that indicate that human microsomes do not possess 25-hydroxylase activity (285). Recently, Axen et al. (9) have purified a pig liver microsomal 25-hydroxylase with an NH2-terminal sequence different from that of CYP2C11 and that is capable of the 25-hydroxylation of both vitamins D2 and D3 . Currently, only the mitochondrial 25-hydroxylase has been purified to homogeneity and subsequently cloned (7 44 351). The cytochrome P-450 involved is known as CYP27 or P-450c27 because it is a bifunctional cytochrome P-450 which in addition to 25-hydroxylating vitamin D3 also carries out side-chain hydroxylation, including 27-hydroxylation of cholesterol-derived intermediates involved in bile acid biosynthesis (from which it derives its name) (252).1 The primary amino acid sequences of three species of CYP27 are depicted in Figure 4. Even though 25-hydroxylation of a variety of vitamin D compounds, including vitamin D3 , has been clearly demonstrated in cells transfected with CYP27 (119), there is still some skepticism in the vitamin D field that a single cytochrome P-450 can explain all the metabolic findings observed over the past two decades of research. The many unexplained observations suggesting that other cytochrome P-450 might perform 25-hydroxylation of vitamin D at nanomolar concentrations of substrate that exist in vivo include the following. 1) Perfused rat liver studies by Fukushima et al. (104) demonstrate kinetics consistent with two 25-hydroxylase enzyme activities: a high-affinity, low-capacity form (presumably microsomal) and a low-affinity, high-capacity form (presumably mitochondrial; CYP27).

Fig. 4.

Fig. 4.Amino acid alignments of all published vitamin D-related cytochromes P-450 from various species: CYP1α (25-OH-D3 1α-hydroxylase); CYP27 (mitochondrial vitamin D3 25-hydroxylase); CYP24 (25-OH-D3 24-hydroxylase). Note high degree of sequence similarity between all family members, particularly toward COOH terminus of each isoform. Conserved cysteine residue in block CMGRRLAELEL in extreme COOH terminus is where heme group is covalently bonded to protein. Slightly more NH2 terminal is putative ferredoxin-binding site involving LPLLKAVVKEVLRL. Another highly conserved site is putative oxygen-binding site ELLLAGVDTVSNTL.


2) Dietary studies show regulation, albeit weak, of the liver 25-hydroxylase in animals given normal intakes of vitamin D after a period of vitamin D deficiency (20), which is not explained by a transcriptional mechanism, since the gene promoter of CYP27 lacks a vitamin D-responsive element (VDRE) (118) or demonstrable responsiveness to 1,25-(OH)2D3 whereas it is regulated by bile acids (354).

3) Clinical studies show no obvious 25-OH-D3 or 1,25-(OH)2D3 deficiency occurs in patients suffering from the genetically inherited disease cerebrotendinous xanthomatosis, in which CYP27 is mutated. [Although a subset of these patients suffer from osteoporosis, this is more likely because of biliary defects leading to altered enterohepatic circulation of 25-OH-D3 (18).]

4) Substrate specificity studies using transfected recombinant human CYP27 show that the enzyme does not 25-hydroxylate vitamin D2 (119); it 24-hydroxylates vitamin D2 instead, evoking the question: Which cytochrome P-450 synthesizes 25-OH-D2?

Observations that are explained by the existence of CYP27 include 1) the occasional reports of extrahepatic 25-hydroxylation of vitamin D3 mentioned above (346), which are consistent with the detection of CYP27 mRNA in a number of extrahepatic tissues including kidney and bone (osteoblast) (10 147), and 2) the abundance of 24-hydroxylated metabolites [e.g., 24-OH-D2 , 1,24-(OH)2D2 , and 24,26-(OH)2D2] in the blood of vitamin D2-intoxicated animals (143 161 178 320). It is worth noting from perusal of Figure 4 that the recently cloned CYP1α is more closely related to CYP27 than it is to CYP24, a surprising fact given that CYP1α and CYP24 are both renal cytochromes P-450 and appear to be reciprocally regulated, thereby implying CYP27 might have evolved to metabolize vitamin D after all. Thus, although CYP27 remains the best-characterized cytochrome P-450 capable of 25-hydroxylation, it may not be the only 25-hydroxylase, and its full physiological importance remains to be established.

The product of the 25-hydroxylation step, 25-OH-D3 , is the major circulating form of vitamin D3 and in humans is present in plasma at concentrations in the range 10–40 ng/ml (25–125 nM) (140). The main reason for the stability of this metabolite is its strong affinity for the vitamin D-binding (globulin) protein of blood (DBP) (70). The metabolic fate of 25-OH-D3 is dependent on the calcium requirements of the animal. An urgent need for calcium results in renal 1α-hydroxylation, whereas an abundance of calcium results in 24-hydroxylation (see sect. ii). These two alternative pathways are discussed in turn below.

D. Renal 1α-Hydroxylation

The enzyme 25-hydroxyvitamin D3-1α-hydroxylase is responsible for the tightly regulated step that involves the introduction of a 1α-hydroxyl group into the A ring of 25-OH-D3 , thereby creating the hormone 1,25-(OH)2D3 . The specific location of this enzyme in the kidney became apparent (100) even before the unequivocal identification of 1,25-(OH)2D3 (139). Experiments involving nephrectomized animals have confirmed that the kidney as the major source of the circulating pool of 1,25-(OH)2D3 . The renal 1α-hydroxylase enzyme comprises a cytochrome P-450, a ferredoxin, and a ferredoxin reductase (112) (see Fig. 2). The cytochrome P-450 for the 1α-hydroxylase enzyme, CYP1α, was recently cloned from rat, mouse, and human (228a 298 316 332), and the amino acid sequences of these are compared with the other known vitamin D-related cytochromes P-450 in Figure 4. Because of the close resemblance with CYP27, it has been suggested that CYP1α (P-4501α) be termed CYP27B1 (298). Both 1α-hydroxylases have short mitochondrial targeting sequences but share many regions of similarity with other members of the family including the classical heme-, ferredoxin-, and oxygen-binding sites indicated in Figure 4. The 1α-hydroxylase is induced by PTH through a cAMP/phosphatidylinositol 4,5-bisphosphate (PIP2)-mediated signal transduction mechanism that is still to be defined at the molecular level (132). The enzyme appears to be downregulated by vitamin D status, possibly through a VDR-mediated transcriptional mechanism involving the hormonal product 1,25-(OH)2D3 , although there were early claims that 1,25-(OH)2D3 might act directly on its own synthesis through an allosteric mechanism (117). Work using the perfused vitamin D-deficient rat kidney (281) elegantly shows that the downregulation of 1α-hydroxylation takes 2–4 h after exposure to 1,25-(OH)2D3 and is blocked by inhibitors of protein synthesis and transcription. In the same model, the disappearance of the 1α-hydroxylase is mirrored by the reciprocal appearance of the renal 25-OH-D3-24-hydroxylase, in effect a complete “switchover” from 1α- to 24-hydroxylating activity in the isolated organ over the 4-h period. The exact reciprocal regulation of the two enzymes, first demonstrated in vivo by Tanaka et al. (336) two decades ago, led some workers to postulate that the 1α- and 24-hydroxylases might share a single cytochrome P-450 polypeptide chain, its catalytic properties modified by NH2-terminal truncation (111) or regulated by the phosphorylation state of the ferredoxin component of the enzyme (300). The cloning of two distinct cytochromes P-450, representing 1α-hydroxylation and 24-hydroxylation, suggests the first hypothesis to be incorrect, the switchover process probably being accomplished by de novo protein synthesis of the required cytochrome P-450. The second hypothesis involving regulation of enzyme activity through ferredoxin phosphorylation remains a possibility for fine-tuning of enzyme activity, although the details of this remain obscure at this point. The cloning of CYP1α has already led to mapping of the human gene to chromosome 12q13.1-q13.3 (316), the same locus as the gene defect for vitamin D dependency rickets type I (192), a disease cured by small doses of exogenous 1,25-(OH)2D3 and which had been previously postulated to be because of a mutated version of the 1α-hydroxylase (99). The next decade will see a clarification of the molecular and clinical aspects surrounding this key regulatory step of calcium homeostasis.

There have been indications that there are extrarenal versions of the 1α-hydroxylase existing in cells of monocyte/macrophage, placental, and keratinocyte lineages (4 22 83). There is strong evidence that the extrarenal enzyme located in macrophages plays a major role in certain granulomatous conditions (e.g., sarcoidosis), causing uncontrolled elevations of blood 1,25-(OH)2D3 levels, which subsequently result in troublesome hypercalcemia and hypercalciuria (3). The normal function of this and other extrarenal 1α-hydroxylases remains obscure at this time, although some have postulated a paracrine or autocrine role for locally produced 1,25-(OH)2D3 . Once again, it is safe to predict that the molecular basis for various reports of extrarenal activity in cultured cells in vitro (4 22 83) and in certain human disease states will be resolved shortly with the cloning of the renal enzyme (298 316 332), as will its mechanisms of regulation.

E. 24-Hydroxylation

The discovery of 24,25-(OH)2D3 (138 327) predated even the identification of 1,25-(OH)2D3 and the recognition of 24-hydroxylation as a metabolic step allowed for the search for the enzyme activity. The relative ease with which 24,25-(OH)2D3 was generated in such large amounts was a clue that the metabolic step was upregulated rather than downregulated by vitamin D administration. The earliest report of the 25-OH-D3-24-hydroxylase was a subcellular localization study (174) using vitamin D-replete chicken kidney tissue which established that the 24-hydroxylase is a mitochondrial cytochrome P-450-containing enzyme, and this was followed by a reconstitution experiment involving partially purified enzyme components (248). Evidence was also emerging both in vivo and in vitro that the 24-hydroxylase was not confined to the kidney but could be found in classical vitamin D target tissues including the intestine and bone (185 347). In the late 1970s, it became evident that 24-hydroxylation was probably only the first step in an inactivation process. First, it was shown to be induced by 1,25-(OH)2D3 itself (172 187 335 337 340), and the product 1,24,25-(OH)3D3 was 10 times less biologically active than 1,25-(OH)2D3 (56 335). Second, 24,25-(OH)2D3 and its 1α-hydroxylated analog, 1,24,25-(OH)3D3 , could be converted to further metabolic products containing a 24-oxo and/or 23-hydroxy groups as well (237 246 330). The perfused rat kidney was helpful in establishing the production of these catabolites (160) but revealed two further pieces to the puzzle of the metabolic role of 24-hydroxylation. First, the perfused rat kidney allowed for a clarification of the temporal relationship of these catabolites, in effect, suggesting the existence of a pathway from 1,25-(OH)2D3 and/or 25-OH-D3 (160), and second, the perfused kidney generated two side chain-cleaved molecules: a 23-alcohol (159) and a 23-acid (208 276), not observed previously in vitro. The 1α-hydroxylated 23-acid, calcitroic acid, is observed in vivo and has been shown to be the principal biliary excretory form of 1,25-(OH)2D3 (94). These discoveries led to the rationalization of many findings from a number of laboratories into a hypothesis that 24-hydroxylation is the first step of a target cell C-24 oxidation pathway (Fig. 3) whose major function is to convert 1,25-(OH)2D3 to calcitroic acid (208 276). The demonstration of vitamin D-inducible, calcitroic acid production in bone (UMR106) and kidney (208) was followed by demonstration of vitamin D-inducible C-24 oxidation pathway activity in a number of vitamin D target cells including intestine (Caco-2), keratinocyte (HPKIA-ras), and breast (T47D and MCF-7) (215 278 344).

In the early 1990s, Okuda's group succeeded in cloning the cytochrome P-450, CYP24 or P-450cc24, representing the 24-hydroxylase (62 148 247). The amino acid sequences of three species of CYP24 are shown in Figure 4. It belongs to the same subfamily as the other two known vitamin D-related cytochromes. The structure of the gene for CYP24 has been described for several species (rat, mouse, and human), and in each case shown to possess two VDRE in its proximal promoter (249 367 368). These VDRE allow for the 1,25-(OH)2D3 VDR-mediated upregulation of CYP24 from undetectable to strongly detectable expression at the mRNA level within 4 h (297). This is consistent with the enzyme activity pattern first observed in the kidney but subsequently reported for a variety of vitamin D target cells after initial exposure to hormone (127 202 209). Metabolic studies with the recombinant CYP24 protein produced in bacteria or insect cells have been equally enlightening. Akiyoshi-Shibata et al. (5) and Beckman et al. (15) have shown that CYP24 is a multicatalytic enzyme capable of several if not all of the steps illustrated in Figure 3. It is likely that CYP24 is able to catalyze three successive oxidations, two at C-24 and one at C-23, to give an intermediate that is subsequently cleaved by an unknown mechanism. All prevailing evidence suggests that C-24 oxidation is a highly efficient process giving rise to molecules of lower biological activity (e.g., calcitroic acid) such that if C-24 hydroxylation is blocked by a general cytochrome P-450 inhibitor such as ketoconazole, 1,25-(OH)2D3 hormone action is extended (265). This argues that the main role of the C-24 oxidation pathway is attenuation of the biological signal inside target cells.

Recently, this hypothesis was tested when St. Arnaud et al. (315) engineered a CYP24-null mouse. At this time, the results have been reported only in abstract form, but it is clear that the defect is not lethal during embryonic development. Instead, at least one-half the mice exhibit hypercalcemia/hypercalciuria in early neonatal life and quickly die before weaning from nephrocalcinosis. The other half of animals survive and appear healthy perhaps due to the upregulation of some alternative vitamin D-catabolic pathway (C-26 hydroxylation or 26,23-lactone formation). All CYP24-null animals exhibit abnormal bone histology, characterized by excessive unmineralized bone matrix and reminiscent of that previously observed in experiments involving exogenous 1,25-(OH)2D3 intoxication (136). This could be caused by an inability of target cells, in this case osteoblasts, to turn off the 1,25-(OH)2D3 signal in the absence of the C-24 oxidation pathway. An alternative explanation that CYP24 is needed for the synthesis of some essential 24-hydroxylated metabolite of vitamin D [e.g., 24,25-(OH)2D3] needed for a yet to be defined role in bone formation has been proposed (314) but seems unlikely for several reasons. Among the data that argue against such an essential role for 24-hydroxylated metabolites are the findings from the study of fluorinated analogs of vitamin D (149 234 338). These analogs are irreversibly blocked in the C-24 and C-23 positions of the side chain with one or more atoms of fluorine and therefore unable to undergo 24- or 23-hydroxylation. However, they are fully biologically active in all known in vivo functions of vitamin D. Furthermore, the biochemical machinery required for transducing the signal from a 24-hydroxylated metabolite has never been satisfactorily demonstrated in bone (or any other) cells. Thus it appears that 24-hydroxylation is not essential for vitamin D to fulfill its many biological roles in vertebrate biology. Therefore, evidence from the CYP24-null mouse seems to confirm our hypothesis that the C-24 oxidation pathway is a complex, self-induced mechanism for limiting the action of 1,25-(OH)2D3 in vitamin D target cells once the initial wave of gene expression has been initiated (see sect. iii). This is discussed in greater detail in section v.

II. ROLE OF VITAMIN D IN CALCIUM HOMEOSTASIS

A. Development of Calcium Homeostatic Mechanisms

Calcium is undoubtedly one of the most tightly regulated substances in plasma of higher animals (81 270). Its concentration is held constantly at 1 mmol ionized calcium or 10 mg/100 ml of total calcium. The ionized calcium concentration of plasma is very close to that found in seawater (270), and it is believed that the evolution of the calcium homeostatic system took place as animals emerged from the sea into fresh water and further onto land. Very likely the dependence of a number of life's essential functions on calcium occurred because of the constancy and abundance of calcium in seawater. Among them are the neural transmission, muscle contraction (and relaxation), exocrine secretion, blood clotting, and the adhesion of cells to each other. The presence of calcium in abundance in seawater also made understandable the use of calcium in the construction of structural elements such as the skeleton. As animals emerged into fresh water, resulting in a drop in ambient calcium concentration, it immediately brought into need the ability to mobilize calcium to meet the needs of the very critical functions such as neural transmission and muscle contraction. This only intensified as animals emerged from fresh water onto land where calcium availability was even more limited than in fresh water. Furthermore, the gravitational forces applied to the terrestrial animals must have increased the need for a structurally sound skeleton. Thus the evolved mammal had many problems to solve before it could live as we now know life. It must be able to aggressively acquire environmental calcium when required (81 80 130 270 283). It had to have a constant source of calcium available to plasma to support nerve and muscle functions. It also had to be able to construct a skeleton of considerable strength to protect the organism and to provide for motility. Finally, the reproductive needs of the animals had to be satisfied, including provision of calcium during embryonic and postembryonic development including construction of an entirely new skeleton of the unborn animals. It is, therefore, clear that the calcium homeostatic system is a very complex one (see Fig. 5) that involves many hormones, most of which are known but some that are not. From this discussion, it will become apparent that the vitamin D endocrine system is the basic one in managing calcium of plasma, with equally important roles for the parathyroid hormone and calcitonin.

Fig. 5.

Fig. 5.Diagrammatic representation of calcium homeostatic system. PTH, parathyroid hormone; 1,25(OH)2D3 , 1α,25-dihydroxyvitamin D3 ; PTG, parathyroid gland; C cells, parafollicular cells of the thyroid that secrete calcitonin (CT).


B. Role of Parathyroid Gland and Its Hormone

For many years it has been clearly recognized that the parathyroid gland is the calcium-sensing organ in the body (266 270 283). Thus, in response to even slight hypocalcemia, the parathyroid glands react within seconds to secrete the 84-amino acid peptide hormone PTH (302). This hormone then initiates the sequence of events that results in the mobilization of calcium to replace that which has been taken from plasma.

There has been a great deal of recent effort expended in the direction of the calcium receptor for the parathyroid glands. The calcium receptor is well known (39 225) and appears to act in a cAMP-dependent mechanism to facilitate the secretion of PTH. This calcium receptor may play an important role in other tissues as well. Interested readers are directed elsewhere (301). The PTH has a lifetime in plasma that can be measured in minutes if not seconds (238). The receptor for the PTH is known and has been cloned (289). This receptor is found throughout the length of the nephron of kidney, is not found in the intestine, and is found in the osteoblasts but not osteoclasts of the skeleton (2). In the kidney, the PTH plays an important role in many functions (101). Well known is that it blocks reabsorption of phosphate causing a phosphate diuresis (39). In the proximal convoluted tubule cells, it activates the 25-hydroxyvitamin D-1α-hydroxylase (25-OH-D3-1α-OHase) that converts 25-hydroxyvitamin D3 to the active hormone, 1,25-(OH)2D3 (109 298 340). As already discussed in section i, the 25-OH-D-1α-OHase has now been cloned by three research groups (298 316 332) that will undoubtedly result in our understanding of the molecular mechanism whereby PTH activates the 1α-OHase. At the present time, it is known that PTH through cAMP (141) activates the 1α-OHase by increasing the mRNA encoding for this important enzyme (298). Whether it acts at the transcriptional level or elsewhere remains to be determined. At the same time the PTH through cAMP activates the 1α-OHase, it markedly suppresses the 25-OH-D-24-OHase, the major enzyme involved in destruction of the vitamin D hormone as described in section i (297). Again, the mechanism of suppression of 24-hydroxylase (24-OHase) by PTH remains unknown, although there is clearly a decrease in the mRNA encoding for the 24-OHase. These two actions result in a marked elevation of plasma levels of 1,25-(OH)2D3 (341).

C. Physiological Actions of 1α,25-Dihydroxyvitamin D3

The consequences of an elevation of this major calcium mobilizing hormone are as follows: 1,25-(OH)2D3 acts by itself to initiate active intestinal calcium transport in the small intestine (35). This system has a relatively long lifetime, being measured in days (123), whereas the other actions of 1,25-(OH)2D3 are much shorter. 1,25-(OH)2D3 also activates osteoblasts. The result of this activation is to either stimulate the osteoclast to resorb bone and/or to activate the reverse transport of calcium from the bone fluid compartment to the plasma compartment (137 328 329 339). The final result is that calcium is mobilized by the skeleton into the plasma compartment by the action of the vitamin D hormone and the PTH. Of considerable importance is that vitamin D-deficient animals having abundant calcium in their bones will not mobilize calcium from the skeleton in response to PTH unless vitamin D is provided (110 271). Similarly, parathyroidectomized animals cannot mobilize calcium in response to 1,25-(OH)2D3 unless PTH is provided (110). Therefore, the presence of both the PTH and the vitamin D hormone are required for this system to operate in vivo. Whether this mechanism is through osteoclastic-mediated bone resorption or a membrane transport phenomenon remains to be determined. In the distal renal tubule, another mechanism proceeds that is dependent on both the PTH and the vitamin D hormone (186 362). Again, these two hormones acting in concert cause the reabsorption of the last 1% of the filtered load of calcium into the plasma compartment. These sources of calcium then cause a rise in serum calcium that then clears the sensing point of the calcium receptor. This then shuts down the secretion of the PTH. It does not appear that the PTH-related protein (PTHrP) functions in this system but may play a role in abnormal calcium mobilization as, for example, in malignancy (212).

D. Role of Calcitonin

The danger to hypercalcemia is calcification of soft tissues especially kidney, heart, aorta, and intestine, causing organ failure and death. To guard against hypercalcemia, not only is the shut-off of the parathyroid gland important but also the turning on of the C cells or the parafollicular cells of the thyroid to secrete the hormone calcitonin. This is a 34-amino acid peptide hormone that is responsible for lowering serum calcium by its action on the skeleton (57). It directly acts on osteoclasts and osteocytes reducing the calcium mobilizing activity and shutting down calcium coming from the skeleton (57). Although other actions of calcitonin have been described in kidney and intestine, by far the most important in the regulation of serum calcium is that which occurs at the skeleton. There have been reports of calcitonin regulating vitamin D metabolism (16 105); however, these largely are secondary to changes in parathyroid secretion, and there is no convincing evidence that calcitonin plays any direct role on regulation of the vitamin D hormonal levels.

E. Vitamin D Metabolites and Other Hormones

There has been considerable interest in other metabolites of vitamin D playing an important role in the regulation of calcium mobilization in suppression of hypercalcemia, or in bone growth. Of particular importance are the many studies carried out on 24,25-(OH)2D3 . This compound has been alleged to be important in the regulation of calcium homeostasis (245) or in the formation of the skeleton (33 230 255) or in counteracting the hypercalcemia activity of 1,25-(OH)2D3 (201). By now, extensive studies have been carried out with fluoro analogs, namely, 24,24-F2-25-OH-D3 to show that hydroxylation on the 24-position has no functional significance (38 149). Thus animals grown for two generations with 24,24-F2-25-OH-D3 as their sole source of vitamin D illustrate that 24-hydroxylation is not required either for skeletal formation, maintenance of calcemia, or growth of bone. More recently, a knockout of the 24-hydroxylase has been reported, and the results are not at all clear as to whether 24-hydroxylation plays a role except in the destruction of the potent hormone 1,25-(OH)2D3 and its precursor, 25-OH-D3 (313). Further work using these models is required before any conclusions can be reached.

Other hormones such as estrogen and glucocorticoids have significant effects on bone and calcium metabolism, but they do not appear to be directly involved in regulating serum calcium concentration. This seems largely to be the role of the vitamin D hormone, the PTH, and calcitonin.

F. Intestinal Calcium and Phosphate Absorption

From a historical point of view, the role of 1,25-(OH)2D3 in intestinal absorption of calcium is perhaps best known. Orr et al. (256) discovered that vitamin D is required for intestinal calcium absorption many decades ago. This was reaffirmed by the work of Nicolaysen and Eeg-Larsen (239), who further demonstrated that the need for calcium increased the ability of the animal to absorb calcium. He postulated the existence of an endogenous factor that would inform the intestine of the skeletal needs for calcium. This basic observation was then shown to be primarily the vitamin D endocrine system, and 1,25-(OH)2D3 has been thought to be the agent that stimulates intestinal calcium absorption to meet the needs of the skeleton (34). It is, therefore, abundantly clear that intestinal calcium absorption remains as one of the basic functions of 1,25-(OH)2D3 . Quite independently of calcium is the role of 1,25-(OH)2D3 in stimulating intestinal absorption of phosphate (126 180). Both are active calcium transport mechanisms, but they appear to be independent of each other (63 126 180). The molecular mechanism of action of 1,25-(OH)2D3 in stimulating intestinal calcium absorption and intestinal phosphate absorption remains unknown, despite many efforts by many investigators. 1α,25-Dihydroxyvitamin D3 stimulates the production of calbindin D9k in mammals (342) and calbindin D28k in birds (68) to appear in the intestine. In the case of the 28k protein in birds, it is absent in deficiency and present in large amounts after stimulation by vitamin D (357). A vitamin D-responsive element has been demonstrated to be present in the calbindin D9k promoter in mammals (77) and the 28k mammalian gene (113) but that has not been shown for the calbindin D28k protein of birds. The exact molecular mechanism for initiating production of the calbindin D28k remains to be determined. Furthermore, if one studies the time course of appearance of the 28k in birds as related to calcium absorption, there is no clear-cut correlation (125 312). The appearance of this protein and calcium transport coincide as a function of time in response to vitamin D or 1,25-(OH)2D3 . However, calcium absorption diminishes while calbindin D28k remains high in the gut. Therefore, there appears to be at least some discrepancy between calbindin D28k and calcium transport. It has suggested that some other protein or proteins are involved. This led to an analysis of a calcium pump in the basolateral membrane that is induced by 1,25-(OH)2D3 (357). However, the degree of induction in the time course of its appearance is not certain to account for the role of vitamin D in calcium transport. In short, the molecular mechanism of action of 1,25-(OH)2D3 in inducing intestinal calcium and phosphate transport is largely unknown. Wasserman and Feher (357) believe there are multiple sites of action of 1,25-(OH)2D3 in intestinal calcium absorption. Considerable work will be required before one can demonstrate the exact role of the calbindin proteins and the calcium pump in the vitamin D-induced calcium transport.

G. Vitamin D and Bone Calcium Mobilization

Even more poorly understood is the role of vitamin D in bone resorption or bone mobilization. The idea that vitamin D could result in the mobilization of calcium from bone was derived from the early work of Bauer et al. (14). A vitamin D-deficient animal on a zero-calcium diet will provide an increase in serum calcium at the expense of skeleton when given vitamin D. This mechanism requires the presence of PTH (110). Furthermore, 1,25-(OH)2D3 is clearly a stimulator of osteoclastic bone resorption in culture (269 319). However, the osteoclast has neither a receptor to the PTH nor a receptor to the vitamin D hormone (223). Instead, a signal appears to arise from interaction of these two hormones with the osteoblast. This signal causes the osteoclast to resorb bone (221 222 328). There is also the possibility that upon 1,25-(OH)2D3 and/or PTH signaling, the osteoblast may cause the transport of calcium from the bone fluid compartment to the plasma compartment (333). The nature of the signal arising from stimulation by the PTH and by the vitamin D hormone on the osteoblast has not been determined.

On the other hand, a great deal of progress has been made in understanding the role of vitamin D and osteoclastic formation and differentiation (328). Through the work of Abe et al. (1) and Tanaka et al. (334) has come the discovery that 1,25-(OH)2D3 plays an important role in causing the differentiation of promyelocyte to the monocyte. The monocyte is considered the precursor of the giant osteoclast. Further differentiation of the osteoclast precursor is catalyzed by the osteoclast differentiation factor that is produced by the osteoblast in response to 1,25-(OH)2D3 (328).

The action of vitamin D in stimulating osteoclastic bone resorption may be to provide bone calcium for the plasma but more likely to cause bone resorption in preparation for coupled formation to complete the bone-remodeling process. This would argue that the vitamin D hormone is involved in this important function that strengthens bone and repairs microfractures that may have occurred during bone usage. At this stage, therefore, 1,25-(OH)2D3 may be viewed as an important hormone that not only plays a role in bone calcium mobilization when required but also plays an important role in initiating bone remodeling and modeling systems required for shaping bone and required for repairing bone (107 319).

H. 1α,25-Dihydroxyvitamin D3 Regulates the Parathyroid Gland

A new addition to the calcium homeostatic system was the discovery that the parathyroid gland is a target of vitamin D action. The initial work carried out by Stumpf et al. (324) showed that the parathyroid gland is a site of localization of highly labeled 1,25-(OH)2D3 . This was followed by two groups who demonstrated the presence of VDR in that tissue (133 145). Finally, a 1,25-(OH)2D3 suppression of parathyroid gland proliferation and of parathyroid production became known. A suppression of parathyroid secretion could be demonstrated in dialysis patients treated with intravenous 1,25-(OH)2D3 . The parathyroid gene was cloned (303), and the VDRE was demonstrated in the promoter region of the gene (85 184). This VDRE acts clearly in a mechanism to suppress transcription of the parathyroid gene. Thus a new loop in the calcium homeostatic system was discovered. Figure 5 shows all of these mechanisms that work together in the regulation of serum calcium concentration.

III. MOLECULAR MECHANISM OF ACTION AT TARGET CELLS

A. Overall Mechanism of Transcriptional Regulation by Vitamin D

The mechanism by which 1,25-(OH)2D3 exerts its effects on transcription is rapidly being uncovered. It is becoming clear that the vitamin D system shares many features with other ligand-activated nuclear receptors such as the retinoic acid receptor (RAR) and thyroid hormone receptor (TR) (95), including many of the same transcriptional cofactors. A summary of what is currently known about the activation process is shown in Figure 6.

Fig. 6.

Fig. 6.Proposed mechanism for transcriptional upregulation of a target gene by vitamin D receptor (VDR). After ligand binding, receptor forms a heterodimer on a response element with retinoid X receptor (RXR). Binding of coactivator protein to heterodimer-DNA complex is followed by histone acetylation and subsequent release of histones from DNA. Transcription factors are then able to initiate transcription of target gene, resulting in production of corresponding protein. RNAP, RNA polymerase; CBP, calcium-binding protein; DRE, vitamin D response element.


In the first place, activation of vitamin D target genes has been shown to require a specific nuclear receptor protein, the VDR. Target genes are upregulated through binding of the VDR protein to specific DNA sequences termed response elements in the promoter regions of these genes. As shown at the top of Figure 6, binding of 1,25-(OH)2D3 to the receptor increases heterodimerization of VDR with a cofactor, the retinoid X receptor (RXR), on a response element (169). Binding of the heterodimer to the response element induces a bend in the DNA of the promoter (151). Binding of 1,25-(OH)2D3 to the receptor also appears to change the conformation at the COOH terminus of the VDR, permitting a region termed the AF-2 domain to interact with other transcription factors, including coactivator proteins such as SRC-1 (216 253 279 356). Recent exciting work indicates that coactivator proteins possess intrinsic histone acetylase activity (59 345). These coactivators bind to transcriptional “integrators” such as calcium-binding protein (CBP) and p300 (59 152 165), which, in addition to other functions, have also been shown to possess histone acetylase activity (179). Thus recruitment of coactivators to a promoter by a liganded receptor appears to result in remodeling of DNA structure through acetylation of histones and their subsequent release from DNA. This in turn leads to opening of the promoter to the transcriptional machinery. The net result of binding of liganded receptor to an upregulated target promoter is therefore to increase the rate of transcription of the gene, leading to increased production of the corresponding protein. It must be stressed at this point that a number of the details of this proposed mechanism are unclear at this point, including the exact sequence of events after receptor binding to the promoter, so the order of events presented in Figure 6 is somewhat speculative.

In contrast to genes that are upregulated by vitamin D, several genes, including PTH (85 251) and interleukin (IL)-2 (6), have been shown to be downregulated by 1,25-(OH)2D3 . The means by which this downregulation is carried out is not clear in all cases, and at least two possibilities exist. The first is that, as has been proposed for the IL-2 promoter, VDR may bind to a downregulatory response element and disrupt the binding of upregulatory transcription factors, leading to a decrease in transcription (6). For other downregulated genes, the situation may be quite different, and binding of VDR to an inhibitory response element may lead to interactions with repressor proteins that decrease transcription of the gene. Interestingly, corepressor proteins such as nuclear receptor corepressor (NCoR) (142) and silencing mediator for retinoic acid receptor and thyroid hormone receptor (SMRT) (60) have recently been found to bind, through intermediary proteins, to histone deacetylase enzymes (131 233). Presumably in this situation the deacetylated histones then bind to the promoter of the downregulated gene and shut off transcription. The VDR has not yet been reported to bind to corepressor proteins, but given the rate at which such proteins are currently being discovered, it is possible that a corepressor that binds to VDR will be uncovered.

Phosphorylation of the receptor may also play a role in the induction of transcription by the VDR (73), perhaps through modulation of the affinity of VDR for the various cofactor proteins involved in transcription. Rapid phosphorylation of the VDR has been shown to occur in organ culture systems upon addition of ligand (41). The phosphorylated residues have been localized to the ligand-binding domain of the protein. The exact functional consequences of this phosphorylation have been difficult to determine. Estrogen receptor phosphorylation by mitogen-activated protein kinase has been shown to have a direct and measurable effect on transcription (167), but this phosphorylation was mapped to the NH2-terminal AF-1 region of the estrogen receptor, which is lacking in VDR. Effects of phosphorylation on activation of the AF-2 domain, at the COOH terminus of the protein, have not yet been clearly shown for the VDR. Indeed, even the kinase (or kinases) responsible for the phosphorylation in vivo has yet to be determined. Some studies have suggested that serine-208 of the VDR can be phosphorylated by casein kinase II and that phosphorylation of this residue may cause increased transcriptional activity (162). Other work has revealed that mutation of this serine failed to affect transcription, although alternate phosphorylation on adjacent serines was noted (135). The exact effects of phosphorylation must await determination of the precise amino acids phosphorylated in vivo and the functional consequences that follow from this.

B. Vitamin D Receptors

The VDR is a member of a superfamily of nuclear receptors (95). Within this family, the VDR has the highest similarity to the subfamily that includes retinoic acid, thyroid hormone, and peroxisome proliferator activator receptor (PPAR) receptors, to which it has sequence and structural resemblance. The VDR has been cloned from several species (12 42 92 164 219) and shows considerable similarity between species in size and sequence. In the rat, for example, the VDR protein consists of 423 amino acids, with a molecular mass of ∼50 kDa, whereas in the human the protein has an additional 4 amino acids at the NH2 terminus, for a total of 427.

Like the other nuclear receptors, the VDR can be divided by function into several domains. An illustration of the different domains of the VDR is shown in Figure 7. At the NH2 terminus is a truncated A/B domain of ∼20 amino acids, to which little function has yet been ascribed for the VDR. After this, the DNA-binding domain, termed the C domain, is located between amino acids 20 and 90. A D or hinge domain is located approximately between amino acids 90 and 130, followed by the COOH-terminal E or ligand-binding domain between amino acids 130 and 423. The ligand-binding domain of the protein is a complex region of the protein, responsible for high-affinity binding of ligand, for dimerization with RXR, and for binding to transcription factors (95). It should be noted that exact delineation of the division between the hinge region and the ligand-binding domain is somewhat uncertain and is based on deletion analysis.

Fig. 7.

Fig. 7.Schematic illustration of structure of VDR protein, showing functional domains of which protein is composed. Regions of receptor thought to interact with transcription factors TFIIB and RXR are shown. COOH-terminal AF-2 domain is shown in black.


Knowledge of the structural and functional properties of the receptor proteins has increased dramatically in recent years, after the advent of systems which allow expression and purification of large quantities of receptor protein from bacteria. This has allowed a piecemeal approach to determination of receptor domain structures; the DNA-binding and ligand-binding domains of several receptors have been expressed separately and purified, and structures have been determined by NMR or X-ray crystallography (32 272). In contrast, to this point, information concerning the VDR has come primarily from site-directed mutagenesis, in which the receptor cDNA is mutated and the mutant protein studied by transfection into mammalian cells (236). No structural data are yet available concerning the VDR, but the other receptor domains may serve as models on which to base hypotheses about the VDR and its properties.

Structures for the DNA-binding domains of other receptors, including the RXR, have been obtained using both NMR and X-ray crystallography. The RXR structure showed that the DNA-binding domain, comprised of two zinc finger motifs, consists of two α-helixes oriented at approximately right angles to one another (194). One helix, termed the orientation helix, thought to be critical for recognition of the receptor response element was proposed to fit into the major groove of the DNA and bind to a specific DNA sequence in the response element. These domains contain two zinc atoms tetrahedrally coordinated to eight conserved cysteine residues. Both the zinc atoms and the cysteine residues are necessary to maintain the three-dimensional structure required for response element recognition and DNA binding. The amino acid sequence of the DNA-binding domain is similar between members of the receptor superfamily, suggesting that these structures can be used as a model for that of the VDR.

There is as yet no direct evidence for the structure of the VDR-RXR complex bound to DNA. However, the crystal structure of the TR and RXR DNA binding domains complexed to a DR4 response element has been determined (272). With the use of this structure as a model, the VDR DNA-binding domain was hypothesized to have specific amino acid contacts with RXR: between the asparagine residue 14 of VDR and RXR residues glutamine-49 and arginine-52. In addition, lysine-68 and glutamate-69 of VDR were modeled to form salt bridges with RXR residues aspartate-39 and arginine-38, respectively. These interactions were thought to account for the optimal spacing of three residues between the direct repeats of the vitamin D response element (DR3).

The crystal structures of the ligand-binding domains of the RXR-α (without ligand) (32), the RAR-γ (with ligand) (279), and the TR-α1 (with ligand) (356) have been determined. The ligand-binding domains of the RXR and RAR had been previously predicted to have a high content of α-helix (64 204), and this was confirmed by structural analysis. All three ligand-binding domains were found to share a common secondary structure of 12 α-helixes, with a small content of β-sheet. A comparison of the structures of the RAR and TR with that of the VDR is shown in Figure 8. The COOH-terminal portion of the proteins, termed the AF-2 domain, has been determined to be critical for transcription. Removal of this portion of the protein results in decreased ligand-binding affinity and loss of transcriptional activation. The structural data indicate that the helixes 11 and 12 of the RAR and TR may undergo a large conformational change in response to ligand binding, folding up around the ligand in what the authors of the RAR paper term a mouse trap action to form a hydrophobic pocket (279). This brings the amino acid residues on helix 12 which make up the AF-2 domain into position to interact with other transcription factors. The VDR shows sequence similarity to the RAR and TR in this region and may be expected to possess a similar three-dimensional structure and to undergo a similar conformational change upon ligand binding.

Fig. 8.

Fig. 8.Sequence alignment of human retinoic acid receptor (RAR)-γ, rat thyroid hormone recetpor (TR)-α and rat VDR. Sequences of ligand-binding domains of these proteins were aligned using program “Pileup” (Genetics Computer Group, University of Wisconsin-Madison). Gaps introduced into sequence to optimize alignment are denoted by dots. Numbering of amino acids of each ligand-binding domain (LBD) is to right of sequences and corresponds to numbering of full-length protein. α-Helical regions of RAR and TR are underlined with solid lines; β-sheet regions are underlined with dashed lines. Regions of amino acid identity are enclosed with boxes. Helixes are denoted by numbers in accordance with TR structure; sheet regions are denoted s1-s4, also according to TR structure. Amino acids in RAR and TR shown by X-ray crystallography to be in contact with ligand are shaded.


The wild-type VDR binds its ligand, 1,25-(OH)2D3 , with extremely high affinity, in the range of 10−10 M (208 282). Both the 1α- and 25-hydroxyl groups are critical for high-affinity binding, and the absence of either results in approximately a 500-fold decrease in affinity of ligand for the receptor (82). The exact amino acid residues in the receptor that contact the ligand remain unknown, although some work has been done recently on affinity labeling of the binding site (273). Previous work with transfected cells has shown that deletion of the NH2-terminal 116 amino acids of the VDR left a protein with measurable ligand binding activity, whereas deletion of the NH2-terminal 160 amino acids did not (220). More recent work with fusion proteins has shown that deletion of the NH2-terminal 124 amino acids gave rise to a functional ligand-binding protein, whereas deletion to amino acid 172 did not (242). From the other end, removal of the COOH-terminal 20 amino acids of the VDR resulted in a 10-fold loss of affinity for ligand, whereas removal of more than this number resulted in complete loss of ligand binding (236). Thus a core domain of ∼300 amino acids was shown to be required for the protein to bind ligand with wild-type affinity. This is in accordance with the ligand-binding domains of the related receptors that have been crystallized, which were all expressed in bacteria as proteins of between 250 and 300 amino acids.

An alignment of the VDR ligand-binding domain sequence with that of the RAR and TR supports the possibility that amino acids in the VDR directly in contact with ligand begin at approximately amino acid 220. In Figure 8, amino acids in contact with ligand in the RAR and TR crystal structures are shaded. Inspection of this figure shows that the contact amino acids begin at helix 3 in both the RAR and TR. Based on sequence alignment, the same region of the VDR begins at approximately amino acid 220, with a leucine residue that is conserved in all three proteins. Interestingly, work in which 10 amino acid segments of the VDR were sequentially deleted found that impairment of ligand binding did not occur until approximately amino acid 230, which is in accordance with this possibility (154).

Site-directed mutagenesis of the VDR ligand-binding domain has been performed recently on several of the cysteine residues in the human protein (235). Alteration of cysteine-288 to glycine resulted in severe attenuation of ligand binding at room temperature, whereas the same mutation at cysteine-337 resulted in a smaller decrease in affinity. The exact significance of this result is uncertain, although a contact between cysteine-237 and carbon-13 of retinoic acid was noted in the binding of retinoic acid to the RAR-γ (279). Interestingly, in Figure 8, the corresponding cysteine residue (cysteine-284) of the rat receptor aligns exactly with ligand-contacting amino acids in the RAR and TR, suggesting that this cysteine may in fact directly contact the ligand.

The VDR is generally expressed at relatively low levels in vivo. Target tissues, such as bone, kidney, and especially intestine, may have relatively high levels of receptor (3,000–6,000 fmol/mg protein), but in other tissues, the levels are generally much lower (72). The receptor has been shown to be present in most tissues that have been examined, including activated immune cells such as T cells, where it may play a role in modulating the levels of cytokines such as IL-2 (6). In contrast to other receptors, such as the glucocorticoid receptor, which is associated with a number of proteins, including heat shock proteins, in the cytoplasm before hormone binding (146), studies with radiolabeled 1,25-(OH)2D3 have shown that the VDR is predominantly nuclear (325). Little information is available concerning whether any proteins associated with the VDR before DNA binding. There are some data suggesting that many hormone-binding receptors, including VDR, contain a binding site for calreticulin in the DNA-binding domain (43). Some reports indicate that calreticulin may bind to VDR and inhibit activation of vitamin D target genes (358), but the physiological significance of this is as yet unclear.

There have been numerous reports in recent years of rapid nongenomic effects of vitamin D on calcium transport, termed transcaltachia (241). The physiological importance of these data is unclear. Whether these effects are mediated by the VDR or a different protein is also unclear. The development recently of a mouse model in which the VDR has been ablated (365) may provide some insight into the relevance of these reports, and into whether the VDR plays any role in them.

Recently, the promoters of the mouse (150) and human (226) VDR have been isolated. Some studies have suggested that the VDR is upregulated by treatment with agents such as forskolin, which induce protein kinase A (183). Further study of these promoters will shed more light on the means by which VDR protein levels are regulated inside target cells.

C. Retinoid X Receptors and Other Coactivators

In the absence of cofactor proteins, the VDR is unable, at physiological concentrations of protein, to bind to most response elements that have been described. This has become clear from both in vitro gel retardation assay experiments and from experiments involving receptor expression in yeast (155 231 310). Work with the VDR and with related receptors indicated that the RXR is the required cofactor.

The RXR was isolated several years ago, and initially its ligand and functional importance were unknown (211). Subsequent work identified 9-cis-retinoic acid as a ligand for RXR (134). It also became clear that RXR plays a critical role in binding to DNA of several different receptors, including the VDR, the TR, the RAR, and the PPAR, among others (98 173). In the absence of RXR, it appears that none of these receptors binds efficiently to their response elements. Like other receptors, the RXR can be divided into functional domains, including an NH2-terminal A/B domain, a C domain which binds to DNA, and a COOH-terminal DE domain that binds ligand and activates transcription (58 210). Heterodimerization with other receptors appears to be mediated through two interfaces: one in the DNA-binding C domain (366) and another in the ligand-binding E domain (359). Interestingly, in structural studies, the RXR ligand-binding E domain was found to crystallize as a dimer, and the dimer interface occurred along helices 9 and 10, exactly where mutagenesis studies suggest RXR heterodimerizes with other receptors (32 261). This region of RXR from amino acids 389 to 429, termed the I box, is sufficient for strong interaction between RXR and TR, but the same sequence provides only weak interaction with VDR. Additional NH2-terminal RXR sequence is required for strong interactions with VDR (261), indicating a potential distinction between VDR and related receptors.

Most vitamin D response elements consist of two half-sites of the sequence AGGTCA separated by three bases (see sect. iii D). Because RXR is required for binding to the response element, the question arose as to whether one half-site was preferred by RXR, and if so, whether this preference was for the upstream or downstream site. For VDR, as for the other receptors that heterodimerize with RXR, it appears that RXR binds to the 5′-half-site and VDR the 3′-half-site (189 366), although there may be some exceptions to this rule. The polarity of this binding has been shown to be important for maximal gene activation, with significantly lowered transcription rates occurring if the response element orientation is reversed relative to the promoter start site. Binding of the VDR-RXR heterodimer to a response element is greatly increased by 1,25-(OH)2D3 when salt concentrations are in the physiological range, i.e., 100–150 mM (169). Interestingly, for many of the receptors that heterodimerize with RXR, the natural ligand of RXR, 9-cis-retinoic acid, may not enhance DNA binding. Some findings indicate that RXR may be unable to bind ligand while in a complex with another receptor (98). In addition, some in vitro reports indicate that the VDR-RXR complex is disrupted by addition of 9-cis-retinoic acid (205). This contrasts with work in transfected cells indicating that both 1α,25-(OH)2D3 and 9-cis-retinoic acid enhanced reporter gene expression from a vitamin D-responsive promoter (291). The reason for these conflicting results is unclear; it may be an experimental artifact, or it may be that the additional proteins that interact with the VDR-RXR heterocomplex in vivo, such as the coactivator/corepressor proteins, affect the conformation of the RXR and permit ligand to bind.

The VDR has been shown to interact directly with a growing number of other transcription factors, including transcription factor IIB (TFIIB). The interaction of VDR with TFIIB has been characterized in several reports (27 206). TFIIB is a 30-kDa protein that was originally isolated as a cofactor associated with TATA-binding protein (120). A region of VDR of ∼70 amino acids in the D domain has been reported to bind to a 43-residue NH2-terminal region of TFIIB (27 203). The sequence of TFIIB in this region is similar to that of a zinc finger, suggesting that the zinc finger motif may function in protein-protein interactions as well as DNA binding. Interestingly, one report suggests that binding of ligand to VDR causes dissociation of TFIIB from the receptor, which may indicate a complex interaction between proteins before the initiation of transcription (217).

Interactions between the VDR and other factors are less well characterized, although this work is progressing, particularly with regard to the coactivator proteins. Recent work indicates that VDR binds to SUG1/TRIP1, a nuclear protein that also binds to other receptors (355). It should be noted that the function of SUG1 in VDR-mediated transcription is unclear. SUG1 binds to the AF-2 domain of the VDR (216 355), but because SUG1 has been found to be a component of the proteasome complex in the nucleus, it may function primarily in receptor degradation.

The VDR has also been shown to bind to a growing number of coactivator proteins such as SRC-1 and TIF-1 (216 355). Other coactivator proteins that have been discovered recently, including ACTR (59) and p/CIP (345), join the rapidly enlarging group of such proteins. The coactivators seem to fall into a family with many features in common, including size (∼1,400 amino acids), structure (NH2-terminal Per-Arnt-Sim/basic helix-loop-helix domains, central receptor interaction domains) and activity (COOH-terminal histone acetylase domains). It seems likely that most, if not all, of the coactivator proteins will be found to interact with VDR and activate transcription. For example, both SRC-1 and TIF-1 have been shown to bind to the VDR (216 355), whereas nuclear receptor coactivator protein (ACTR) has been shown to enhance transcription from a vitamin D target gene in transfected cells (59). However, it is not yet clear whether all coactivators interact with VDR equally well; some may function more effectively than others in the vitamin D system. An additional possibility is that tissue-specific expression may be a significant factor in determining which coactivator works with the VDR. For example, the coactivator ACTR was shown to be expressed at relatively low levels in kidney, an important vitamin D target tissue (59). This may indicate that other coactivators play a more important role in transcriptional activation of vitamin D-controlled genes, at least in this tissue.

Whether VDR interacts with corepressor proteins to bring about downregulation of target genes is currently unclear. Like the coactivator proteins, the corepressor proteins appear to have many features in common. The corepressors SMRT and NCoR (142), for example, have considerable sequence similarity, although NCoR possesses a large NH2-terminal region lacking in SMRT. In addition, both proteins appear to repress transcription through interactions with the Sin3 proteins, which bind to histone deacetylase enzymes (131 233). However, unlike related receptors such as retinoic acid and TR, it is unclear whether the VDR is bound by corepressor proteins. At least one corepressor protein, NCoR, has been reported not to bind to VDR at all. NCoR binds to a conserved region of the TR and RAR, termed the CoR box, to exert its effects; binding of ligand to the TR or RAR causes dissociation of NCoR from receptor. NCoR does not bind to the CoR box region of the VDR, despite the similarity of the VDR sequence in this region to that of TR and RAR (142). This may explain why the VDR does not seem to exhibit dominant negative silencing seen with both the TR and RAR (262 286). The possibility remains that other corepressor proteins will be found to bind to the VDR and that this interaction is responsible for silencing at least some of the genes that are downregulated by vitamin D.

D. Vitamin D Responsive Elements

As mentioned in section iii A, responsive elements are the sequences of DNA, isolated from the promoters of vitamin D-responsive genes, that are bound by the VDR. One of the first such response elements isolated was that of the rat osteocalcin gene (84). Work with this and other vitamin D response elements, and with the response elements of other receptors, led to the demonstration that many response elements consisted of two repeats of the half-site sequence AGGTCA separated by several nonspecified bases (75 173 349). The variable that determines receptor specificity seems to be the number of nucleotides separating the half-sites. The RXR has been shown to have a preference for two half-sites separated by a single nucleotide, termed a DR-1 response element, whereas the VDR binds to a DR-3 element, the TR to a DR-4 element, and the RAR to a DR-5 element (349). Interestingly, almost all of the naturally occurring VDRE isolated from genes upregulated by the VDR have fallen into the DR-3 category. The negatively regulated genes, such as those for PTH and IL-2, may not follow this rule. Initial work with the PTH gene indicated that the response element in this case consisted of only a single half-site. However, more recent work indicates that the PTH VDRE may in fact consist of two half-sites (76), which may leave only a few exceptions, such as the IL-2 gene, to this rule (6).

One interesting aspect of the interaction between receptor and response element is the bend in the response element DNA that is induced by binding of the VDR-RXR heterodimer (151). This has been reported for binding of related receptors, such as the RAR, to their response elements (71). The degree of bending induced by the protein complex is reportedly on the order of 30°. Although this bend is much lower in magnitude than that reported for binding of the TATA-binding protein to DNA (240), it is still appreciable. It is interesting to speculate on the possible role of this induced bending, particularly in light of the finding that coactivator proteins and their cofactors acetylate histones. It may be that the combination of DNA bending and histone dissociation acts to make the promoter of the target gene more accessible to transcription factors before transcriptional activation.

Some recent reports have suggested that a greater degree of flexibility exists for binding of VDR to response elements. Work in vitro has suggested that VDR can bind to DR-3 response elements, such as the osteopontin VDRE, as homodimers (291). Other work suggests that TR and RAR can act as heterodimeric partners with VDR on a response element, instead of RXR (292). In addition, response elements that do not fall into the DR-3 category have been suggested to bind VDR, including DR-6 elements (264) and one element composed of an inverted repeat of two half-site elements separated by nine bases (292). At present, the significance of this work is unclear. For instance, it has been reported that although VDR can bind to the osteopontin response element as a homodimer, the homodimer is transcriptionally inactive and the VDR-RXR heterodimer is required for transcriptional activity (197). Thus the results of such in vitro experiments must be interpreted with caution and, if possible, in vivo. For example, the report of VDR heterodimerization with TR and RAR may potentially be evaluated through the use transgenic animals in which RXR has been ablated, to determine whether VDR-responsive genes can still be expressed. Unfortunately, ablation of RXR-α, the preferred partner for VDR, is a lethal mutation (90), but it is possible that inducible knockouts of the RXR-α will provide insight into whether these other receptors can heterodimerize with VDR and rescue the expression of VDR target genes.

E. Vitamin D-Dependent Genes, Their Roles, and Gene Complexity

Many different genes have been shown to be responsive to vitamin D. Most of these genes play a direct role in calcium endocrinology or bone formation and were the earliest examples of vitamin D response elements isolated. Examples of these include osteocalcin (84), osteopontin (243), PTH (251), the hydroxylases CYP24 (368) and CYP1α (298 316 332), and the calbindin genes (69). Many of these have been reviewed recently (75). The roles of many of these genes in calcium endocrinology have become more evident in recent years. For example, the bone protein osteocalcin is secreted by osteoblasts and represents the single greatest protein constituent of bone by mass. Expression of this protein is upregulated by vitamin D, and it may play a significant role in maintaining bone integrity (73). On the other hand, several genes have been shown to be downregulated by vitamin D, including the gene for PTH, which plays a critical role in controlling the levels of 1,25-(OH)2D3 in serum.

Several of the hydroxylase enzymes that act directly on 1,25-(OH)2D3 formation and degradation have been shown to be controlled at the transcriptional level by 1,25-(OH)2D3 . In particular, the hydroxylase enzyme CYP24 is one of the most highly regulated genes that responds to vitamin D. This enzyme represents the major means by which 1,25-(OH)2D3 is catabolized in vivo, and its importance is emphasized by recent gene knockout experiments in which loss of the enzyme was found to be lethal to at least 50% of the recipient mice (313). The CYP24 gene is highly upregulated by 1,25-(OH)2D3 and contains two sets of VDRE in its promoter (61 168 368). Both sets of elements are required for the gene to display maximum inducibility by vitamin D (168 368). It therefore appears that the presence of multiple copies of response elements are one way in which vitamin D target genes can be efficiently upregulated.

In contrast to the preceding example, the hydroxylase enzyme CYP1α, which is responsible for catalyzing the formation of 1,25-(OH)2D3 , is highly downregulated by its product. This enzyme, expressed in the proximal convoluted tubule of the kidney, is regulated with extreme stringency, although the exact details of the way in which this regulation is exerted are unclear. The recent cloning and expression of this enzyme open the door to a more complete understanding of its regulation, and details of the way in which the gene and protein expression are regulated are sure to be forthcoming (298 316 332).

Another set of genes that have been shown to be regulated by vitamin D are the calbindins. In mammals, the calbindin D9K is expressed primarily in intestine, whereas the calbindin D28K is expressed in kidney and other tissues. The calbindin D9K may play a role in absorption of calcium from the gut, whereas the calbindin D28K may be involved in reabsorption of calcium from glomerular filtrate, although this has not been definitively shown. The promoters of both of these genes have been cloned (74 153 331), and both have been reported to contain VDRE. Recent work on targeted gene disruption of the D28K has recently been reported in the mouse (311). This is not a lethal mutation, and the effects of this knockout on regulation of calcium levels will be of great interest in deciphering the role of this protein in calcium transport.

More recent and somewhat more surprising discoveries of genes that are regulated by vitamin D include the discovery of VDRE in the promoters for a number of genes not traditionally considered vitamin D targets. These include potential response elements in the promoters of several genes, including integrin β3 (52), fibronectin (264), atrial natriuretic factor (163 198), c-fos (48 290), PTHrP (182), and p21 (199). The significance of the vitamin D regulation of many of these genes is unclear and may prove an interesting area of study. In this regard, the recently developed mouse strain in which the receptor has been ablated may provide an interesting vehicle for assessing the contribution of the VDR to regulation of these genes (365).

F. Target Cell Metabolic Enzymes

The metabolism of vitamin D by target cells has been shown to occur in a number of tissues, including kidney, intestine, and bone. In kidney and intestine in particular, upregulation of the 24-hydroxylase enzyme in response to 1,25-(OH)2D3 treatment is rapid, occurring within 4 h, as shown by both Northern and enzymatic analyses (297). Cultured bone cells have also been shown to possess highly inducible 24-hydroxylase activity (202). These cells have been used to isolate catabolic products of 1,25-(OH)2D3 that can be placed in a logical order on a degradative pathway leading from the active compound to inactive excretion products (82). Thus it appears that the target cells for vitamin D contain the means to regulate its activity at the cellular level. The additional level of control that this target cell metabolism provides over local concentrations of 1,25-(OH)2D3 may be an important means by which tissues regulate the responsiveness of genes to vitamin D.

IV. RECENTLY DISCOVERED FUNCTIONS OF 1α,25-DIHYDROXYVITAMIN D3

A. Discovery of New Target Organs for 1α,25-Dihydroxyvitamin D3

The first insight into possible new functions of 1,25-(OH)2D3 beyond its actions in the regulation of calcium and phosphorus as described in section ii came as a result of studies involving the cellular and subcellular localization of radiolabeled 1,25-(OH)2D3 at high specific activity. This allowed the administration of physiological doses of 1,25-(OH)2D3 and its detection by frozen section autoradiography in target tissues (324 325). The result of this study in the late 1970s illustrated that not only was radiolabeled 1,25-(OH)2D3 localized in the nuclei of the enterocyte of the small intestine, the distal renal tubule of the kidney, and the osteoblasts of bone, but were found in tissues not previously considered targets of vitamin D action. These localizations were found specific inasmuch as the localizations could be prevented by the administration of unlabeled 1,25-(OH)2D3 but not 25-OH-D3 . Of great interest was the finding of 3H-labeled 1,25-(OH)2D3 in the nuclei of the islet cells of the pancreas, keratinocytes of skin, ovarian tissue, mammary epithelium, epithelial cells of the epididymis, certain neuronal tissue, promyelocytes, macrophages, and T lymphocytes. This unexpected development focused on the possibility that 1,25-(OH)2D3 may carry out functions not previously appreciated and in tissues not previously considered targets of vitamin D action. This result was followed by the demonstration of the VDR in the parathyroid gland (133 145), in skin (305), in the thymus (267), and in ovarian cells (88). Following this lead came the demonstration that the vitamin D hormone carries out functions in many of these tissues, and still other functions remain to be discovered.

B. Role of Vitamin D Hormone in the Parathyroid Gland

Perhaps the most well-established new function of 1,25-(OH)2D3 is in the parathyroid gland. Specific localization of 1,25-(OH)2D3 in the parathyroid gland (325) and the presence of VDR was provided by two independent laboratories (133 145). However, there had been considerable controversy as to whether the 1,25-(OH)2D3 could suppress secretion or production of the PTH (114 115 302). Early work failed to provide support for the idea that 1,25-(OH)2D3 could, in fact, suppress parathyroid secretion in isolated parathyroid glands. Furthermore, 25-OH-D3 was continuing to be used as a suppressant of secondary hyperparathyroidism in renal failure patients on dialysis (274). Notably, however, large amounts of this compound were needed to provide any degree of suppression, usually of the order of 20 μg/day. Suppression of parathyroid levels in such patients with orally administered 1,25-(OH)2D3 was also not dramatic. However, when an intravenous route was used, excellent treatment of secondary hyperparathyroidism of renal failure resulted (308). These results strongly suggested that 1,25-(OH)2D3 may have a direct action through its receptor in the parathyroid glands. Thus PTH secretion by isolated parathyroid glands or cells could be suppressed by the direct administration of 1,25-(OH)2D3 (8284). This suppression required some time to achieve. With the cloning of the preproparathyroid gene came a search for the possibility that this gene might be under control of 1,25-(OH)2D3 through its receptor. Analysis of mRNA encoding the PTH by Northern blot suggested a transcriptional suppression by 1,25-(OH)2D3 (284 296 304). Subsequently, work by Demay et al. (85) and Silver et al. (303) have shown unequivocally the presence of a VDRE in the promoter region of the parathyroid gene. Although this was first thought to be a single 6-base response element, subsequent work has revealed that in actual fact it represents a DR-3 in which two repeat sequences are indeed found at 123–108 bases 5′ from the transcriptional start site (76). This responsive element system causes a suppression of expression of the preproparathyroid gene promoter. Details of the mechanism are discussed in section iii.

In addition to the role of the vitamin D hormone in suppressing the parathyroid gene is the idea that the vitamin D hormone also suppresses proliferation of the parathyroid gland cells. The discovery by Suda and co-workers (1 334) that 1,25-(OH)2D3 can suppress the growth and stimulate cellular differentiation of the promyelocytes brought forth the idea that 1,25-(OH)2D3 might also suppress proliferation of the parathyroid gland cells. This rationale provides further impetus to the idea that renal osteodystrophy patients should be treated with 1,25-(OH)2D3 to prevent the development of hypertrophied parathyroid glands as well as to suppress secretion of the PTH.

C. Role of Vitamin D Hormone in Skin

After the discovery by Suda and co-workers (1 334) of the differentiative action of 1,25-(OH)2D3 on the promyelocytes came the application of this finding to the keratinocytes. Hosomi et al. (144) probably provided the first report that 1,25-(OH)2D3 induces keratinocyte differentiation. These results were found by the in vitro addition of 1,25-(OH)2D3 . Similar results were reported by Holick and co-workers (309), and it is now common knowledge that the keratinocytes are induced to differentiate by the in vitro addition of 1,25-(OH)2D3 . Exactly how important this differentiation effect of 1,25-(OH)2D3 is in vivo is difficult to assess. Certainly, vitamin D-deficient animals do not have a problem with keratinocyte differentiation. Thus hyperproliferation of the keratinocyte and failure to differentiate is not found in vitamin D-deficient animals (79; DeLuca, personal observations). It is possible that the differentiation may be triggered by more than one system providing a redundancy and hence the failure to observe a lesion in skin as a result of vitamin D deficiency. Together with the differentiation of the keratinocyte comes an inhibition of proliferation. This inhibition of proliferation has been utilized in the treatment of hyperproliferative diseases of skin as, for example, psoriasis (144 309). Both 1,25-(OH)2D3 and analogs especially MC903 of Leo Pharmaceuticals (also termed Dovenex) can be used as a significant therapy against psoriasis with as many as 70% patients responding to this treatment (44).

A number of studies have been carried out on the mechanism whereby 1,25-(OH)2D3 brings about the differentiation of the keratinocyte. During the differentiation process, there is an increase in the protein involukrin (326). There is an increase in transglutamanase activity (326) and cornified envelope formation (263). However, exactly how 1,25-(OH)2D3 induces differentiation of the keratinocyte and inhibits proliferation is unknown and remains to be investigated.

Finally, it has been proposed that the keratinocyte functions in a paracrine fashion in which 1,25-(OH)2D3 is produced by the keratinocyte itself, and this compound serves as a paracrine to stimulate differentiation of the keratinocyte (21). If this is true, it remains to be explained how vitamin D-deficient animals are able to obtain mature and functional keratinocytes. There have been very dramatic statements made that the keratinocyte is fully able to metabolize 25-OH-D3 to 1,25-(OH)2D3 but that it lacks the ability to 25-hydroxylate 1α-OH-D3 (21). In work carried out in the rat by two different laboratories (277 299), there was a failure to detect any 3H-labeled 1,25-(OH)2D3 in skin or other tissues in nephrectomized rats given the highest specific activity (160 Ci/mmol) of 3H-labeled 25-OH-D3 available. Thus, in normal rats, the keratinocyte was unable to produce significant amounts of 1,25-(OH)2D3 in vivo. On the other hand, there are a number of scientists who believe that this does, in fact, occur in vivo (21). Thus the role of 1,25-(OH)2D3 in the keratinocyte under normal physiological circumstances is unknown, and certainly the idea that the keratinocyte in vivo is able to produce 1,25-(OH)2D3 is not universally accepted. Recently, three laboratory groups have succeeded in cloning the 25-OH-D-1α-OHase (102 103 298 316 332), and one report used PCR techniques to clone it from human keratinocytes (102). This could support the presence of this enzyme in keratinocytes; however, Northern analysis of mouse skin (330) did not confirm 1α-OHase mRNA in this tissue. Additional work will be required before this area is clarified. However, certainly one has to consider the keratinocyte as a potential target of 1,25-(OH)2D3 to the extent that analogs are therapeutically effective against the hyperproliferative keratinocyte disease psoriasis.

D. Role of 1α,25-Dihydroxyvitamin D3 in the Immune System

The presence of the VDR in activated T lymphocytes was reported by Provvedini et al. (267), who also showed that resting lymphocytes do not express the VDR. Certainly the thymus, which is a repository of immature lymphocytes, is a source of VDR, since it was used as a source of receptor for measurements of 1,25-(OH)2D3 (138). Among the antigen-presenting cells, the macrophage that is derived from the monocyte has receptors for 1,25-(OH)2D3 (20 264). There is no doubt, therefore, that the immune system does indeed possess receptors to 1,25-(OH)2D3 . These results suggest a role for 1,25-(OH)2D3 in the immune system. However, the role is just now beginning to be defined. There have been many reports using peripheral blood lymphocytes that suggest that production of IL-2, tumor necrosis factor-α (TNF-α), and interferon-γ is suppressed by 1,25-(OH)2D3 (193). Furthermore, there are reports of increased synthesis of transforming growth factor-β in a number of immortalized cell lines in response to 1,25-(OH)2D3 in vitro (177). In fact, there are many conflicting reports regarding the in vitro response of the lymphocytes to 1,25-(OH)2D3 . This review focuses on what is known concerning the in vivo responses of the immune system in the absence and presence of 1,25-(OH)2D3 and its analogs.

In a study of delayed hypersensitivity, which is a T-helper cell lymphocyte-mediated response, vitamin D deficiency markedly reduces the ability of the mouse to respond (364). However, large amounts of 1,25-(OH)2D3 and its analogs will also suppress the delayed hypersensitivity response (363). These results illustrate that the T-helper cell lymphocyte system is vitamin D responsive but that both immunostimulation and immunosuppression can be found in the in vivo situation. Currently, there is little or no evidence to support the idea that B cell-mediated immunity is affected by 1,25-(OH)2D3 .

The most dramatic results obtained to date in the immune system are those found in several autoimmune diseases. Most notably is the autoimmune disease known as experimental autoimmune encephalomyelitis (EAE) that can be induced in B10.PL mice or SJL mice by the injection of myelin basic protein together with small amounts of pertussis toxin (49). This treatment induces a multiple sclerosis-like disease (EAE). This disease is aggravated by agents that stimulate T-helper (Th)-1 cells and that increase interferon-γ and TNF-α secretion (280). Injections once every other day of 1,25-(OH)2D3 in some cases produced a reduction in the severity and in other cases produced very slight responses (36 196). However, since 1,25-(OH)2D3 has a lifetime measured certainly in hours, the administration of this compound three times a week or every other day would not be expected to be sufficient (116). More recently, EAE in B10.PL mice could be completely prevented by the addition of 1,25-(OH)2D3 to the diet of such mice. Furthermore, if the hormonal substance was begun after the EAE had begun, the development of symptoms could be prevented (Fig. 9) (49). Current results strongly suggest that the vitamin D hormone and its analogs are functioning by stimulating Th-2 T-helper cells to produce transforming growth factor-β1 and IL-4 that might serve to suppress the TNF-α and interferon-γ production by Th-1 cells (51). Similar results were obtained with another autoimmune disease, rheumatoid arthritis induced by Borrelia burgdorferi or Lyme's disease (50) or arthritis produced by the injection of collagen into susceptible mice (50). Some improvement has also been reported for the autoimmune induction of diabetes in the nonobese diabetic mouse (218). The idea, therefore, that the vitamin D compounds may play a role in modulating the immune system is strongly supported. Furthermore, there is the idea that 1,25-(OH)2D3 and its analogs might become useful in the prevention or treatment of autoimmune diseases.

Fig. 9.

Fig. 9.1,25-(OH)2D3 prevents progression of experimental autoimmune encephalomyelitis (EAE) in B10.PL mice. When individual mice showed EAE symptoms of 1 or greater, they were given an intraperitoneal injection containing 300 ng 1,25-(OH)2D3 dissolved in ethanol (•) or mock injected with an equivalent amount of ethanol (○). At time of treatment, diet was also changed to experimental diet containing no additional vitamin D or to same diet containing 20 ng⋅day−1⋅mouse−1 1,25-(OH)2D3 .


Of some interest is the idea that the immunomodulatory action of vitamin D might be useful in the management of transplant rejection. Initial work indicates that the vitamin D hormone and its analogs would only be marginally useful in prevention of transplant rejection and might only be considered as an adjunct to the well-established cyclosporin method of preventing transplant rejection (195). However, more recently, the administration of 1,25-(OH)2D3 and its analogs in the diet and thus their provision in a continuous manner has shown that it can be much more effective than cyclosporin itself in prevention of transplant rejection (D. A. Hullett, M. T. Cantorna, C. Redaelli, J. Humpal-Winter, H. W. Sollinger, and H. F. DeLuca, unpublished data). It has been used effectively to prolong transplant survival beyond that achieved by cyclosporin in mouse embryonic cardiac transplants and in a vascularized cardiac transplant model in the rat. Of great interest is that the prevention of transplant rejection by the vitamin D compounds was not accompanied by a bone loss phenomenon as is the case with cyclosporin A and, furthermore, was not accompanied by the danger of opportunist infection (M. T. Cantorna, D. A. Hullett, C. Redaelli, C. R. Brandt, J. Humpal-Winter, H. W. Sollinger, and H. F. DeLuca, unpublished data). However, the exact role of how the vitamin D hormone affects T-lymphocyte cell populations, cytokine secretion, and production is not entirely known. There has been a suppressive element reported for the IL-2 gene for the VDR. Whether this is a significant mechanism or not remains to be determined.

E. Islet Cells of the Pancreas

The intriguing finding of a nuclear localization of 1,25-(OH)2D3 in the islet cells of the pancreas has led to a speculation that vitamin D might be involved in glucose regulation of insulin secretion (324 325). The presence of a VDR in these cells by now is well accepted, but it is unclear as to what, if any, role vitamin D plays in the functioning of the islet cells. Initial results reveal that vitamin D-deficient rats were unable to respond to a glucose challenge by secreting appropriate amounts of insulin (65). Although this could be corrected by the administration of 1,25-(OH)2D3 , other studies suggested that this was mediated by the action of vitamin D in raising plasma calcium concentration (66). The problem, therefore, is that it is unclear whether the regulation of insulin secretion may have a vitamin D component because of the direct actions of calcium concentration on the functioning of the islet cells. Nevertheless, the idea that inappropriate responses of the islet cells to a glucose challenge suggests that a role of the vitamin D hormone is involved. These findings must also be coupled to the experiments carried out in the nonobese diabetic mouse in which the onset of diabetes is delayed or retarded by the administration of 1,25-(OH)2D3 (218). The relationship between vitamin D and diabetes is certainly worthy of additional investigation.

F. Role of Vitamin D and Reproduction

Because of the focus of scientists on the role of vitamin D in calcium homeostasis and regulation of phosphate metabolism, the possibility that vitamin D might be involved in reproduction has been largely ignored in the past. However, during the course of producing vitamin D-deficient embryos came the observation that female reproduction is markedly and significantly diminished in vitamin D deficiency (122). Thus a reduction in fertility of 80% was found and could not be corrected by correcting the hypocalcemia (191). This defect, therefore, is quite clearly one related to an absence of the vitamin D molecule. Similar female reproductive failure was noted in the transgenic receptor knockout mice described by Yoshizawa et al. (365). Why there is a failure in the female reproductive system is not known except that it is clear that ovarian cells contain VDR (88). Furthermore, ovarian cells in vivo also accumulate 1,25-(OH)2D3 (324 325). Other organs involved in female reproduction, i.e., the hypothalamus, also contain VDR. It has been suggested that the ovary therefore is a target of vitamin D action. In experiments carried out in the authors' laboratory, the VDR levels of the ovary are increased during estrus and diminished during proestrus (348). These results suggest that female rats may not ovulate properly. The infertility brought about by vitamin D deficiency in the female rat can be easily corrected by the administration of 1,25-(OH)2D3 (191). These results suggest a role for the vitamin D hormone and its receptor in ovarian function; however, the mechanism remains to be found.

In the case of male reproduction, vitamin D deficiency also reduces the effectiveness of the male (190). However, this diminished male fertility can be corrected by merely providing additional calcium, raising plasma calcium concentration which in turn restores fertility (190). Thus there does not appear to be a role for the vitamin D hormone in spermatogenesis and male reproduction.

G. Does Vitamin D Play an Essential Role During Embryonic Development?

This question intrigued investigators in this field a couple of decades ago. The interest in this possibility was intensified when Suda and colleagues (1 334) demonstrated for the first time the role of 1,25-(OH)2D3 in promyelocyte differentiation. This was followed by work which illustrated that cancer cell lines also responded by differentiating in response to 1,25-(OH)2D3 (91). This then provided the basis for believing that the vitamin D hormone is also a developmental hormone. However, vitamin D-deficient rats are able to reproduce albeit at only 20% of normal, but the embryos develop approximately normally to parturition (121). Furthermore, the provision of sufficient nutrients to the pups allows them to develop approximately normally until weaning (37). It is at this time that the dependency on vitamin D becomes very obvious. These results strongly suggest that vitamin D does not play an essential role in differentiation and development during embryogenesis. It does not exclude, however, that the vitamin D hormone plays an important role in terminal differentiation of specific cell types, as for example the differentiation and formation of osteoclasts. It does argue, however, that 1,25-(OH)2D3 is not involved in the development of major organ systems. These results are supported very strongly by the receptor knockout mice and by the vitamin D-dependency rickets type II cases which are the human example of receptor knockouts. In these cases, the fetuses develop normally and, in fact, appropriately fed, the animals will survive to day 55 (365). If in the absence of receptor the major organ systems develop normally, it provides little support for the idea that vitamin D is an important developmental vitamin. One might argue that there are nongenomic actions of 1,25-(OH)2D3; however, there is insufficient evidence for this concept at this time as well (195). Therefore, it is the view of these authors that 1,25-(OH)2D3 is not a major developmental hormone at least involved in the development of major organ systems during embryogenesis.

H. Summary

Scientists are just beginning to investigate the new and previously unappreciated functions of the vitamin D hormone. They include functions in the parathyroid glands, the keratinocytes of skin, the T cells and macrophages of the immune system, the islet cells of the pancreas, and ovarian cells of the female. There is the possibility that additional functions will be found in such systems as the islet cells of the pancreas, mammary cell epithelium, and cells in the hair follicle system.

V. VITAMIN D ANALOGS

A. Development of New Analogs of 1α,25-Dihydroxyvitamin D3

The discovery of 1,25-(OH)2D3 and the other metabolites of vitamin D in the early 1970s immediately led to their chemical synthesis and to the first generation of compounds suitable for hormone replacement therapy (13 294). In addition to 1,25-(OH)2D3 , most notable among these were the prodrugs 1α-OH-D3 and 1α-OH-D2 , which were designed to overcome the need for a functional kidney containing the 1α-hydroxylase enzyme, instead, depending only on the need for 25-hydroxylation (193 258). Synthetic 1,25-(OH)2D3 and 1α-OH-D3 proved to be valuable calcemic agents capable of the correction of hypocalcemia and bone abnormalities (rickets and osteomalacia) resulting from a variety of causes from simple vitamin D deficiency to chronic renal failure. There were early indications that patients with osteoporosis might benefit from the use of an active vitamin D preparation. As a result, it is not surprising that clinical trials of 1α-OH-D3 (129 254), 1α-OH-D2 (106), and 1,25-(OH)2D3 (7 9 108 257 343 351) have been undertaken. Modest gains in bone mineral density and reductions in fracture rates were reported in many of these studies, and this subject has been reviewed recently by Cali and Russell (44) and Seeman et al. (293). Consequently, 1,25-(OH)2D3 and 1α-OH-D3 are currently used to treat osteoporosis, particularly in Japan and Europe, where dietary calcium intakes are often lower and hypercalcemic side effects not so prevalent.

Although vitamin D analogs continued to be synthesized in the late 1970s, the emergence of the newer nonclassical functions of 1,25-(OH)2D3 , particularly the discovery of its role in controlling cell proliferation and differentiation (228 252) (see sect. iv), led to an acceleration in the pace of the search for new analogs. The chemical synthesis of new analogs of 1,25-(OH)2D3 became a major priority with the hope that the calcemic properties of 1,25-(OH)2D3 might be separated from the antiproliferative cell-differentiating properties (119 157). Hundreds of molecules have now been synthesized in pursuit of this goal, and a couple of recent reviews have attempted to classify or list all published vitamin D analogs in a detailed fashion (20 31 47 104). For the sake of brevity, we have selected some of the more promising of these that are under development by the pharmaceutical industry in Table 1.

Those selected are not the only analogs being developed, but all are at some advanced stage of animal or clinical testing. From perusal of the compounds in Table 1, the reader will immediately notice that the side chain appears to be a popular target for modification, and there is a rational basis for this, namely, that the VDR is relatively tolerant of changes in this part of the molecule. However, since 1980, the synthetic chemist has experimented with modification of every part of the vitamin D molecule in an attempt to accentuate either the antiproliferative or calcemic properties. The full extent of these synthetic modifications is best displayed as the annotated structure illustrated in Figure 10. Probably the most successful 1,25-(OH)2D3 analog developed thus far is the Leo antipsoriatic drug calcipotriol, also known as MC903, which was the first vitamin D analog to reach the market for any condition other than hypocalcemia or bone disease (45). When given orally, calcipotriol is ineffective due to the fact that it is rapidly broken down (24 354). When given topically as an ointment, calcipotriol survives long enough to cause improvement in 70% of psoriasis patients (18 181). Both 1,25-(OH)2D3 and calcipotriol are believed to be effective in psoriasis because they block hyperproliferation of keratinocytes, increase differentiation of keratinocytes, and help suppress local inflammatory factors through their immunomodulatory properties. Like several other analogs [1,24S-(OH)2D2 and 1,24-(OH)2D3] currently being pursued, calcipotriol has no 25-hydroxyl group but instead possesses a 24-hydroxyl group which appears to act as a surrogate 25-hydroxyl function. Other analogs being tested, such as 22-oxacalcitriol (OCT) (232), 1,25-(OH)2-16-ene-23-yne-D3 (11), and EB1089 (23 161), all contain modifications at C-22 and C-23 which affect the metabolic stability of the molecules without adversely altering the VDR binding. 22-Oxacalcitriol is in development for use in secondary hyperparathyroidism, while EB1089 is in clinical trial for treatment of breast cancer. Another successful modification has been the addition of carbons to the backbone or terminus of the side chain (termed homologation), examples being EB1089 and KH1060 (124). The latter analog epitomizes the latest generation of vitamin D analogs, in that it contains a combination of modifications; in the case of KH1060, four modifications: an extra C at C-24a; additional terminal carbons at C-26a and C27a; a 22-oxa group; and epimerization at the C-17 to C-20 bonds giving a side chain with the opposite orientation. The synthesis of these 20-epi-analogs (46 320), as they are termed, represents a successful experiment in vitamin D design because it led to a family of compounds, many with improved VDR binding compared with the natural side chain. Very recently, several groups (188 350) have used this experience to go one step further and synthesize so-called “butterfly” or “Gemini” analogs with both 20-epi (20S) and normal (20R) side chains. Although little is known at this time about the biological success of this experiment, it is certain to lead to other more exotic double-side chain molecules with symmetrical and asymmetrical structures around C-20 (Fig. 11). In the asymmetrical double-side chain analog (Fig. 11) featuring one 23-oxa side chain and one normal side chain, X-ray structural analysis (188) has revealed the orientation of the two side chains and a measure of the range of conformational structures that can be accommodated by the ligand binding pocket of the VDR, since it is known to bind both 20S- and 20R-analogs.
Fig. 10.

Fig. 10.Structure of 1,25-(OH)2D3 showing location of carbon atoms or carbon-carbon bonds modified in various vitamin D analogs constructed to date. Each circled carbon or bond is connected to a rectangular box which indicates 1 or more successful modifications introduced into analog at that site. For sake of space, model does not include all modifications made but seeks to give reader a flavor of extent of modifications attempted covering whole molecule. Certain popular analogs contained in Table 1 frequently contain combinations of such modifications.


Fig. 11.

Fig. 11.X-ray crystallographic structure depicting side chains and D-ring region of a double-side chain vitamin D analog. (Figure courtesy of Dr. Martin Calverley. Adapted from Vitamin D: Chemistry, Biology and Clinical Application of the Steroid Hormone, edited by A. W. Norman, R. Bouillon, and M. Thomasset. Riverside: Univ. of California, 1997.)


Equally interesting and perhaps more practical from a biological activity perspective are analogs with modifications of the cis-triene, either changing the conformation of the double bonds to create molecules with an altered cis-triene (250) or analogs that lack the C-10 to C-19 methylene group altogether such as 19-nor-1,25-(OH)2D3 or 19-nor-1,25-(OH)2D2 (260). In a sense, these latter compounds represent an extension of the experiment begun over 60 years ago (97) when German chemists successfully stabilized the labile cis-triene of one of the photoirradiation products of vitamin D and created the world's first vitamin D analog, dihydrotachysterol (DHT). This analog DHT has an A-ring rotated 180° and has lost the double bond between C-10 and C-19 but retains the C-19 as a methyl group not a methylene, as is the case in almost all other vitamin D compounds. Dihydrotachysterol is metabolized to 25-hydroxylated and 1,25-dihydroxylated forms, both of which bind to VDR and are biologically active, explaining why DHT was once the analog of choice for hypocalcemia, particularly that associated with chronic renal failure (158 268). The experiments with 19-nor-1,25-(OH)2D2 and DHT show that the 10–19 double bond and the C-19 carbon are probably superfluous for vitamin D-dependent gene expression. As with several other analogs, 19-nor-1,25-(OH)2D2 appears effective in the treatment of hyperparathyroidism associated with chronic renal failure (307).

Further excursions into A-ring or C- and D-ring modifications are currently underway. One molecule into clinical trial for the treatment of osteoporosis is the Chugai analog ED-71, which is modified at C-2 of the A ring by the addition of a hydroxypropoxy group (241), a feature which appears to extend its lifetime in vivo. An extensive collaboration between the De Clerq and Bouillon laboratories has generated many new analogs with modification in the C and D rings, although at present many of these are only partially tested biologically (29 78). One interesting spin-off from the latter research is the synthesis of novel E-ring vitamin D analogs in which the C-8–14–13–17 backbone is retained with the loss of the C and D rings of the vitamin D nucleus but involving the creation of a new stabilizing E ring utilizing C-13, C-12, C-21, C-20, and C-17. The resulting 1,25-dihydroxylated E-ring analog KS176 has 10-fold lower VDR binding and 5-fold lower DBP binding than 1,25-(OH)2D3 but is virtually inactive in calcemic assays in vivo (29). It will be interesting to see why E-ring analogs show good in vitro activity but appear to be devoid of in vivo activity.

B. Factors That Alter the Action of Vitamin D Analogs

From the synthesis of literally hundreds of analogs has emerged a clearer understanding of the factors that are important in changing the action of the specific analog relative to 1,25-(OH)2D3 . These are listed in the following sections.

1. Activating enzymes

The natural hormone of vitamin D is activated by sequential steps in the liver and kidney as described in section i. The distribution and pharmacokinetics of 1,25-(OH)2D3 depend on these steps occurring efficiently at substrate concentrations that circulate bound to DBP and/or are available in the respective tissues. The strategy with most synthetic analogs is to synthesize a molecule with both 25- and 1α-hydroxyl groups and miss out the need for these activation steps. The disadvantage of this approach is that the body loses its ability to regulate the amount of the active principle in the blood and target cell, with the success of the therapy being heavily dependent on the pharmacokinetics of the specific analog.

On the other hand, the use of prodrugs such as 1α-OH-D3 circumvents some of these problems by introducing a step of activation back into the process, albeit by a different regulatory enzyme. This activating enzyme is in essence a “slow-release” mechanism for delivering the same “active form,” 1,25-(OH)2D3 , with altered pharmacokinetics. This is because the activating enzyme has been changed from the tightly regulated renal 1α-hydroxylase to the loosely regulated liver 25-hydroxylase. Although this was believed to be a simple strategy to get around dependence on the easily damaged renal enzyme when it was conceived in the 1970s, the emergence of CYP27 as one of the enzymes responsible for the 25-hydroxylation of vitamin D3 and 1α-OH-D3 has complicated this story. CYP27 assumes an even greater importance for the activation of 1α-OH-D3 because several studies have shown that this cytochrome P-450 prefers 1α-OH-D3 as a substrate over vitamin D3 itself (8 119). CYP27 is found in extrahepatic sites including kidney, bone, and endothelial lining cells. In theory, this makes the activation of 1α-OH-D3 possible in many sites of the body in addition to the liver and opens up the possibility that 1,25-(OH)2D3 may be formed from 1α-OH-D3 in target cells like the osteoblast (147) to act locally on vitamin D-dependent processes. This theory remains to be proven in vivo.

Another aspect of the activation strategy that was overlooked at first glance was the potential of producing multiple active forms from the same administered prodrug. The analog 1α-OH-D2 has consistently proven to be less toxic than its D3 counterpart in animal studies and clinical trials (93 306). The explanation for these findings remains obscure but may be due to the fact that 1α-OH-D2 gives rise to two biologically active products, 1,25-(OH)2D2 and 1,24-(OH)2D2 (320), each with its own set of slightly different pharmacokinetic parameters compared with 1,25-(OH)2D3 (175). CYP27 is capable of the efficient synthesis of 1,24-(OH)2D2 (119). The enzyme responsible for the synthesis of 1,25-(OH)2D2 remains unknown. Whether this strategy of delivering a prodrug which because of its structure is activated by different enzymes with altered distribution or gives rise to more than one active metabolite offers therapeutic advantages remains moot at this time.

2. Binding to DBP

It has been recognized since the discovery of the various vitamin D metabolites that the affinity of an analog for DBP is inversely correlated with its rate of clearance from the bloodstream. Those metabolites with the strongest affinity for DBP [e.g., 25-OH-D3 , 24,25-(OH)2D3 , and 25-OH-D3-26,23-lactone] possess the longest half-life values in the blood on the order of days (17 26). Those metabolites such as 1,25-(OH)2D3 have lower affinity and half-life values on the order of hours. These basic concepts were confirmed by our experience with analogs of 1,25-(OH)2D3 .

Clearly, some modifications are detrimental to DBP binding such as the addition of a 1α-hydroxyl group. Other modifications such as the addition of the double bond between C-22 and C-23 or the addition of a C-24 methyl group, as in the metabolites of vitamin D2 , do little to affect binding to mammalian DBP but dramatically lower the binding of these compounds to avian DBP (17). However, the findings with D2 compounds with mammalian DBP go against the general rule, and most 1α-hydroxylated, side chain-modified analogs have inferior DBP binding compared with 1,25-(OH)2D3 . Calcipotriol, OCT, EB1089, KH1060, and 20-epi-1,25-(OH)2D3 all bind DBP with an affinity at least one order of magnitude less than 1,25-(OH)2D3 (89 171). One analog, the A ring-substituted, 1α-hydroxylated compound ED-71, binds DBP more strongly than the natural hormone, a fact which dramatically increases its survival time in the blood and hence its usefulness (241).

The consequences of altered DBP binding are altered tissue distribution of analogs and changes in the rates of their metabolic clearance and cell uptake. Reduction in the affinity of DBP binding can lead to transport on alternate carriers, either albumin or lipoproteins. It has been claimed that OCT derives part of its usefulness for treatment of hyperparathyroidism from its accumulation in parathyroid tissue in vivo, possibly caused by an altered plasma transport on lipoprotein (176).

Second, the rate of metabolic clearance is affected by altering DBP affinity. Consistent with the findings for vitamin D metabolites, a reduction in DBP binding for a given vitamin D analog leads to even more rapid excretion than that observed for 1,25-(OH)2D3 . Calcipotriol binds DBP with an affinity two to three orders of magnitude lower than 1,25-(OH)2D3 (171). When given systemically, calcipotriol is metabolized and excreted so rapidly that it has no biological effects even at 50 times the dose used for the natural hormone. As described above, it is only effective because it is topically administered in an ointment and is rapidly cleared (and metabolized) before it can reach tissues involved in calcium homeostasis. Thus DBP binding is an important parameter in dictating in vivo biological activity.

Third, the affinity of the analog for DBP influences the rate of cell uptake. Given that it is the “free” fraction of the plasma-borne ligand which is believed to gain access to the target cell and function by binding to the receptor (VDR) in the classical steroid hormone model (see sect. iii), it is therefore no surprise that this might also be the case for vitamin D analogs. Many analogs have the same affinity for VDR as 1,25-(OH)2D3 , which when combined with a lower affinity for DBP than 1,25-(OH)2D3 creates an even greater downhill gradient into target cells. This has been clearly demonstrated in cultured cells in vitro where the composition of the extracellular fluid can be easily modified. Whereas DBP in fetal calf serum slightly retards entry of 1,25-(OH)2D3 , it has no discernible effect on the entry of 20-epi-1,25-(OH)2D3 , a notoriously poor binder to DBP (86). There is no doubt that poor DBP binding contributes positively to the promising profile of many vitamin D analogs in biological assays in vitro.

On the other hand, when this effect of DBP, greater access to target cells, is considered along side its other effect, more rapid clearance at the liver, in the animal in vivo, the net effect is usually not so advantageous. This is presumably because these two effects are counteractive and the vitamin D analog cannot be active unless it reaches the target cell in sufficient concentrations to gain greater access and produce its target cell biological effects. In the case of analogs such as ED-71, chemists have created a longer lasting analog that may give reduced biological activity but for a longer period of time. In essence, they have altered the pharmacokinetics again.

3. Binding to VDR-RXR-VDRE complexes

Section iii of this review has established that 1,25-(OH)2D3 is able to work through a VDR-mediated genomic mechanism to stimulate transcriptional activity at vitamin D-dependent genes. Whether 1,25-(OH)2D3 works through other nongenomic mechanisms to produce physiologically relevant effects is a theory that, in our opinion, remains unproven (244).

Much evidence exists to support the viewpoint that vitamin D analogs operate by mimicking 1,25-(OH)2D3 and using a genomic mechanism. The first clue that vitamin D analogs can work through a VDR-mediated transcriptional mechanism came 15 years ago from the bone resorption studies reported by Stern (318). Stern (318) showed that there exists a strong correlation between chick intestinal VDR binding of an analog and its potency in a 45Ca rat bone resorption assay. This suggests that a vitamin D analog is only as good as its affinity for the VDR. Certainly, there are some apparent exceptions to this general rule, but these are not discrepancies that cannot be explained by considerations such as DBP binding or pharmacokinetics. Preliminary results with the analogs KH1060, EB1089, and 20-epi-1,25-(OH)2D3 (25) suggested that they might be active in immunoregulatory roles at concentrations orders of magnitude below their affinities for the VDR (e.g., at as low as 10−15 M for KH1060 whereas it binds VDR at 10−11 M). More recent results suggest that KH1060 (55) is consistently only one or two orders of magnitude more potent than 1,25-(OH)2D3 in gene transactivation assays, a difference that could be explained by using the transcriptional model of analog action in conjunction with pharmacokinetic considerations rather than discarding the genomic hypothesis all together.

With the finding that the liganded VDR probably functions transcriptionally as a vitamin D-VDR-RXR heterodimer (128 322) but can form VDR-VDR homodimers (53 54) has come the question as to whether vitamin D analogs might differ from 1,25-(OH)2D3 and favor association with VDR-VDR homodimers. Although it remains possible that vitamin D analogs bound to VDR-VDR homodimers might preferentially act at certain VDRE in selected vitamin D-dependent genes (e.g., those associated with immunoregulatory or cell regulatory responses), currently the data do not support this (128 322). It should be noted that in a recent review (53) it was theorized that 1,25-(OH)2D3 (and thus vitamin D analogs) might be able to act through as many as 14 different heterodimeric response element combinations to modulate gene expression, but the physiological relevance of the majority of these has yet to be established.

To add to the complexity of the target cell action of 1,25-(OH)2D3 , and thus that of vitamin D analogs, is the type and context of the VDRE involved (40). The work of Williams and co-workers (40 166 360) shows that compared with 1,25-(OH)2D3 , KH1060 and EB1089 show different patterns of gene activation in bone marrow, osteoblastic cells (ROS17/2, ROS25/1, and UMR106), and intestine (HT-29 and Caco-2) which appear to be gene and cell specific. These data are probably explained by the changes in available transcription factors acting at VDRE in different tissues and at different stages of cell differentiation (317). Furthermore, there have been reports of a nonclassical VDRE in the promoter of the c-fos gene (48) that could theoretically respond differently to “noncalcemic” and “calcemic” analogs, but this has not been demonstrated experimentally. Furthermore, Morrison and Eisman (229) showed that a noncalcemic analog such as calcipotriol is capable of transactivating a calcemic VDRE such as the human osteocalcin promoter VDRE placed in front of the chloramphenicol acetyltransferase reporter gene and stably transfected into ROS17/2 cells. One interpretation of this experiment is that a noncalcemic analog with good VDR affinity is just as calcemic as 1,25-(OH)2D3 if it can be delivered to the target cell.

A more recent hypothesis put forward to explain the unique properties of certain vitamin D analogs is that they generate different conformations of the ligand-VDR-RXR heterodimer which result in altered binding at the VDRE. Such subtle changes in protein conformation may still be detected by indirect biochemical means such as altered sensitivity to protease digestion. Recent studies have revealed differences between 1,25-(OH)2D3-VDR-RXR and 20-epi-1,25-(OH)2D3-VDR-RXR or KH1060-VDR-RXR complexes to trypsin digestion, suggesting subtle changes in the conformation of the ligand binding domain after ligand binding (259 352). The increased resistance to trypsin digestion provided to VDR by KH1060 binding is mimicked by the most active metabolites of KH1060, suggesting that certain receptor conformations might correlate with increased biological activity (353). The use of truncated VDR mutants has allowed the work on KH1060 to be extended and to potentially reveal the amino acid residues involved in these conformational differences and in ligand binding (200). In the future, it is to be expected that the generation of large quantities of the ligand-binding domain of the VDR using bacterial or baculovirus-generated material (323) will permit the determination of the three-dimensional structure with and without ligand [1,25-(OH)2D3 or vitamin D analog] by NMR or X-ray crystallographic analysis. For now, we can draw on the experience gained from related fields. If the findings from X-ray crystallographic studies of various retinoid receptors (apo-RXR and holo-RAR ) are relevant to VDR, then the orientation of the COOH terminus containing the AF-2 domain will turn out to be dramatically altered upon vitamin D analog binding and to be very important to the interaction with the other proteins of the transcription initiation complex (279). Moras and co-workers (361) have already used computer modeling based on his RAR model to predict the orientation of ligands including vitamin D analogs in the binding pocket of the VDR. He predicts that 1,25-(OH)2D3 and various analogs bind with side chain deep into the ligand-binding pocket and A ring close to the putative “trap-door” (AF-2 domain). Therefore, the observed differences in protease resistance induced by potent 20-epi compounds such as KH1060 (259 352) could be important clues that the AF-2 domain of the heterodimer-analog complex is in a slightly different conformation to that of the heterodimer-1,25-(OH)2D3 complex. As a result, these complexes may recruit additional transcription factors (see sect. iii) that may result in qualitative or quantitative differences in gene expression.

There is some evidence from surface plasmon resonance studies that a series of Roche analogs causes differences in the association and dissociation rates of VDR-RXR heterodimers to artificial VDRE-containing oligonucleotide anchored to gold-coated chips (67). Whether this translates into increased stability of specific complexes on the vitamin D-dependent gene promoter in vivo remains to be proven. In summary, researchers continue to probe the many steps of the gene transcription activation process in the hope of finding a fundamental difference between 1,25-(OH)2D3 and its analogs that may explain their improved potency or their action at only selected genes. Thus far, they have more questions than answers.

4. Target cell and liver catabolic enzymes

As stated in section i, much evidence has accumulated to support the hypothesis that 1,25-(OH)2D3 is subject to target-cell catabolism and side chain cleavage via a 24-oxidation pathway to calcitroic acid. CYP24 carries out multiple steps in the C-24 oxidation process, is vitamin D inducible, and is present in many (if not all) vitamin D target cells. We have postulated that the purpose of this catabolic pathway is to desensitize the target cell to continuing hormonal stimulation by 1,25-(OH)2D3 (203). One can also ask the question: Are vitamin D analogs subject to the same catabolic metabolism? And, if not, are other drug-catabolizing systems present within vitamin D target cells to inactivate the vitamin D analog?

We now have a good idea as to the answer to these questions, but the answers are analog specific. Certainly, there are metabolism-sensitive analogs such as calcipotriol and OCT that are metabolized very rapidly by target cell enzymes to clearly defined and unique metabolites (213 215) that resemble products of the 24-oxidation pathway for 1,25-(OH)2D3 . The provision of such features as a 22-oxa group seems to accelerate the rate of metabolism of OCT, and the 24-hydroxyl group of calcipotriol is a site for deactivation. The rapid metabolism of calcipotriol by cultured keratinocytes in vitro (215) correlates well with the in vivo findings that it is rapidly broken down when administered topically and fails to reach tissues involved in calcium homeostasis (24). One does not need to invoke hepatic metabolism as the principal site of inactivation, even though this tissue can also break down calcipotriol (24 215).

Analogs that appear to be slowly metabolized by target cells include EB1089 and 1,24S-(OH)2D2 (156 170 295). Both are considered more calcemic than either calcipotriol or OCT but not excessively so. These data are consistent with metabolism-resistant analogs having a longer half-life inside the target cell as compared with 1,25-(OH)2D3 and a longer half-life in pharmacokinetic studies (30 175), and possibly a higher biological activity in vivo, although this is tempered by inferior DBP binding and more rapid clearance from the bloodstream. Interestingly, both EB1089 and 1,24S-(OH)2D2 are blocked at the C-24 position and cannot enter the C-24 oxidation pathway, leaving C-26 as the new site of hydroxylation, albeit at a slower rate (156 170 295). Another potent analog blocked in the C-24 position is 1,25-(OH)2-16-ene-23-yne-D3 , which is also subject to 26-hydroxylation (288). Studies with EB1089 using transfected human CYP27 suggest that this cytochrome is not responsible for the 26-hydroxylation observed (295). However, Suda and co-workers (227) have concluded that the other candidate cytochrome P-450, CYP24 may be responsible for 26-hydroxylation of the 24-blocked analog 24,24-F2-1,25-(OH)2D3 inside kidney cells.

Another metabolic complication brought on by the use of vitamin D analogs is the potential for creating “active” rather than “inactive” products as a result of target cell enzymes. Three examples of this have now been documented, and this may be more common than we think. The potent calcemic analog 26,27-F6-1,25-(OH)2D3 is converted both in vivo and in vitro, presumably by the cytochrome CYP24 into 26,27-F6-1,23,25-(OH)3D3 . Sasaki et al. (287) have shown that this metabolite which accumulates in bone in vivo is fivefold more active than 1,25-(OH)2D3 in a reporter gene expression system. Thus a metabolite may explain the long-lasting effects of the parent analog. In a second example, Dilworth et al. (87) recently showed that KH1060 is metabolized very rapidly in vitro and in vivo into as many as 22 metabolites but that some of these metabolites retain significant biological activity in similar reporter gene assay systems. The most abundant of these metabolites include 26-hydroxy and 24α-hydroxy KH1060, which can be detected in the blood in vivo, are stable in vitro, and as stated earlier can form trypsin-resistant complexes with VDR (87 353). At the very least, these active metabolites of KH1060 may explain its high potency when compared with 1,25-(OH)2D3 using in vitro assays. In addition, the findings may also explain the disappointing performance of KH1060 in vivo, since the metabolites are not allowed to accumulate and are rapidly cleared from the bloodstream. The third example of possible activation of an analog is 3-epimerization (275). Using in vitro cultured cell and perfused kidney systems where artificial media and high analog concentrations are employed, several researchers have observed the 3-epimerization of vitamin D analogs [e.g., 1,25-(OH)2D3 itself] (214 224 275). As with the example of KH1060, the claim is not that the metabolic product 3-epi-1,25-(OH)2D3 is more active than the parent compound, just that it retains significant biological activity and is more stable, not being subject to the C-24 oxidation pathway which is the fate that befalls the hormone. The nature of the enzyme involved is currently undefined but is reminiscent of the behavior of steroidal 3α- and 3β-oxidoreductases. Moreover, the formation of 1α- and 1β-hydroxylated DHT in rats and humans in vivo, referred to in section v A, resembles the 3-epimerization of vitamin D analogs and suggests that the process may be more than just an in vitro artifact. Nevertheless, the main hurdle for proponents of the 3-epimerization hypothesis is that they must show the metabolic step can occur in vivo.

The final metabolic consideration is hepatic modification. The liver is believed to be devoid of VDR and the inducible cytochrome P-450; CYP24 and cultured liver cells (Hep G2 and Hep 3B) do not appear to degrade the hormone 1,25-(OH)2D3 very rapidly (213 215). However, the story with vitamin D analogs appears to be quite different. Analogs such as calcipotriol and OCT are broken down by liver enzymes, these presumably contributing to the rapid metabolic clearance observed in vivo (213 215). As the analog structure is modified away from the vitamin D prototype, it becomes vulnerable to other more general catabolic enzymes found in the liver. One interesting example is the enzyme that saturates the double bond between C-22 and C-23 in the side chain of calcipotriol but does nothing to the same bond in the side chain of vitamin D2 (215). The simple replacement of the 24-methyl of the D2 side chain with a 24-hydroxyl group in calcipotriol is enough to allow the saturase enzyme to function. As vitamin D analogs continue to deviate more from the classical structure, they become more and more susceptible to general purpose or unrelated enzyme systems.

It should be noted with regard to molecular mechanisms of action at the target cell level that metabolism is often disregarded or given too little emphasis. Metabolic assumptions are made when testing biological activity in vitro that are not always valid. These include 1) the analog is biologically active as administered and 2) the analog has the same stability as 1,25-(OH)2D3 in the in vitro target cell model used, whether in in vitro organ culture, cultured target cell, or host cell/reporter gene construct. The validity of this approach is made even more tenuous when data acquired with different in vitro models, where metabolic considerations may or may not apply, are compared with data acquired in vivo, where metabolic considerations do apply. The reader is cautioned that invalid comparisons of in vivo and in vitro data abound in this field. The outcome is that a number of promising in vitro-tested vitamin D analogs have been found to be inactive when tested in vivo.

5. Relative importance of key factors on vitamin D analog action

It is evident that vitamin D analogs differ from 1,25-(OH)2D3 in a variety of ways: activating enzymes, DBP affinity, VDR-RXR-VDRE-gene association, and catabolic enzyme susceptibility. Any one or a combination of these differences could be important in making a specific analog appear calcemic or noncalcemic or work in vitro but not in vivo, or vice versa. Although it is attractive at this time to consider as most important the subtle VDR conformational differences induced by analogs as compared with 1,25-(OH)2D3 , there is a wealth of information that argues that metabolism, transport, and pharmacokinetics play a major role in vivo.

With our existing in vitro cell systems, it is difficult to study any one of these factors to the total exclusion of all others. Thus it seems unlikely that we shall be able to determine the relative importance of all of the factors listed until we have a fully functional cell-free in vitro model for transcription so that analogs can be tested without the complication of metabolic enzymes.

C. Future Directions in Vitamin D Drug Design

The growing availability of recombinant cytochromes P-450 will allow for a search for potential inhibitors. Such specific inhibitors of CYP24 may be of value in blocking 1,25-(OH)2D3 catabolism. Modeling of the vitamin D binding pocket of VDR, DBP, and the three vitamin D-related cytochrome P-450 will become a major goal now that all these specific proteins have been cloned and overexpressed. Although the full-length proteins are slightly beyond the current limits of NMR or X-ray crystallography, the ligand-binding domains or pockets are not. It is likely that technical problems with these procedures will be overcome shortly and the full-length proteins can be tackled. The membrane-associated region of cytochromes P-450 poses problems, but enormous strides have been made based on models built with soluble prokaryotic isoforms. Although the first priority is to study 1,25-(OH)2D3 binding to the VDR, the binding of a variety of analogs to all these proteins is likely to follow. In the process, we will be able to explore the permitted boundaries for optimal protein binding. More rational vitamin D analog design will follow to take advantage of structural idiosyncrasies of all of these key proteins. Meanwhile, the not-so-rational synthesis of new analogs is likely to continue.

We thank Pat Mings, Department of Biochemistry, University of Wisconsin-Madison, for expert technical help with the assembly of this document; David Prosser, Department of Biochemistry, Queen's University, who provided the amino acid alignment contained in Fig. 4 and helped with some artwork; and Martin Calverley for giving permission for reproduction of the X-ray structure of a double side-chain vitamin D analog.

FOOTNOTES

  • We were supported by National Institutes of Health Program Project Grant DK-14881 (to H. F. DeLuca), by a fund from the Wisconsin Alumni Research Foundation (to H. F. DeLuca), and by a grant from the Medical Research Council of Canada (to G. Jones).

  • 1The term 27-hydroxylation has been suggested by a consortium of cytochrome P-450 specialists to describe terminal hydroxylation of steroids given that methyl groups at C-26 and C-27 are indistinguishable. The old nomenclature for this was 26-hydroxylation, and the literature contains numerous references to 26-hydroxylation of vitamin D compounds giving rise to 26-hydroxylated vitamin D metabolites, e.g., 25,26-(OH)2D3 .

REFERENCES

  • 1 ABE E., MIURA C., SAKAGAMI H., TAKEDA M., KONNO K., YAMAZAKI T., YOSHIKI S., SUDA T.Differentiation of mouse myeloid leukemia cells induced by 1α,25-dihydroxyvitamin D3 .Proc. Natl. Acad. Sci. USA78198149904994
    Crossref | PubMed | ISI | Google Scholar
  • 2 ABOU-SAMRA A. B., JUPPNER H., KONG X. F., SCHIPANI E., IIDA-KLEIN A., KARGA H., URENA P., GARDELLA T. F., POTTS J. T.Jr., and H. M. KRONENBERG. Structure, function, and expression of the receptor for parathyroid hormone and parathyroid hormone-related peptide.Adv. Nephrol.231994247264
    PubMed | Google Scholar
  • 3 ADAMS, J. S. Extrarenal production and action of active vitamin D metabolites in human lymphoproliferative disases. In: Vitamin D, edited by D. Feldman, F. H. Glorieux, and J. W. Pike. San Diego, CA: Academic, 1997, p. 903–921.
    Google Scholar
  • 4 ADAMS J. S., GACAD M. A.Characterization of 1α-hydroxylation of vitamin D3 sterols by cultured alveolar macrophages from patients with sarcoidosis.J. Exp. Med.1611985755765
    Crossref | PubMed | ISI | Google Scholar
  • 5 AKIYOSHI-SHIBATA M., SAKAKI T., OHYAMA Y., NOSHIRO M., OKUDA K., YABUSAKI Y.Further oxidation of hydroxycalcidiol by calcidiol 24-hydroxylase: a study with the mature enzyme expressed in Escherichia coli.Eur. J. Biochem.2241994335343
    Crossref | PubMed | Google Scholar
  • 6 ALROY I., TOWERS T. L., FREEDMAN L. P.Transcriptional repression of the interleukin-2 gene by vitamin D3: direct inhibition of NFATp/AP-1 complex formation by a nuclear hormone receptor.Mol. Cell. Biol.15199457895799
    Crossref | ISI | Google Scholar
  • 7 ANDERSSON S., DAVIS D. L., DAHLBACK H., JORNVALL H.Cloning, structure and expression of the mitochondrial cytochrome P-450 sterol 26-hydroxylase, a bile acid biosynthetic enzyme.J. Biol. Chem.264198982228229
    PubMed | ISI | Google Scholar
  • 8 ANDERSSON S., HOLMBERG I., WIKVALL K.25-Hydroxylation of C27-steroids and vitamin D3 in rat liver microsomes.J. Biol. Chem.258198367776781
    PubMed | ISI | Google Scholar
  • 9 AXEN E., BERGMAN T., WIKVALL K.Microsomal 25-hydroxylation of vitamin D2 and vitamin D3 in pig liver.J. Steroid Mol. Biol.51199497106
    Crossref | PubMed | ISI | Google Scholar
  • 10 AXEN E., POSTLIND H., WIKVALL K.Effects on CYP expression in rat kidney and liver by 1α,25-dihydroxyvitamin D3 , 1α-hydroxylase activity.Biochem. Biophys. Res. Commun.2151995136141
    Crossref | PubMed | ISI | Google Scholar
  • 11 BAGGIOLINI E. G., PARTRIDGE J. J., SHIVEY S. J., TRUITT G. A., VSKOKOVIC M. R.Cholecalciferol 23-yne derivatives, their pharmaceutical compositions, their use in the treatment of calcium-related diseases, and their antitumor activity, US 4,804,502.Chem. Abstr.111198958160d
    Google Scholar
  • 12 BAKER A. R., McDONNELL D. P., HUGHES M., CRISP T. M., MANGELSDORF D. J., HAUSSLER M. R., PIKE J. W., SHINE J., O'MALLEY B. W.Cloning and expression of full-length cDNA encoding human vitamin D receptor.Proc. Natl. Acad. Sci. USA85198832943298
    Crossref | PubMed | ISI | Google Scholar
  • 13 BARTON D. H., HESSE R. H., PECHET M. M., RIZZARDO E.A convenient synthesis of 1-hydroxy-vitamin D3 .J. Am. Chem. Soc.95197327482749
    Crossref | PubMed | ISI | Google Scholar
  • 14 BAUER G. C. H., CARLSSON A., LINDQUIST B.Evaluation of accretion, resorption, and exchange reactions in the skeleton.Kungl. Fysiograf. Sallskapets I. Lund Forhandlingar251955318
    Google Scholar
  • 15 BECKMAN M., TADIKONDA P., WERNER E., PRAHL J. M., YAMADA S., DeLUCA H. F.Human 25-hydroxyvitamin D3-24-hydroxylase, a multicatalytic enzyme.Biochemistry35199684658472
    Crossref | PubMed | ISI | Google Scholar
  • 16 BECKMAN M. J., GOFF J. P., REINHARDT T. A., BEITZ D. C., HORST R. L.In vivo regulation of rat intestinal 24-hydroxylase: potential new role of calcitonin.Endocrinology1351994119511955
    Crossref | ISI | Google Scholar
  • 17 BELSEY R. E., DeLUCA H. F., POTTS J. T.Selective binding properties of vitamin D transport proteins in chick plasma in vitro.Nature2471974208209
    Crossref | PubMed | ISI | Google Scholar
  • 18 BERGINER V. M., SHANY S., ALKALAY D., BERGINER J., DEKEL S., SALEN G., TINT G. S., GAZIT D.Osteoporosis and increased bone fractures in cerebrotendinous xanthomatosis.Metabolism4219936974
    Crossref | PubMed | ISI | Google Scholar
  • 19 BHALLA A. K., AMENTO E. P., CLEMENS T. L., HOLICK M. F., KRANE S. M.Specific high-affinity receptors for 1,25-dihydroxyvitamin D3 in human peripheral blood mononuclear cells. Presence in monocytes and induction in T lymphocytes following activation.J. Clin. Endocrinol.57198313081311
    Crossref | PubMed | ISI | Google Scholar
  • 20 BHATTACHARYYA M. H., DeLUCA H. F.The regulation of the rat liver calciferol-25-hydroxylase.J. Biol. Chem.248197329692973
    PubMed | ISI | Google Scholar
  • 21 BIKLE, D. D. Vitamin D and skin. In: Vitamin D, edited by D. Feldman, F. H. Glorieux, and J. W. Pike. San Diego, CA: Academic, 1997, chapt. 25, p. 379–394.
    Google Scholar
  • 22 BIKLE D. D., NEMANIC M. K., GEE E. A., ELIAS P.1,25-Dihydroxyvitamin D3 production by human keratinocytes: kinetics and regulation.J. Clin. Invest.781986557566
    Crossref | PubMed | ISI | Google Scholar
  • 23 BINDERUP, E., M. J. CALVERLY, and L. BINDERUP. Synthesis and biological activity of 1-hydroxylated vitamin D analogues with poly-unsaturated side chains. In: Vitamin D: Gene Regulation, Structure-function Analysis and Clinical Application, edited by A. W. Norman, R. Bouillon, and M. Thomasset. Berlin: de Gruyter, 1991, p. 192–193.
    Google Scholar
  • 24 BINDERUP L., BRAMM E.Effects of a novel vitamin D analog MC903 on cell proliferation and differentiation in vitro and on calcium metabolism in vivo.Biochem. Pharmacol.371988889895
    Crossref | PubMed | ISI | Google Scholar
  • 25 BINDERUP L., LATINI S., BINDERUP E., BRETTING C., CALVERLEY M. J., HANSEN K.20-Epi-vitamin D3 analogues: a novel class of potent regulators of cell growth and immune responses.Biochem. Pharmacol.42199115691575
    Crossref | PubMed | ISI | Google Scholar
  • 26 BISHOP J. E., COLLINS E. D., OKAMURA W. H., NORMAN A. W.Profile of ligand specificity of the vitamin D binding protein for 1,25-dihydroxyvitamin D3 and its analogs.J. Bone Miner. Res.9199412771288
    Crossref | PubMed | ISI | Google Scholar
  • 27 BLANCO J. C. G., WANG I. M., TSAI S. Y., TSIA M. J., O'MALLEY B. W., JURUTKA P. W., HAUSSLER M. R., OZATO K.Transcription factor TFIIB and the vitamin D receptor cooperatively activate ligand-dependent transcription.Proc. Natl. Acad. Sci. USA92199515351539
    Crossref | PubMed | ISI | Google Scholar
  • 28 BLUNT J. W., DeLUCA H. F., SCHNOES H. K.25-Hydroxycholecalciferol. A biologically active metabolite of vitamin D3 .Biochemistry7196833173322
    Crossref | PubMed | ISI | Google Scholar
  • 29 BOUILLON, R. Biological activities of CD-ring modified non-steroidal 1,25(OH)2 vitamin D analogs: E-ring analogs. In: Vitamin D. Chemistry, Biology and Clinical Applications of the Steroid Hormone, edited by A. W. Norman, R. Bouillon, and M. Thomasset. Riverside: Univ. of California, 1997, p. 59–64.
    Google Scholar
  • 30 BOUILLON R., ALLEWAERT K., XIANG D. Z., TAN B. K., VAN BAELEN H.Vitamin D analogs with low affinity for the vitamin D binding protein: enhanced in vitro and decreased in vivo activity.J. Bone Miner. Res.6199110511057
    Crossref | PubMed | ISI | Google Scholar
  • 31 BOUILLON R., OKAMURA W. H., NORMAN A. W.Structure-function relationships in the vitamin D endocrine system.Endocr. Rev.161995200257
    PubMed | ISI | Google Scholar
  • 32 BOURGUET W., RUFF M., CHAMBON P., GRONEMEYER H., MORAS D.Crystal structure of the ligand-binding domain of the human nuclear receptor RXR-α.Nature3751995377382
    Crossref | PubMed | ISI | Google Scholar
  • 33 BOYAN, B. D., D. D. DEAN, V. L. SYLVIA, and Z. SCHWARTZ. Cartilage and vitamin D: genomic and nongenomic regulation by 1,25(OH)2D3 and 24,25(OH)2D3 . In: Vitamin D, edited by D. Feldman, F. H. Glorieux, and J. W. Pike. San Diego, CA: Academic, 1997, chapt. 26, p. 395–421.
    Google Scholar
  • 34 BOYLE, I. T., R. W. GRAY, J. L. OMDAHL, and H. F. DeLUCA. Calcium control of an in vivo biosynthesis of 1,25-dihydroxyvitamin D3: Nicolaysen's endogenous factor. In: Endocrinology 1971, edited by S. Taylor. London: Heinemann, 1972, p. 468–476.
    Google Scholar
  • 35 BOYLE I. T., MIRAVET L., GRAY R. W., HOLICK M. F., DeLUCA H. F.The response of intestinal calcium transport to 25-hydroxyvitamin D in nephrectomized rats.Endocrinology901972605608
    Crossref | PubMed | ISI | Google Scholar
  • 36 BRANISTEANU D. D., WAER M., SOBIS H., MARCELIS S., VANDEPUTTE M., BOUILLON R.Prevention of murine experimental allergic encephalomyelitis: cooperative effects of cyclosporine and 1α,25(OH)2D3 .J. Neuroimmunol.611995151160
    Crossref | PubMed | ISI | Google Scholar
  • 37 BROMMAGE R., DeLUCA H. F.A maternal defect is responsible for the growth failure in vitamin D-deficient rat pups.Am. J. Physiol.256(Endocrinol. Metab. 19)1984E216E220
    Google Scholar
  • 38 BROMMAGE R., DeLUCA H. F.Evidence that 1,25-dihydroxyvitamin D3 is the physiologically active metabolite of vitamin D3 .Endocr. Rev.61985491511
    Crossref | PubMed | ISI | Google Scholar
  • 39 BROWN E. M., GAMBA G., RICCARDI R., LOMBARDI M., BUTTERS R., KIFOR O., SUN A., HEDIGER M. A., LYTTON J., HEBERT J.Cloning and characterization of an extracellular Ca2+-sensing receptor from bovine parathyroid.Nature3661993575580
    Crossref | PubMed | ISI | Google Scholar
  • 40 BROWN G., BUNCE C. M., ROWLANDS D. C., WILLIAMS G. R.All-trans retinoic acid and 1,25-dihydroxyvitamin D3 co-operate to promote differentiation of the human promyeloid leukemia cell line HL60 to monocytes.Leukemia81994806815
    PubMed | ISI | Google Scholar
  • 41 BROWN T. A., DeLUCA H. F.Phosphorylation of the 1,25-dihydroxyvitamin D3 receptor. A primary event in 1,25-dihydroxyvitamin D3 action.J. Biol. Chem.26519901002510029
    PubMed | ISI | Google Scholar
  • 42 BURMESTER J., WIESE R., MAEDA N., DeLUCA H. F.Structure and regulation of the rat 1,25-dihydroxyvitamin D3 receptor.Proc. Natl. Acad. Sci. USA85198894999502
    Crossref | PubMed | ISI | Google Scholar
  • 43 BURNS K., DUGGAN B., ATKINSON E. A., FAMULSKI K. S., NEMER M., BLEACKLEY R. C., MICHALAK M.Modulation of gene expression by calreticulin binding to the glucocorticoid receptor.Nature3671994476480
    Crossref | PubMed | ISI | Google Scholar
  • 44 CALI J. J., RUSSELL D. W.Characterization of human sterol 27-hydroxylase. A mitochondrial cytochrome P-450 that catalyzes multiple oxidation reactions in bile acid biosynthesis.J. Biol. Chem.266199177747778
    PubMed | ISI | Google Scholar
  • 45 CALVERLEY M. J.Synthesis of MC-903, a biologically active vitamin D metabolite analog.Tetrahedron43198746094619
    Crossref | ISI | Google Scholar
  • 46 CALVERLEY, M. J., E. BINDERUP, and L. BINDERUP. The 20-epi modification in the vitamin D series: selective enhancement of “non-classical” receptor-mediated effects. In: Vitamin D: Gene Regulation, Structure-Function Analysis and Clinical Application, edited by A. W. Norman, R. Bouillon, and M. Thomasset. Berlin: de Gruyter, 1991, p. 163–164.
    Google Scholar
  • 47 CALVERLEY, M. J., and G. JONES. Vitamin D. In: Antitumor Steroids, edited by B. R. T. Toronto. Orlando, FL: Academic, 1992, p. 193–270.
    Google Scholar
  • 48 CANDELIERE G. A., JURUTKA P. W., HAUSSLER M. R., ST-ARNAUD R.A composite element binding the vitamin D receptor, retinoid X receptor α, and a member of CTF/NF-1 family of transcription factors mediates the vitamin D responsiveness of the c-fos promoter.Mol. Cell. Biol.16199658492
    Crossref | PubMed | ISI | Google Scholar
  • 49 CANTORNA M. T., HAYES C. E., DeLUCA H. F.1,25-Dihydroxyvitamin D3 reversibly blocks the progression of relapsing encephalomyelitis.Proc. Natl. Acad. Sci. USA93199678617864
    Crossref | PubMed | ISI | Google Scholar
  • 50 CANTORNA M. T., HAYES C. E., DeLUCA H. F.1,25-Dihydroxyvitamin D3 prevents and ameliorates symptoms in two experimental models of human arthritis.J. Nutr.12819986872
    Crossref | PubMed | ISI | Google Scholar
  • 51 CANTORNA M. T., WOODWARD W. D., HAYES C. E., DeLUCA H. F.1,25-Dihydroxyvitamin D3 is a positive regulator for the two anti-encephalitogenic cytokines TGF-β1 and IL-4.J. Immunol.160199853145319
    PubMed | ISI | Google Scholar
  • 52 CAO X., ROSS F. P., ZHANG L., MacDONALD P. N., CHAPPEL J., TEITELBAUM S. L.Cloning of the promoter for the avian integrin β3 subunit gene and its regulation by 1,25-dihydroxyvitamin D3 .J. Biol. Chem.26819932737127380
    PubMed | ISI | Google Scholar
  • 53 CARLBERG C.Mechanisms of nuclear signalling by vitamin D3: interplay with retinoid and thyroid hormone signalling.Eur. J. Biochem.2311995517527
    Crossref | PubMed | Google Scholar
  • 54 CARLBERG C., BENDIK I., WYSS A., MEIER E., STURZENBECKER L. J., GRIPPO J. F., HUNZIKER W.Two nuclear signalling pathways for vitamin D.Nature3611993657660
    Crossref | PubMed | ISI | Google Scholar
  • 55 CARLBERG C., MATHIASEN I. S., SAURAT J. H., BINDERUP L.The 1,25-dihydroxyvitamin D3 (VD) analogues MC903, EB1089, and KH1060 activate the VD receptor: homodimers show higher ligand sensitivity than heterodimers with retinoid X receptors.J. Steroid Biochem. Mol. Biol.511994137142
    Crossref | PubMed | ISI | Google Scholar
  • 56 CASTILLO L., TANAKA Y., DeLUCA H. F., IKEKAWA N.On the physiological role of 1,24,25-trihydroxyvitamin D3 .Miner. Electrolye Metab.11978198207
    Google Scholar
  • 57 CHAMBERS T. J., MAGNUS C. J.Calcitonin alters behaviour of isolated osteoclasts.J. Pathol.13619822739
    Crossref | PubMed | ISI | Google Scholar
  • 58 CHAMBON P.A decade of molecular biology of retinoic acid receptors.FASEB J.101996940954
    Crossref | PubMed | ISI | Google Scholar
  • 59 CHEN H., LIN R. J., SCHLITZ R. L., CHAKRAVART D., NASH A., NAGY L., PRIVALSKY M. L., NAKATANI Y., EVANS R. M.Nuclear receptor coactivator ACTR is a novel histone acetyltransferase and forms a multimeric activation complex with P/CAF and CBP/p300.Cell901997569580
    Crossref | PubMed | ISI | Google Scholar
  • 60 CHEN J. D., EVANS R. M.A transcriptional co-repressor that interacts with nuclear hormone receptors.Nature3771995454457
    Crossref | PubMed | ISI | Google Scholar
  • 61 CHEN K. S., DeLUCA H. F.Cloning of the human 1α,25-dihydroxyvitamin D3 24-hydroxylase gene and identification of two vitamin D-responsive elements.Biochim. Biophys. Acta1263199519
    Crossref | PubMed | ISI | Google Scholar
  • 62 CHEN K.-S., PRAHL J. M., DeLUCA H. F.Isolation and expression of the human 1,25-dihydroxyvitamin D3 24-hydroxylase cDNA.Proc. Natl. Acad. Sci. USA90199345434547
    Crossref | PubMed | ISI | Google Scholar
  • 63 CHEN T. C., CASTILLO L., KORYCKA-DAHL M., DeLUCA H. F.Role of vitamin D metabolites in phosphate transport of rat intestine.J. Nutr.104197410561060
    Crossref | PubMed | ISI | Google Scholar
  • 64 CHENG L., NORRIS A. W., TATE B. F., ROSENBERGER M., GRIPPO J. F., LI E.Characterization of the ligand binding domain of human retinoid X receptor α expressed in Escherichia coli.J. Biol. Chem.26919941866218667
    PubMed | ISI | Google Scholar
  • 65 CHERTOW B. S., SIVITZ W. I., BARANETSKY N. G., CLARK S. A., WAITE A., DeLUCA H. F.Cellular mechanisms of insulin release. The effects of vitamin D deficiency and repletion of the rat insulin secretion.Endocrinology113198315111518
    Crossref | PubMed | ISI | Google Scholar
  • 66 CHERTOW B. S., SIVITZ W. I., BARANETSKY N. G., CORDLE M. B., DeLUCA H. F.Islet insulin release and net calcium retention in vitro in vitamin D-deficient rats.Diabetes351986771775
    Crossref | PubMed | ISI | Google Scholar
  • 67 CHESKIS B., LEMON B. D., USKOKOVIC M. R., LOMEDICO P. T., FREEDMAN L. P.Vitamin D3-retinoid X receptor dimerization, DNA binding, and transactivation are differentially affected by analogs of 1,25-dihydroxyvitamin D3 .Mol. Endocrinol.9199518141824
    PubMed | Google Scholar
  • 68 CHRISTAKOS, S., J. D. BECK, and S. J. HYLLNER. Calbindin-D28K . In: Vitamin D, edited by D. Feldman, F. H. Glorieux, and J. W. Pike. San Diego, CA: Academic, 1997, chapt. 13, p. 209–221.
    Google Scholar
  • 69 CHRISTAKOS S., GILL R., LEE S., LI H.Molecular aspects of the calbindins.J. Nutr.1221992678682
    Crossref | PubMed | ISI | Google Scholar
  • 70 COOKE, N. C., and J. G. HADDAD. Vitamin D binding protein. In: Vitamin D, edited by D. Feldman, F. H. Glorieux, and J. W. Pike. San Diego, CA: Academic, 1997, p. 87–101.
    Google Scholar
  • 71 COTT D. K., HALL R. K., GRANNER D. K.Retinoid receptors cause distortion of the retinoic acid response element in the phosphoenolpyruvate carboxykinase gene promoter.Biochem. J.3101995483490
    Crossref | PubMed | ISI | Google Scholar
  • 72 DAME M. C., PIERCE E. A., DeLUCA H. F.Identification of the porcine intestinal 1,25-dihydroxyvitamin D3 receptor on sodium dodecyl sulfate/polyacrylamide gels by renaturation and immunoblotting.Proc. Natl. Acad. Sci. USA82198578257829
    Crossref | PubMed | ISI | Google Scholar
  • 73 DARWISH H. M., BURMEISTER J. K., MOSS V. E., DeLUCA H. F.Phosphorylation is involved in transcriptional activation by the 1,25-dihydroxyvitamin D3 receptor.Biochim. Biophys. Acta116719932936
    Crossref | PubMed | ISI | Google Scholar
  • 74 DARWISH H. M., DeLUCA H. F.Identification of a 1,25-dihydroxyvitamin D3-response element in the 5′-flanking region of the rat calbindin D9k gene.Proc. Natl. Acad. Sci. USA891992603607
    Crossref | PubMed | ISI | Google Scholar
  • 75 DARWISH H. M., DeLUCA H. F.Recent advances in the molecular biology of vitamin D action.Prog. Nucleic Acid Res. Mol. Biol.531996321344
    Crossref | PubMed | Google Scholar
  • 76 DARWISH H. M., DeLUCA H. F.Analysis of binding of the 1,25-dihydroxyvitamin D3 receptor to positive and negative vitamin D response elements.Arch. Biochem. Biophys.3341996223234
    Crossref | PubMed | ISI | Google Scholar
  • 77 DARWISH H. M., KRISINGER J., STROM M., DeLUCA H. F.Molecular cloning of the cDNA and chromosomal gene for vitamin D-dependent calcium-binding protein of rat intestine.Proc. Natl. Acad. Sci. USA84198761086111
    Crossref | PubMed | ISI | Google Scholar
  • 78 DE CLERQ, P., C. D'HALLEWEYN, and S. GABRIELS. Development of CD-ring modified non-steroidal 1,25-(OH)2 vitamin D analogs. In: Vitamin D. Chemistry, Biology and Clinical Applications of the Steroid Hormone, edited by A. W. Norman, R. Bouillon, and M. Thomasset. Riverside: Univ. of California, 1997, p. 3–10.
    Google Scholar
  • 79 DeLUCA, H. F. Active compounds. In: The Vitamins. Chemistry, Physiology, Pathology, Methods, edited by W. H. Sebrell, Jr., and R. S. Harris. New York: Academic, 1971, chapt. IV, p. 223–232.
    Google Scholar
  • 80 DeLUCA, H. F. Vitamin D: metabolism and function. In: Monographs on Endocrinology, edited by F. Gross, M. M. Grumbach, A. Labhart, M. B. Lipsett, T. Mann, L. T. Samuels, and J. Zander. New York: Springer-Verlag, 1979, p. 1–78.
    Google Scholar
  • 81 DeLUCA, H. F. The transformation of a vitamin into a hormone: the vitamin D story. In: The Harvey Lectures. New York: Academic, 1981, ser. 75, p. 333–379.
    Google Scholar
  • 82 DeLUCA H. F., SCHNOES H. K.Metabolism in mechanism of action of vitamin D3 .Annu. Rev. Biochem.451976631666
    Crossref | PubMed | ISI | Google Scholar
  • 83 DELVIN E. E., ARABIAN A.Kinetics and regulation of 25-hydroxycholecalciferol 1α-hydroxylase from cells isolated from human term decidua.Eur. J. Biochem.1641987659662
    Crossref | Google Scholar
  • 84 DEMAY M. B., GERARDI J. M., DeLUCA H. F., KRONENBERG H. M.DNA sequences in the rat osteocalcin gene that bind the 1,25-dihydroxyvitamin D3 receptor and confer responsiveness to 1,25-dihydroxybitamin D3 .Proc. Natl. Acad. Sci. USA871990369373
    Crossref | PubMed | ISI | Google Scholar
  • 85 DEMAY M. B., KIERNAN M. S., DeLUCA H. F., KRONENBERG H. M.Sequences in the human parathyroid hormone gene that bind to the 1,25-dihydroxyvitamin D3 receptor and mediate transcriptional repression in response to 1,25-dihydroxyvitamin D3 .Proc. Natl. Acad. Sci. USA89199280978101
    Crossref | PubMed | ISI | Google Scholar
  • 86 DILWORTH F. J., CALVERLEY M. J., MAKIN H. L. J., JONES G.Increased biological activity of 20-epi-1,25-dihydroxyvitamin D3 is due to reduced catabolism and altered protein binding.Biochem. Pharmacol.471994987993
    Crossref | PubMed | ISI | Google Scholar
  • 87 DILWORTH F. J., WILLIAMS G. R., KISSMEYER A.-M., LOGSTED-MIELSEN J., BINDERUP E., CALVERLEY M. J., MAKIN H. L. J., JONES G.The vitamin D analog KH1060 is rapidly degraded both in vivo and in vitro via several pathways: principal metabolites generated retain significant biological activity.Endocrinology138199754855496
    Crossref | PubMed | ISI | Google Scholar
  • 88 DOKOH S., DONALDSON C. A., MARION S. L., PIKE J. W., HAUSSLER M. R.The ovary: a target organ for 1,25-dihydroxyvitamin D3 .Endocrinology1121983200206
    Crossref | PubMed | ISI | Google Scholar
  • 89 DUSSO A. S., NEGREA L., GUNAWARDHANA S., LOPEZ-HILKER S., FINCH J., MORI T., NISHII Y., SLATOPOLSKY E., BROWN A. J.On the mechanisms for the selective action of vitamin D analogs.Endocrinology128199116871692
    Crossref | PubMed | ISI | Google Scholar
  • 90 DYSON E., SUCOV H. M., KUBALAK S. W., SCHMID-SCHONBEIN G. W., DELANL F. A., EVANS R. M., ROSS J. J., CHIEN K. R.Atrial-like phenotype is associated with embryonic ventricular failure in retinoid X receptor α −/− mice.Proc. Natl. Acad. Sci. USA92199573867390
    Crossref | PubMed | ISI | Google Scholar
  • 91 EISMAN J. A., KOGA M., SUTHERLAND R. L., BARKLA D. H., TUTTON P. J. M.1,25-Dihydroxyvitamin D3 and the regulation of human cancer cell replication.Proc. Soc. Exp. Biol. Med.1911989221226
    Crossref | PubMed | ISI | Google Scholar
  • 92 ELAROUSSI M. A., PRAHL J. M., DeLUCA H. F.The avian vitamin D receptors: primary structures and their origins.Proc. Natl. Acad. Sci. USA9119941159611600
    Crossref | PubMed | ISI | Google Scholar
  • 93 ERBEN R. G., BANTE U., BIRNER H., STANGASSINGER M.1-Hydroxyvitamin D2 partially dissociates between preservation of cancellous bone mass and effects on calcium homeostasis in ovarectomized rats.Calcif. Tissue Int.601997449456
    Crossref | PubMed | ISI | Google Scholar
  • 94 ESVELT R. P., SCHNOES H. K., DeLUCA H. F.Isolation and characterization of 1α-hydroxy-23-carboxytetranorvitamin D3: a major metabolite of 1,25-dihydroxyvitamin D3 .Biochemistry18197939773983
    Crossref | PubMed | ISI | Google Scholar
  • 95 EVANS R. M.The steroid and thyroid hormone receptor superfamily.Science2401988889895
    Crossref | PubMed | ISI | Google Scholar
  • 96 FELDMAN, D., F. H. GLORIEUX, and W. PIKE. Vitamin D. San Diego, CA: Academic, 1997, p. 1–1285.
    Google Scholar
  • 97 FIESER, L. F., and M. FIESER. Steroids. New York: Reinhold, 1959, p. 144–146.
    Google Scholar
  • 98 FORMAN B. M., UMESONO K., CHEN H., EVANS R. M.Unique response pathways are established by allosteric interactions among nuclear hormone receptors.Cell811995541550
    Crossref | PubMed | ISI | Google Scholar
  • 99 FRASER D., KOOH S. W., KIND P., HOLICK M. F., TANAKA Y., DeLUCA H. F.Pathogenesis of hereditary vitamin D dependency rickets.N. Engl. J. Med.2891973817822
    Crossref | PubMed | ISI | Google Scholar
  • 100 FRASER D. R., KODICEK E.Unique biosynthesis by kidney of a biologically active vitamin D metabolite.Nature2281970764766
    Crossref | PubMed | ISI | Google Scholar
  • 101 FRIEDLANDER G., AMIEL C.Cellular mode of action of parathyroid hormone.Adv. Nephrol.231994265279
    PubMed | Google Scholar
  • 102 FU G. K., LIN D., ZHANG M. Y. H., BIKLE D. D., SHACKLETON C. H. L., MILLER W. L., PORTALE A. A.Cloning of human 25-hydroxyvitamin D-1α-hydroxylase and mutations causing vitamin D-dependent rickets type 1.Mol. Endocrinol.11199719611970
    PubMed | Google Scholar
  • 103 FU G. K., PORTALE A. A., MILLER W. L.Complete structure of the human gene for the vitamin D 1α-hydroxylase, P450c1α.DNA Cell Biol.16199714991507
    Crossref | PubMed | ISI | Google Scholar
  • 104 FUKUSHIMA M., SUZUKI Y., TOHIRA Y., NISHII Y., SUZUKI M., SASAKI S., SUDA T.25-Hydroxylation of 1α-hydroxyvitamin D3 in vivo and in the perfused rat liver.FEBS Lett.651976211214
    Crossref | PubMed | ISI | Google Scholar
  • 105 GALANTE L., COLSTON K. W., MacAULEY S. J., MacINTYRE I.Effect of calcitonin on vitamin D metabolism.Nature2381972271273
    Crossref | PubMed | ISI | Google Scholar
  • 106 GALLAGHER J. C., BISHOP C. W., KNUTSON J. C., MAZESS R. B., DeLUCA H. F.Effects of increasing doses of 1 alpha-hydroxyvitamin D2 on calcium homeostasis in postmenopausal osteopenic women.J. Bone Miner. Res.91994607614
    Crossref | PubMed | ISI | Google Scholar
  • 107 GALLAGHER J. C., RIGGS B. L.Action of 1,25-dihydroxyvitamin D3 on calcium balance and bone turnover and its effect on vertebral fracture rate.Metabolism3919903034
    Crossref | PubMed | ISI | Google Scholar
  • 108 GALLAGHER J. C., RIGGS B. L., RECKER R. R., GOLDGAR D.The effect of calcitriol on patients with postmenopausal osteoporosis with special reference to fracture frequency.Proc. Soc. Exp. Biol. Med.1911989287292
    Crossref | PubMed | ISI | Google Scholar
  • 109 GARABEDIAN M., HOLICK M. F., DeLUCA H. F., BOYLE I. T.Control of 25-hydroxycholecalciferol metabolism by the parathyroid glands.Proc. Natl. Acad. Sci. USA69197216731676
    Crossref | PubMed | ISI | Google Scholar
  • 110 GARABEDIAN M., TANAKA Y., HOLICK M. F., DeLUCA H. F.Response of intestinal calcium transport and bone calcium mobilization to 1,25-dihydroxyvitamin D3 in parathyroidectomized rats.Endocrinology94197410221027
    Crossref | PubMed | ISI | Google Scholar
  • 111 GHAZARIAN, J. G. The renal mitochondrial hydroxylases of the vitamin D3 endocrine complex. How are they regulated at the molecular level? J. Bone Miner. Res. 5: 897–903, 1990.
    Google Scholar
  • 112 GHAZARIAN J. G., JEFCOATE C. R., KNUTSON J. C., ORME-JOHNSON W., DeLUCA H. F.Mitochondrial cytochrome P-450. A component of chick kidney 25-hydroxycholecalciferol-1α-hydroxylase.J. Biol. Chem.249197430263033
    PubMed | ISI | Google Scholar
  • 113 GILL R. K., CHRISTAKOS S.Identification of sequence elements in mouse calbindin-D28k gene that confer 1,25-dihydroxyvitamin D3- and butyrate-inducible responses.Proc. Natl. Acad. Sci. USA84199329842988
    Crossref | ISI | Google Scholar
  • 114 GOLDEN P., GREENWALT A., MARTIN K., BELLORIN-FONT E., MAZEY R., KLAHR S., SLATOPOLSKY E.Lack of direct effect of 1,25-dihydroxycholecalciferol on parathyroid hormone secretion by normal bovine parathyroid glands.Endocrinology1071980602607
    Crossref | PubMed | ISI | Google Scholar
  • 115 GOLDEN, P., R. MAZEY, A. GREENWALT, K. MARTIN, and E. SLATOPOLSKY. Vitamin D: a direct effect on the parathyroid gland? Miner. Electrolyte Metab. 2: 1–6, 1979.
    Google Scholar
  • 116 GRAY R. W., CALDAS A. E., WILZ D. R., LEMANN J.Jr., G. A. SMITH, and H. F. DeLUCA. Metabolism and excretion of 3H-1,25-(OH)2-vitamin D3 in healthy adults.J. Clin. Endocrinol. Metab.461978756765
    Crossref | PubMed | ISI | Google Scholar
  • 117 GRAY R. W., OMDAHL J. L., GHAZARIAN J. G., DeLUCA H. F.25-Hydroxycholecalciferol-1α-hydroxylase: subcellular distribution and properties.J. Biol. Chem.247197275287532
    PubMed | ISI | Google Scholar
  • 118 GUO Y.-D., JONES G.Gene promoter for human liver vitamin D3-25-hydroxylase (CYP27) lacks a classical VDRE but contains silencer sequence (Abstract).J. Bone Miner. Res.91994S289
    Google Scholar
  • 119 GUO Y.-D., STRUGNELL S., BACK D. W., JONES G.Transfected human liver cytochrome P-450 hydroxylates vitamin D analogs at different side-chain positions.Proc. Natl. Acad. Sci. USA90199386688672
    Crossref | PubMed | ISI | Google Scholar
  • 120 HA I., LANE W. S., REINBERG D.Cloning of a human gene encoding the general transcription initiation factor IIB.Nature3521991689695
    Crossref | PubMed | ISI | Google Scholar
  • 121 HALLORAN B. P., DeLUCA H. F.Vitamin D deficiency and reproduction in rats.Science20419797374
    Crossref | PubMed | ISI | Google Scholar
  • 122 HALLORAN B. P., DeLUCA H. F.Effect of vitamin D deficiency on fertility and reproductive capacity in the female rat.J. Nutr.110198015731580
    Crossref | PubMed | ISI | Google Scholar
  • 123 HALLORAN B. P., DeLUCA H. F.Intestinal calcium transport: evidence for two distinct mechanisms of action of 1,25-dihydroxyvitamin D3 .Arch. Biochem. Biophys.2081981477486
    Crossref | PubMed | ISI | Google Scholar
  • 124 HANSEN, K., M. J. CALVERLEY, and L. BINDERUP. Synthesis and biological activity of 22-oxa vitamin D analogues. In: Vitamin D: Gene Regulation, Structure-Function Analysis and Clinical Application, edited by A. W. Norman, R. Bouillon, and M. Thomasset. Berlin: de Gruyter, 1991, p. 161–162.
    Google Scholar
  • 125 HARMEYER J., DeLUCA H. F.Calcium-binding protein and calcium absorption after vitamin D administration.Biochem. J.170196993102
    Google Scholar
  • 126 HARRISON H. E., HARRISON H. C.Intestinal transport of phosphate: action of vitamin D, calcium, and potassium.Am. J. Physiol.201196110071012
    Abstract | ISI | Google Scholar
  • 127 HAUSSLER M. R.Vitamin D receptors: nature and function.Annu. Rev. Nutr.61986527562
    Crossref | PubMed | ISI | Google Scholar
  • 128 HAUSSLER M. R., JURUTKA P. W., HSIEH J. C., THOMPSON P. D., SELZNICK S. H., HAUSSLER C. A., WHITFIELD G. K.New understanding of the molecular mechanism of receptor-mediated genomic actions of the vitamin D hormone.Bone17, Suppl.199533S38S
    Crossref | ISI | Google Scholar
  • 129 HAYASHI S., USUI E., OKUDA K.Sex-related difference in vitamin D3 25-hydroxylase of rat liver mitochondria.J. Biochem.1031988863866
    Crossref | PubMed | ISI | Google Scholar
  • 130 HEANEY, R. P. Vitamin D: role in the calcium economy. In: Vitamin D, edited by D. Feldman, F. H. Glorieux, and J. W. Pike. San Diego, CA: Academic, 1997, chapt. 31, p. 485–497.
    Google Scholar
  • 131 HEINZEL T., LAVINSKY R. M., MULLEN T.-M., SODERSTROM M., LAHERTY C. D., TORCHIA J., YANK W.-M., BRARD G., NGO S. D., DAVIE J. R.A complex containing N-CoR, mSin3 and histone deacetylase mediates transcriptional repression.Nature38719974348
    Crossref | PubMed | ISI | Google Scholar
  • 132 HENRY, H. L. The 25-hydroxyvitamin D3 1α-hydroxylase. In: Vitamin D, edited by D. Feldman, F. H. Glorieux, and J. W. Pike. San Diego, CA: Academic, 1997, p. 57–68.
    Google Scholar
  • 133 HENRY H. L., NORMAN A. W.Studies on the mechanism of action of calciferol. VII. Localization of 1,25-dihydroxy-vitamin D3 in chick parathyroid glands.Biochem. Biophys. Res. Commun.621975781788
    Crossref | PubMed | ISI | Google Scholar
  • 134 HEYMAN R. A., MANGELSDORF D. J., DYCK J. A., STEIN R. B., EICHELE G., EVANS R. M., THALLER C.9-Cis retinoic acid is a high affinity ligand for the retinoid X receptor.Cell681992397406
    Crossref | PubMed | ISI | Google Scholar
  • 135 HILLIARD, G. M., IV, R. G. COOK, N. L. WEIGEL, and J. W. PIKE. 1,25-Dihydroxyvitamin D3 modulates phosphorylation of serine 205 in the human vitamin D receptor: site-directed mutagenesis of this residue promotes alternative phosphorylation. Biochemistry 33: 4300–4311, 1994.
    Google Scholar
  • 136 HOCK J. M., GUNNESS-HEY M., POSER J., OLSON H., BELL N. H., RAISZ L. G.Stimulation of undermineralized matrix formation by 1,25-dihydroxyvitamin D3 in long bones of rats.Calcif. Tissue Int.3819867986
    Crossref | PubMed | ISI | Google Scholar
  • 137 HOLICK M. F., GARABEDIAN M., DeLUCA H. F.1,25-Dihydroxycholecalciferol: metabolite of vitamin D3 active on bone in anephric rats.Science176197211461147
    Crossref | PubMed | ISI | Google Scholar
  • 138 HOLICK M. F., SCHNOES H. K., DeLUCA H. F., GRAY R. W., BOYLE I. T., SUDA T.Isolation and identification of 24,25-dihydroxycholecalciferol: a metabolite of vitamin D3 made in the kidney.Biochemistry11197242514255
    Crossref | PubMed | ISI | Google Scholar
  • 139 HOLICK M. F., SCHNOES H. K., DeLUCA H. F., SUDA T., COUSINS R. J.Isolation and identification of 1,25-dihydroxycholecalciferol. A metabolite of vitamin D active in the intestine.Biochemistry10197127992804
    Crossref | PubMed | ISI | Google Scholar
  • 140 HOLLIS, B. W. Detection of vitamin D and its major metabolites. In: Vitamin D, edited by D. Feldman, F. H. Glorieux, and J. W. Pike. San Diego, CA: Academic, 1997, p. 587–606.
    Google Scholar
  • 141 HORIUCHI N., SUDA T., TAKAHASHI H., SHAZAWA E., OGATA E.In vivo evidence for the intermediary role of 3′,5′cyclic AMP in parathyroid hormone-induced stimulation of 1α,25-dihydroxyvitamin D3 synthesis in rats.Endocrinology1011977969974
    Crossref | PubMed | ISI | Google Scholar
  • 142 HORLEIN A. J., NAAR A. M., HEINZEL T., TORCHIA J., GLOSS B., KUROKAWA R., RYAN A., KAMEI Y., SODERSTROM M., GLASS C. K.Ligand-dependent repression by the thyroid hormone receptor mediated by a nuclear receptor co-repressor.Nature3771995397404
    Crossref | PubMed | ISI | Google Scholar
  • 143 HORST R. L., KOSZEWSKI N. J., REINHARDT T. A.1α-Hydroxylation of 24-hydroxyvitamin D2 represents a minor physiological pathway for the activation of vitamin D2 in mammals.Biochemistry291990578582
    Crossref | PubMed | ISI | Google Scholar
  • 144 HOSOMI J., HOSOI J., ABE E., SUDA T., KUROKI T.Regulation of terminal differentiation of cultured mouse epidermal cells by 1α,25-dihydroxyvitamin D3 .Endocrinology113198319501957
    Crossref | PubMed | ISI | Google Scholar
  • 145 HUGHES M. R., HAUSSLER M. R.1,25-Dihydroxyvitamin D3 receptors in parathyroid glands. Preliminary characterization of cytoplasmic and nuclear binding components.J. Biol. Chem.252197810651073
    Google Scholar
  • 146 HUTCHINSON K. A., SCHERRER L. C., CZAR M. J., STANCATO L. F., CHOW Y. H., JOVE R., PRATT W. B.Regulation of glucocorticoid receptor function through assembly of a receptor-heat shock protein complex.Ann. NY Acad. Sci.68419933548
    Crossref | PubMed | ISI | Google Scholar
  • 147 ICHIKAWA F., SATO K., NANJO M., NISHII Y., SHINKI T., TAKAHASHI N., SUDA T.Mouse primary osteoblasts express vitamin D3 25-hydroxylase mRNA and convert 1α-hydroxyvitamin D3 into 1α,25-hydroxyvitamin D3 .Bone161995129135
    Crossref | PubMed | ISI | Google Scholar
  • 148 ITOH S., YOSHIMURA T., IEMURA O., YAMADA E., TSUJIKAWA K., KOHAMA Y., MIMURA T.Molecular cloning of 25-hydroxyvitamin D3 24-hydroxylase (CYP24) from mouse kidney: its inducibility by vitamin D3 .Biochim. Biophys. Acta126419952628
    Crossref | PubMed | ISI | Google Scholar
  • 149 JARNAGIN K., BROMMAGE R., DeLUCA H. F., YAMADA S., TAKAYAMA H.1- But not 24-hydroxylation of vitamin D is required for growth and reproduction in rats.Am. J. Physiol.244(Endocrinol. Metab. 7)1983E290E297
    Google Scholar
  • 150 JEHAN F., DeLUCA H. F.Cloning and characterization of the mouse vitamin D receptor promoter.Proc. Natl. Acad. Sci. USA9419971013810143
    Crossref | PubMed | ISI | Google Scholar
  • 151 JEHAN-KIMMEL C., DARWISH H. M., STRUGNELL S. A., WIEFLING B. A., DeLUCA H. F.Binding of vitamin D receptor to vitamin D response elements induces DNA distortion (Abstract).J. Bone Miner. Res.111996S312
    PubMed | ISI | Google Scholar
  • 152 JENSTER G., SPENCER T. E., BURCIN M. M., TSAI S. Y., TSAI M.-J., O'MALLEY B. W.Steroid receptor induction of gene transcription: a two-step model.Proc. Natl. Acad. Sci. USA94199778797884
    Crossref | PubMed | ISI | Google Scholar
  • 153 JEUNG E. B., LEUNG P. C., KRISINGER J.The human calbindin-D9k gene. Complete structure and implications on steroid hormone regulation.J. Mol. Biol.235199412311238
    Crossref | PubMed | ISI | Google Scholar
  • 154 JIN C. H., KERNER S. A., HONG M. H., PIKE J. W.Transcriptional activation and dimerization functions in the human vitamin D receptor.Mol. Endocrinol.101996945957
    PubMed | Google Scholar
  • 155 JIN C. H., PIKE J. W.Human vitamin D receptor-dependent transactivation in Saccharomyces cerevisiae requires retinoid X receptor.Mol. Endocrinol.101996196205
    PubMed | Google Scholar
  • 156 JONES G., BYFORD V., MAKIN H. L. J., KREMER R., RICE R. H., DEGRAFFENRIED L.-A., KNUTSON J. C., BISHOP C. W.Anti-proliferative activity and target cell catabolism of the vitamin D analog 1,24(S)-(OH)2D2 in normal and immortalized human epidermal cells.Biochem. Pharmacol.521996133140
    Crossref | PubMed | ISI | Google Scholar
  • 157 JONES G., CALVERLEY M. J.A dialogue on analogues: newer vitamin-D drugs for use in bone disease, psoriasis, and cancer.Trends Endocrinol. Metab.41993297303
    Crossref | PubMed | ISI | Google Scholar
  • 158 JONES G., EDWARDS N., VRIEZEN D., PORTEOUS C., TRAFFORD D. J. H., CUNNINGHAM J., MAKIN H. L. J.Isolation and identification of seven metabolites of 25-hydroxydihydrotachysterol 3 formed in the isolated perfused rat kidney: a model for the study of side-chain metabolism of vitamin D.Biochemistry27198870707079
    Crossref | PubMed | ISI | Google Scholar
  • 159 JONES G., KANO K., YAMADA S., FURUSAWA T., TAKAYAMA H., SUDA T.Identification of 24,25,26,27-tetranor-23-(OH)D3 as a product of renal metabolism of 24,25-(OH)2D3 .Biochemistry23198437493754
    Crossref | PubMed | ISI | Google Scholar
  • 160 JONES G., KUNG M., KANO K.The isolation and identification of two new metabolites of 25-hydroxyvitamin D3 produced in the kidney.J. Biol. Chem.25819831292012929
    PubMed | ISI | Google Scholar
  • 161 JONES G., SCHNOES H. K., LEVAN L., DeLUCA H. F.Isolation and identification of 24-hydroxyvitamin D2 and 24,25-dihydroxyvitamin D2 .Arch. Biochem. Biophys.2021980450457
    Crossref | PubMed | ISI | Google Scholar
  • 162 JURUTKA P. W., HSIEH J. C., NAKAJIMA S., HAUSSLER C. A., WHITFIELD G. K., HAUSSLER M. R.Human vitamin D receptor phosphorylation by casein kinase II at Ser-208 potentiates transcriptional activation.Proc. Natl. Acad. Sci. USA93199635193524
    Crossref | PubMed | ISI | Google Scholar
  • 163 KAHLEN J. P., CARLBERG C.Functional characterization of a 1,25-dihydroxyvitamin D3 receptor binding site in the rat atrial natriuretic factor promoter.Biochem. Biophys. Res. Commun.2181996882886
    Crossref | PubMed | ISI | Google Scholar
  • 164 KAMEI Y., KAWADA T., FUKUWATARI T., ONO T., KATO S., SUGIMOTO E.Cloning and sequencing of the gene encoding the mouse vitamin D receptor.Gene1521995281282
    Crossref | PubMed | ISI | Google Scholar
  • 165 KAMEI Y., XU L., HEINZEL T., TORCHIA J., KUROKAWA R., GLOSS B., LIN S. C., HEYMAN R., ROSE D. W., GLASS C. K.A CBP integrator complex mediates transcriptional activation and AP-1 inhibition by nuclear receptors.Cell851996403414
    Crossref | PubMed | ISI | Google Scholar
  • 166 KANE K. F., LANGMAN M. J. S., WILLIAMS G. R.Antiproliferative responses of two human colon cancer cell lines to vitamin D3 are differentially modified by 9-cis retinoic acid.Cancer Res.561996623632
    PubMed | ISI | Google Scholar
  • 167 KATO S., ENDOH H., MASUHIRO Y., KITAMOTO T., UCHIYAMA S., SASAKI H., MASUSHIGE S., GOTOH Y., NISHIDA E., KAWASHIMA H.Activation of the estrogen receptor through phosphorylation by mitogen-activated protein kinase.Science270199514911494
    Crossref | PubMed | ISI | Google Scholar
  • 168 KERRY D. M., DWIVEDI P. P., HAHN C. N., MORRIS H. A., OMDAHL J. L., MAY B. K.Transcriptional synergism between vitamin D-responsive elements in the rat 25-hydroxyvitamin D3 24-hydroxylase (CYP24) promoter.J. Biol. Chem.27119962971529721
    Crossref | PubMed | ISI | Google Scholar
  • 169 KIMMEL-JEHAN C., JEHAN F., DeLUCA H. F.Salt concentration determines 1,25-dihydroxyvitamin D3 dependency of vitamin D receptor-retinoid X receptor-vitamin D-responsive element complex.Arch. Biochem. Biophys.34119977580
    Crossref | PubMed | ISI | Google Scholar
  • 170 KISSMEYER A.-M., BINDERUP E., BINDERUP L., HANSEN C. M., ANDERSEN N. R., SCHROEDER N. J., MAKIN H. L. J., SHANKAR V. N., JONES G.The metabolism of the vitamin D analog EB1089: identification of in vivo and in vitro metabolites and their biological activities.Biochem. Pharmacol.53199710871097
    Crossref | PubMed | ISI | Google Scholar
  • 171 KISSMEYER A.-M., MATHIASEN I. S., LATINI S., BINDERUP L.Pharmacokinetic studies of vitamin D analogues: relationship to vitamin D binding protein (DBP).Endocrine31995263266
    Crossref | PubMed | ISI | Google Scholar
  • 172 KLEINER-BOSSALLER A., DeLUCA H. F.Formation of 1,24,25-trihydroxyvitamin D3 from 1,25-dihydroxyvitamin D3 .Biochim. Biophys. Acta3381974489495
    Crossref | ISI | Google Scholar
  • 173 KLIEWER S. A., UMESONO K., MANGELSDORF D. J., EVANS R. M.Retinoid X receptor interacts with nuclear receptors in retinoic acid, thyroid hormone and vitamin D3 signalling.Nature3551992446449
    Crossref | PubMed | ISI | Google Scholar
  • 174 KNUTSON J. C., DeLUCA H. F.25-Hydroxyvitamin D3-24-hydroxylase. Subcellular location and properties.Biochemistry13197415431548
    Crossref | PubMed | ISI | Google Scholar
  • 175 KNUTSON J. C., LEVAN L. W., VALLIERE C. R., BISHOP C. W.Pharmacokinetics and systemic effect on calcium homeostasis of 1,24-dihydroxyvitamin D2 in rats: comparison with calcitriol and calcipotriol.Biochem. Pharmacol.531997829837
    Crossref | PubMed | ISI | Google Scholar
  • 176 KOBAYASHI T., TSUGAWA N., OKANO T., MASUDA S., TAKEUCHI A., KUBODERA N., NISHII Y.The binding properties with blood proteins and tissue distribution of 22-oxa-1,25-dihydroxyvitamin D3 , a noncalcemic analogue of 1,25-dihydroxyvitamin D3 in rats.J. Biochem.1151994373380
    Crossref | PubMed | ISI | Google Scholar
  • 177 KOLI K., KESKI-OJA J.1,25-Dihydroxyvitamin D3 enhances the expression of transforming growth factor β1 and its latent form binding protein in cultured breast carcinoma cells.Cancer Res.55199515401546
    PubMed | ISI | Google Scholar
  • 178 KOSZEWSKI N. J., REINHARDT T. A., NAPOLI J. L., BEITZ D. C., HORST R. L.24,26-Dihydroxyvitamin D2: a unique physiological metabolite of vitamin D2 .Biochemistry27198857855790
    Crossref | PubMed | ISI | Google Scholar
  • 179 KOUZARIDES T., BANNISTER A. J.The CBP co-activator is a histone acetyltransferase.Nature3841997641643
    ISI | Google Scholar
  • 180 KOWARSKI S., SCHACHTER D.Effects of vitamin D on phosphate transport and incorporation into mucosal constituents of rat intestinal mucosa.J. Biol. Chem.2441969211217
    PubMed | ISI | Google Scholar
  • 181 KRAGBALLE K., GJERTSEN B. T., DE HOOP D., KARLSMARK T., VAN DE KERKHOF P. C., LARKO O., NIEBOER C., ROED-PETERSEN J., STRAND A., TIKJOB G.Double-blind, right/left comparison of calcipotriol and betamethasone valerate in treatment of psoriasis vulgaris.Lancet3371991193196
    Crossref | PubMed | ISI | Google Scholar
  • 182 KREMER R., SEBAG M., CHAMPIGNY C., MEEROVITCH K., HENDY G. N., WHITE J., GOLTZMAN D.Identification and characterization of 1,25-dihydroxyvitamin D3-responsive repressor sequences in the rat parathyroid hormone-related peptide gene.J. Biol. Chem.27119961631016316
    Crossref | PubMed | ISI | Google Scholar
  • 183 KRISHNAN A. V., FELDMAN D.Cyclic adenosine 3′,5′-monophosphate up-regulates 1,25-dihydroxyvitamin D3 receptor gene expression and enhances hormone action.Mol. Endocrinol.61992198206
    PubMed | Google Scholar
  • 184 KRONENBERG H., IGARASHI T., FREEMAN M. W., OKAZAKI T., BRAND S. J., WIREN K. M., POTTS J. T.Jr. Structure and expression of the human parathyroid hormone gene.Rec. Prog. Horm. Res.421986641663
    PubMed | Google Scholar
  • 185 KUMAR R.Metabolism of 1,25-dihydroxyvitamin D3 .Physiol. Rev.641984478504
    Link | ISI | Google Scholar
  • 186 KUMAR, R. Vitamin D and the kidney. In: Vitamin D, edited by D. Feldman, F. H. Glorieux, and J. W. Pike. San Diego, CA: Academic, 1997, chapt. 17, p. 275–292.
    Google Scholar
  • 187 KUMAR R., SCHNOES H. K., DeLUCA H. F.Rat intestinal 25-hydroxyvitamin D3- and 1α,25-dihydroxyvitamin D3-24-hydroxylase.J. Biol. Chem.253197838043809
    PubMed | ISI | Google Scholar
  • 188 KUREK-TYRLIK, A., F. Z. MAKAEV, J. WICHA, V. ZHABINSKII, and M. J. CALVERLEY. Pivoting 20-normal and 20-epi calcitriols: synthesis and crystal structure of a “doucle side chain” analogue. In: Vitamin D. Chemistry, Biology and Clinical Applications of the Steroid Hormone, edited by A. W. Norman, R. Bouillon, and M. Thomasset. Riverside: Univ. of California, 1997, p. 31–32.
    Google Scholar
  • 189 KUROKAWA R., DIRENZO J., BOEHM M., SUGARMAN J., GLOSS B., ROSENFELD M. G., HEYMAN R. A., GLASS C. K.Regulation of retinoid signalling by receptor polarity and allosteric control of ligand binding.Nature3711994528531
    Crossref | PubMed | ISI | Google Scholar
  • 190 KWIECINSKI G. G., PETRIE G. I., DeLUCA H. F.Vitamin D is necessary for reproductive functions in the male rat.J. Nutr.1191989741744
    Crossref | PubMed | ISI | Google Scholar
  • 191 KWIECINSKI G. G., PETRIE G. I., DeLUCA H. F.1,25-Dihydroxyvitamin D3 restores fertility of vitamin D-deficient female rats.Am J. Physiol.256(Endocrinol. Metab. 19)1989E483E487
    Google Scholar
  • 192 LABUDA M., MORGAN K., GLORIEUX F. H.Mapping autosomal recessive vitamin D dependency type I to chromosome 12q14 by linkage analysis.Am. J. Human Genet.4719902836
    PubMed | ISI | Google Scholar
  • 193 LAM H. Y., SCHNOES H. K., DeLUCA H. F.1-Hydroxyvitamin D2: a potent synthetic analog of vitamin D2 .Science186197410381040
    Crossref | PubMed | ISI | Google Scholar
  • 194 LEE M. S., KLIEWER S. A., PROVENCAL J., WRIGHT P. E., EVANS R. M.Structure of the retinoid X receptor α DNA-binding domain: a helix required for homodimeric DNA binding.Science260199311171121
    Crossref | PubMed | ISI | Google Scholar
  • 195 LEMIRE, J. The role of vitamin D3 in immunosuppression: lessons from autoimmunity and transplantation. In: Vitamin D, edited by D. Feldman, F. H. Glorieux, and J. W. Pike. San Diego, CA: Academic, 1997, chapt. 69, p. 1167–1181.
    Google Scholar
  • 196 LEMIRE J. M.Immunomodulatory role of 1,25-dihydroxyvitamin D3 .J. Cell Biochem.4919922631
    Crossref | PubMed | ISI | Google Scholar
  • 197 LEMON B. D., FREEDMAN L. P.Selective effects of ligands on vitamin D3 receptor- and retinoid X receptor-mediated gene activation in vivo.Mol. Cell. Biol.16199610061016
    Crossref | PubMed | ISI | Google Scholar
  • 198 LI Q., GARDNER D. G.Negative regulation of the human atrial natriuretic peptide gene by 1,25-dihydroxyvitamin D3 .J. Biol. Chem.269199449344939
    PubMed | ISI | Google Scholar
  • 199 LIU M., LEE M. H., COHEN M., BOMMANKANTI M., FREEDMAN L. P.Transcriptional activation of the Cdk inhibitor p21 by vitamin D3 leads to the induced differentiation of the myelomonocytic cell line U937.Genes Dev.101996142153
    Crossref | PubMed | ISI | Google Scholar
  • 200 LIU Y.-Y., COLLINS E. D., NORMAN A. W., PELEG S.Differential interaction of 1,25-dihydroxyvitamin D3 analogues and their 20-epi homologues with the vitamin D receptor.J. Biol. Chem.272199733363345
    Crossref | PubMed | ISI | Google Scholar
  • 201 LLACH F., BRICKMAN A. S., COBURN J. W.Unique effects of 24,25-dihydroxyvitamin D3 in uremic patients.Contr. Nephrol.181980212217
    Crossref | PubMed | Google Scholar
  • 202 LOHNES D., JONES G.Side chain metabolism of vitamin D3 in osteosarcoma cell line UMR-106. Characterization of products.J. Biol. Chem.26219871439414401
    PubMed | ISI | Google Scholar
  • 203 LOHNES, D., and G. JONES. Further metabolism of 1,25-dihydroxyvitamin D3 in target cells. Proceedings of the First International Congress on Vitamins and Biofactors in Life Science. J. Nutr. Sci. Vitaminol. Special Issue: 75–78, 1992.
    Google Scholar
  • 204 LUPISELLA J. A., DRISCOLL J. E., METZLER W. J., RECZEK P. R.The ligand binding domain of the human retinoic acid receptor γ is predominantly α-helical with a Trp residue in the ligand binding site.J. Biol. Chem.27019952488424890
    Crossref | PubMed | ISI | Google Scholar
  • 205 MacDONALD P. N., DOWD D. R., NAKAJIMA S., GALLIGAN M. A., REEDER M. C., HAUSSLER C. A., OZATO K., HAUSSLER M. R.Retinoid X receptors stimulate and 9-cis retinoic acid inhibits 1,25-dihydroxyvitamin D3-activated expression of the rat osteocalcin gene.Mol. Cell. Biol.13199359075917
    Crossref | PubMed | ISI | Google Scholar
  • 206 MacDONALD P. N., SHERMAN D. R., DOWD D. R., JEFCOAT J. R., DELISLE R. K.The vitamin D receptor interacts with general transcription factor IIβ.J. Biol. Chem.270199547484752
    Crossref | PubMed | ISI | Google Scholar
  • 207 MADHOK T. C., DeLUCA H. F.Characteristics of the rat liver microsomal enzyme system converting cholecalciferol to 25-hydroxycholecalciferol. Evidence for participation of cytochrome P-450.Biochem. J.1841979491499
    Crossref | PubMed | ISI | Google Scholar
  • 208 MAKIN G., LOHNES D., BYFORD V., RAY R., JONES G.Target cell metabolism of 1,25-dihydroxyvitamin D3 to calcitroic acid. Evidence for a pathway in kidney and bone involving 24-oxidation.Biochem. J.2621989173180
    Crossref | PubMed | ISI | Google Scholar
  • 209 MALLOY, P. J., J. W. PIKE, and D. FELDMAN. Hereditary 1,25-dihydroxyvitamin D resistant rickets. In: Vitamin D, edited by D. Feldman, F. H. Glorieux, and J. W. Pike. San Diego, CA: Academic, 1997, p. 765–787.
    Google Scholar
  • 210 MANGELSDORF D. J., EVANS R. M.The RXR heterodimers and orphan receptors.Cell831995841850
    Crossref | PubMed | ISI | Google Scholar
  • 211 MANGELSDORF D. J., ONG E. S., DYCK J. A., EVANS R. M.Nuclear receptor that identifies a novel retinoic acid response pathway.Nature3451990224229
    Crossref | PubMed | ISI | Google Scholar
  • 212 MARTIN, T. J. The roles of PTHrP in normal physiology and in cancer. In: The First International Forum on Calcified Tissue and Bone Metabolism, edited by T. Suda. Tokyo: Chugai, 1993, p. 48–53.
    Google Scholar
  • 213 MASUDA S., BYFORD V., KREMER R., MAKIN H. L. J., KUBODERA N., NISHII Y., OKAZAKI A., OKANO T., KOBAYASHI T., JONES G.In vitro metabolism of the vitamin D analog, 22-oxacalcitriol, using cultured osteosarcoma, hepatoma and keratinocyte cell lines.J. Biol. Chem.271199687008708
    Crossref | PubMed | ISI | Google Scholar
  • 214 MASUDA S., OKANO T., KAMAO M., KREMER R., JONES G., KOBAYASHI T.Different metabolism of 1,25-dihydroxyvitamin D3 in target cells (Abstract).J. Bone Miner. Res.111996S420
    ISI | Google Scholar
  • 215 MASUDA S., STRUGNELL S., CALVERLY M. J., MAKIN H. L. J., KREMER R., JONES G.In vitro metabolism of the anti-psoriatic vitamin D analog, calcipotriol, in two cultured human keratinocyte models.J. Biol. Chem.269199447944803
    PubMed | ISI | Google Scholar
  • 216 MASUYAMA H., BROWNFIELD C. M., ST.-ARNAUD R., MacDONALD P. N.Evidence for ligand-dependent intramolecular folding of the AF-2 domain in vitamin D receptor-activated transcription and coactivator interaction.Mol. Endocrinol.11199715071517
    Crossref | PubMed | Google Scholar
  • 217 MASUYAMA H., JEFCOAT S. C.Jr., and P. N. MacDONALD. The N-terminal domain of transcription factor TFIIB is required for direct interaction with the vitamin D receptor and participates in vitamin D-mediated transcription.Mol. Endocrinol.111997218228
    Crossref | PubMed | Google Scholar
  • 218 MATHIEU C., LAUREYS J., SOBIS H., VANDEPUTTE M. V., BOUILLON R.1,25-Dihydroxyvitamin D3 prevents insulitis in NOD mice.Diabetes41199214911495
    Crossref | PubMed | ISI | Google Scholar
  • 219 McDONNELL D. P., MANGELSDORF D. J., PIKE J. W., HAUSSLER M. R., O'MALLEY B. W.Molecular cloning of complementary DNA encoding the avian receptor for vitamin D.Science235198712141217
    Crossref | PubMed | ISI | Google Scholar
  • 220 McDONNELL D. P., SCOTT R. A., KERNER S. A., O'MALLEY B. W., PIKE J. W.Functional domains of the human vitamin D3 receptor regulate osteocalcin gene expression.Mol. Endocrinol.31989635644
    Crossref | PubMed | Google Scholar
  • 221 McSHEEY P. M., CHAMBERS T. J.Osteoblast-like cells in the presence of parathyroid hormone release soluble factor that stimulates osteoclastic bone resorption.Endocrinology119198616541659
    Crossref | PubMed | ISI | Google Scholar
  • 222 McSHEENY P. M. J., CHAMBERS T. J.1,25-Dihydroxyvitamin D3 stimulates rat osteoblastic cells to release a soluble factor that increases osteoclastic bone resorption.J. Clin. Invest.801987425429
    Crossref | PubMed | ISI | Google Scholar
  • 223 MERKE J., KLAUS G., HUGEL U., WALDHERR R., RITZ E.No 1,25-dihydroxyvitamin D3 receptors on osteoclasts of calcium-deficient chicken despite demonstrable receptors on circulating monocytes.J. Clin. Invest.771986312314
    Crossref | PubMed | ISI | Google Scholar
  • 224 MILLER B. E., CHIN D. P., JONES G.1,25-Dihydroxyvitamin D3 metabolism in a human osteosarcoma cell line and human bone cells.J. Bone Miner. Res.51990597607
    Crossref | PubMed | ISI | Google Scholar
  • 225 MITHAL A., KIFOR O., KIFOR I., VASSILEV P., BUTTERS R., KRAPCHO K., SIMIN R., FULLER F., HEVERT S. C., BROWN E. M.The reduced responsiveness of cultured bovine parathyroid cells to extracellular Ca2+ is associated with marked reduction in the expression of extracellular Ca2+-sensing receptor messenger ribonucleic acid and protein.Endocrinology136199530873092
    Crossref | PubMed | ISI | Google Scholar
  • 226 MIYAMOTO K., KESTERSON R. A., YAMAMOTO H., TAKETANI Y., NISHIWAKII E., TATSUMI S., INOUE Y., MORITA K., TAKEDA E., PIKE J. W.Structural organization of the human vitamin D receptor chromosomal gene and its promoter.Mol. Endocrinol.11199711651179
    Crossref | PubMed | Google Scholar
  • 227 MIYAMOTO Y., SHINKI T., YAMAMOTO K., OHYAMA Y., IWASAKI H., HOSOTANI R., KASAMA T., AKAYAMA H., YAMADA S., SUDA T.1,25-Dihydroxyvitamin D3-24-hydroxylase (CYP24) hydroxylates the carbon at the end of the side chain (C-26) of the C-24-fluorinated analog of 1,25-dihydroxyvitamin D3 .J. Biol. Chem.27219971411514119
    Crossref | PubMed | ISI | Google Scholar
  • 228 MIYAURA C., ABE E., KURIBAYASHI T., TANAKA H., KONNO K., NISHII Y., SUDA T.1,25-Dihydroxyvitamin D3 induces differentiation of human myeloid leukemia cells.Biochem. Biophys. Res. Commun.1021981937943
    Crossref | PubMed | ISI | Google Scholar
  • 228a MONKAWA T., YOSHIDA T., WAKINO S., SHINKI T., ANAZAWA H., DeLUCA H. F., SUDA T., HAYASHI M., SARUTA T.Molecular cloning of cDNA and genomic DNA for human 25-hydroxyvitamin D3 1α-hydroxylase.Biochem. Biophys. Res. Commun.2391997527533
    Crossref | PubMed | ISI | Google Scholar
  • 229 MORRISON N. A., EISMAN J. A.Nonhypercalcemic 1,25-(OH)2D3 analogs potently induce the human osteocalcin gene promoter stably transfectd into rat osteosarcoma cells (ROSCO-2).J. Bone Miner. Res.61991893899
    Crossref | PubMed | ISI | Google Scholar
  • 230 MORTENSEN B. M., GORGELADZE J. O., LYNGDAL P. T., AARSETH H. P., GAUTVIK K. M.Alterations in serum and urine parameters reflecting bone turnover in uremic patients during treatment with 1,25-dihydroxyvitamin D3 and 24,25-dihydroxyvitamin D3 .Miner. Electrolyte Metab.1919937885
    PubMed | Google Scholar
  • 231 MUNDER M., HERZBERG I. M., ZIEROLD C., MOSS V. E., HANSON K., CLAGETT-DAME M., DeLUCA H. F.Identification of the porcine intestinal accessory factor that enables DNA sequence recognition by vitamin D receptor.Proc. Natl. Acad. Sci. USA92199527952799
    Crossref | PubMed | ISI | Google Scholar
  • 232 MURAYAMA E., MIYAMOTO K., KUBODERA N., MORI T., MATSUNAGA I.Synthetic studies of vitamin D analogues. VIII. Synthesis of 22-oxavitamin D3 analogues.Chem. Pharm. Bull.34198644104413
    Crossref | PubMed | ISI | Google Scholar
  • 233 NAGY L., KAO H.-Y., CHAKRAVARTI D., LIN R. J., HASSIG C. A., AYER D. E., SCHREIBER S. L., EVANS R. M.Nuclear receptor repression mediated by a complex containing SMRT, mSin3A, and histone deacetylase.Cell891997373380
    Crossref | PubMed | ISI | Google Scholar
  • 234 NAKADA M., TANAKA Y., SCHNOES H. K., DeLUCA H. F., KOBAYASHI Y., IKEKAWA N.Biological activities and binding properties of 23,23-difluoro-25-hydroxyvitamin D3 and its 1-hydroxy derivative.Arch. Biochem. Biophys.2411985173178
    Crossref | PubMed | ISI | Google Scholar
  • 235 NAKAJIMA S., HSIEH J. C., JURUTKA P., GALLIGAN M. A., HAUSSLER C. A., WHITFIELD G. K., HAUSSLER M. R.Examination of the potential functional role of conserved cysteine residues in the hormone binding domain of the human 1,25-dihydroxyvitamin D3 receptor.J. Biol. Chem.271199651435149
    Crossref | PubMed | ISI | Google Scholar
  • 236 NAKAJIMA S., HSIEH J. C., MacDONALD P. N., GALLIGAN M. A., HAUSSLER C. A., WHITFIELD G. K., HAUSSLER M. R.The C-terminal region of the vitamin D receptor is essential to form a complex with a receptor auxiliary factor required for high affinity binding to the vitamin D-responsive element.Mol. Endocrinol.81994159172
    PubMed | Google Scholar
  • 237 NAPOLI J. L., HORST R. L.C(24)- and C(23)-oxidation, converging pathways of intestinal 1,25-dihydroxyvitamin D3 metabolism: identification of 24-keto-1,24,25-trihydroxyvitamin D3 .Biochemistry22198358485853
    Crossref | PubMed | ISI | Google Scholar
  • 238 NAVEH-MANY, T., and J. SILVER. Parathyroid hormone synthesis, secretion and action. In: Kidney Stones: Medical and Surgical Management, edited by F. L. Coe, M. J. Favus, C. Pak, J. Parks, and G. Preminger. New York: Raven, 1996, p. 175–199.
    Google Scholar
  • 239 NICOLAYSEN R., EEG-LARSEN N.The biochemistry and physiology of vitamin D.Vitamin Horm.1119532960
    Crossref | PubMed | ISI | Google Scholar
  • 240 NIKOLOV D. B., CHEN H., HALAY E. D., HOFFMAN A., ROEDER G., BURLEY S. K.Crystal structure of a human TATA box-binding protein/TATA element complex.Proc. Natl. Acad. Sci. USA93199648624867
    Crossref | PubMed | ISI | Google Scholar
  • 241 NISHII Y., SATO K., KOBAYASHI T.The development of vitamin D analogues for the treatment of osteoporosis.Osteoporosis Int.1, Suppl.1993S190S193
    Crossref | Google Scholar
  • 242 NISHIKAWA J.-I., KITAUA M., MASAYOSHI I., NISHIHARA T.Vitamin D receptor contains multiple dimerization interfaces that are functionally different.Nucleic Acids Res.231995606611
    Crossref | PubMed | ISI | Google Scholar
  • 243 NODA M., VOGEL R. L., CRAIG A. M., PRAHL J., DeLUCA H. F., DENHARDT D. T.Identification of a DNA sequence responsible for binding of the 1,25-dihydroxyvitamin D3 receptor and 1,25-dihydroxyvitamin D3 enhancement of mouse secreted phosphoprotein 1 (Spp-1 or osteopontin) gene expression.Proc. Natl. Acad. Sci. USA87199099959999
    Crossref | PubMed | ISI | Google Scholar
  • 244 NORMAN, A. W. Rapid biological responses initiated by 1,25-dihydroxyvitamin D3: a case study of transcaltachia (rapid hormonal stimulation of intestinal calcium transport). In: Vitamin D, edited by D. Feldman, F. H. Glorieux, and J. W. Pike. San Diego, CA: Academic, 1997, chapt. 15, p. 233–258.
    Google Scholar
  • 245 NORMAN A. W., HENRY H. L., MALLUCHE H. H.24R,25-Dihydroxyvitamin D3 and 1,25-dihydroxyvitamin D3 are both indispensable for calcium and phosphorus homeostasis.Life Sci.271980229231
    Crossref | PubMed | ISI | Google Scholar
  • 246 OHNUMA N., NORMAN A. W.Identification of a new C-23 oxidation pathway of metabolism for 1,25-dihydroxyvitamin D3 present in intestine and kidney.J. Biol. Chem.257198282618271
    PubMed | ISI | Google Scholar
  • 247 OHYAMA Y., NOSHIRO M., OKUDA K.Cloning and expression of cDNA encoding 25-hydroxyvitamin D3 24-hydroxylase.FEBS Lett.2781991195198
    Crossref | PubMed | ISI | Google Scholar
  • 248 OHYAMA Y., OKUDA K.Isolation and characterization of a cytochrome P-450 from rat kidney mitochondria that catalyzes the 24-hydroxylation of 25-hydroxyvitamin D3 .J. Biol. Chem.266199186908695
    PubMed | ISI | Google Scholar
  • 249 OHYAMA Y., OZONO K., UCHIDA M., SHINKI T., KATO S., SUDA T., YAMAMOTO O., NOSHIRO M., KATO Y.Identification of a vitamin D-responsive element in the 5'-flanking region of the rat 25-hydroxyvitamin D3 24-hydroxylase gene.J. Biol. Chem.26919941054510550
    PubMed | ISI | Google Scholar
  • 250 OKAMURA W. H., MIDLAND M. M., NORMAN A. W., HAMMOND M. W., DORMANEN M. C., NEMERE I.Biochemical significance of the 6-s-cis conformation of the steroid hormone 1,25-dihydroxyvitamin D3 based on the provitamin D skeleton.Ann. NY Acad. Sci.7611995344348
    Crossref | PubMed | ISI | Google Scholar
  • 251 OKAZAKI T., IGARASHI T., KRONENBERG H. M.5′-Flanking region of the parathyroid hormone gene mediates negative regulation by 1,25-(OH)2 vitamin D3 .J. Biol. Chem.263198822032208
    PubMed | ISI | Google Scholar
  • 252 OKUDA K. I., USUI E., OHYAMA Y.Recent progress in enzymology and molecular biology of enzymes involved in vitamin D metabolism.J. Lipid Res.36199516411652
    PubMed | ISI | Google Scholar
  • 253 ONATE S. A., TSAI S. Y., TSAI M. J., O'MALLEY B. W.Sequence and characterization of a coactivator for the steroid hormone receptor superfamily.Science270199513541357
    Crossref | PubMed | ISI | Google Scholar
  • 254 ORIMO H., SHIRAKI M., HAYASHI T., NAKAMURA T.Reduced occurrence of vertebral crush fractures in senile osteoporosis treated with 1-(OH)-vitamin D3 .Bone Miner. Res.319874752
    Google Scholar
  • 255 ORNOY A., GOODWIN D., NOFF D., EDELSTEIN S.24,25-Dihydroxyvitamin D is a metabolite of vitamin D essential for bone formation.Nature2761978517519
    Crossref | PubMed | ISI | Google Scholar
  • 256 ORR W. J., HOLT L. E.Jr., L. WILKINS, and F. H. BOONE. The calcium and phosphorus metabolism in rickets, with special reference to ultraviolet ray therapy.Am. J. Dis. Child.261923362372
    Google Scholar
  • 257 OTT S., CHESNUT C. H.Calcitriol treatment is not effective in post-menopausal osteoporosis.Ann. Intern. Med.1181989267274
    Crossref | ISI | Google Scholar
  • 258 PAAREN H. E., HAMER D. E., SCHNOES H. K., DeLUCA H. F.Direct C-1 hydroxylation of vitamin D compounds: convenient preparation of 1-hydroxyvitamin D3 , 1,25-dihydroxyvitamin D3 and 1-hydroxyvitamin D2 .Proc. Natl. Acad. Sci. USA75197820802081
    Crossref | PubMed | ISI | Google Scholar
  • 259 PELEG S., SASTRY M., COLLINS E. D., BISHOP J. E., NORMAN A. W.Distinct conformational changes induced by 20-epi analogues of 1,25-dihydroxyvitamin D3 are associated with enhanced activation of the vitamin D receptor.J. Biol. Chem.27019951055110558
    Crossref | PubMed | ISI | Google Scholar
  • 260 PERLMAN K. L., SICINSKI R. R., SCHNOES H. K., DeLUCA H. F.1,25-Dihydroxy-19-nor-vitamin D3 , a novel vitamin D-related compound with potential therapeutic activity.Tetrahedron Lett.31199018231824
    Crossref | ISI | Google Scholar
  • 261 PERLMANN T., UMESONO K., RANGARAJAN P. N., FORMAN B. N., EVANS R. M.Two distinct dimerization interfaces differentially modulate target gene specificity of nuclear hormone receptors.Mol. Endocrinol.9199611661179
    Google Scholar
  • 262 PIEDRAFITA F. J., ORTIZ M. A., PFAHL M.Thyroid hormone receptor β mutants associated with generalized resistance to thyroid hormone show defects in their ligand-sensitive repression function.Mol. Endocrinol.9199515331548
    PubMed | Google Scholar
  • 263 PILLAI S., BIKLE D. D.Role of intracellular free calcium in the cornified envelope formation of keratinocytes: differences in the mode of action of extracellular calcium and 1,25 dihydroxyvitamin D.J. Cell. Physiol.146199194100
    Crossref | PubMed | ISI | Google Scholar
  • 264 POLLY P., CARLBERG C., EISMAN J. A., MORRISON N. A.Identification of a vitamin D response element in the fibronectin gene that is bound by a vitamin D3 receptor homodimer.J. Cell. Biochem.601996322333
    Crossref | PubMed | ISI | Google Scholar
  • 265 POLS H. A. P., BIRKENHAGEN J. C., SCHILTE J. P., VISSER T. J.Evidence that the self-induced metabolism of 1,25-dihydroxyvitamin D3 limits the homologous up-regulation of its receptor in rat osteosarcoma cells.Biochim. Biophys. Acta9701988122129
    Crossref | PubMed | ISI | Google Scholar
  • 266 POTTS, J. T., Jr Biological surprises and evolutionary links in the calcium field. Parathyroid hormone (PTH): past, present and future. In: PTH and PTHrP, edited by T. Suda. Tokyo: Chugai, 1993, p. 6–11.
    Google Scholar
  • 267 PROVVEDINI D. M., RULOT C. M., SOBOL R. E., TSOUKAS C. D., MANOLAGAS S. C.1α,25-Dihydroxyvitamin D, receptors in human thymic and tonsillar lymphocytes.J. Bone Miner. Res.21987239241
    Crossref | PubMed | ISI | Google Scholar
  • 268 QAW F., CALVERLEY M. J., SCHROEDER N. J., TRAFFORD D. J. H., MAKIN H. L. J., JONES G.In vivo metabolism of the vitamin D analog, dihydrotachysterol. Evidence for formation of 1,25- and l,25-dihydroxy-dihydrotachysterol metabolites and studies of their biological activity.J. Biol. Chem.2681993282292
    PubMed | ISI | Google Scholar
  • 269 RAISZ L. G., TRUMMEL C. L., HOLICK M. F., DeLUCA H. F.1,25-Dihydroxycholecalciferol: a potent stimulator of bone resorption in tissue culture.Science1751972768769
    Crossref | PubMed | ISI | Google Scholar
  • 270 RASMUSSEN H., DeLUCA H. F.Calcium homeostasis.Ergebnisse Physiol.531963107173
    Crossref | Google Scholar
  • 271 RASMUSSEN H., DeLUCA H. F., ARNAUD C., HAWKER C., VON STEDINGK M.The relationship between vitamin D and parathyroid hormone.J. Clin. Invest.42196319401946
    Crossref | PubMed | ISI | Google Scholar
  • 272 RASTINEJAD F., PERLMANN T., EVANS R. M., SIGLER P. B.Structural determinants of nuclear receptor assembly on DNA direct repeats.Nature3751995203211
    Crossref | PubMed | ISI | Google Scholar
  • 273 RAY R., SWAMY N., MacDONALD P. N., RAY S., HAUSSLER M. R., HOLICK M. F.Affinity labelling of the 1α,25-dihydroxyvitamin D3 receptor regulate osteocalcin gene expression.J. Biol. Chem.271199620122017
    Crossref | PubMed | ISI | Google Scholar
  • 274 RECKER R., SCHOENFELD P., LETTERI J., SLATOPOLSKY E., GOLDSMITH R., BRICKMAN A.The efficacy of calcifediol in renal osteodystrophy.Arch. Intern. Med.1381978857863
    Crossref | PubMed | Google Scholar
  • 275 REDDY, G. S., M.-L. SIU-CALDERA, I. SCHUSTER, N. ASTECKER, K.-Y. TSERNG, K. R. MULRALIDHARAN, W. H. OKAMURA, J. A. McCLANE, and M. R. USKOKOVIC. Target tissue specific metabolism of 1,25-(OH)2D3 through A-ring modification. In: Vitamin D. Chemistry, Biology and Clinical Applications of the Steroid Hormone, edited by A. W. Norman, R. Bouillon, and M. Thomasset. Riverside: Univ. of California, 1997, p. 139–146.
    Google Scholar
  • 276 REDDY G. S., TSERNG K.-Y.Calcitroic acid, end product of renal metabolism of 1,25-dihydroxyvitamin D3 through C-24 oxidation pathway.Biochemistry28198917631769
    Crossref | PubMed | ISI | Google Scholar
  • 277 REEVE L., TANAKA Y., DeLUCA H. F.Studies on the site of 1,25-dihydroxyvitamin D3 synthesis in vivo.J. Biol. Chem.258198336153617
    PubMed | ISI | Google Scholar
  • 278 REINHARDT T. A., HORST R. L.Self induction of 1,25-dihydroxyvitamin D3 metabolism limits receptor occupancy and target cell responsiveness.J. Biol. Chem.26419891591715921
    PubMed | ISI | Google Scholar
  • 279 RENAUD J. P., ROCHEL N., RUFF M., VIVAT V., CHAMBON P., GRONEMEYER H., MORAS D.Crystal structure of the RAR-γ ligand-binding domain bound to all-trans retinoic acid.Nature3781995681689
    Crossref | PubMed | ISI | Google Scholar
  • 280 RENNO T., KRAKOWSKI M., PICCIRILLO C., LIN J., OWENS T.TNF-α expression by resident microglia and infiltrating leukocytes in the central nervous system of mice with experimental allergic encephalomyelitis: regulation by Th1 cytokines.J. Immunol.1541995944953
    PubMed | ISI | Google Scholar
  • 281 ROSENTHAL A. M., JONES G., KOOH S. W., FRASER D.25-Hydroxyvitamin D3 metabolism by the isolated perfused rat kidney.Am. J. Physiol.239(Endocrinol. Metab. 2)1980E12E20
    Google Scholar
  • 282 ROSS T. K., PRAHL J. M., DeLUCA H. F.Overproduction of rat 1,25-dihydroxyvitamin D3 receptor in insect cells using the baculovirus expression system.Proc. Natl. Acad. Sci. USA88199165556559
    Crossref | PubMed | ISI | Google Scholar
  • 283 RUBIN, R. P., G. B. WEISS, and J. W. PUTNEY, Jr (Editors). Calcium in Biological Systems. New York: Plenum, 1985, p. 1–737.
    Google Scholar
  • 284 RUSSELL J., LETTIERI D., SHERWOOD L. M.Suppression by 1,25(OH)2D3 of transcription of the pre-proparathyroid hormone gene.Endocrinology119198628642867
    Crossref | PubMed | ISI | Google Scholar
  • 285 SAAREM K., BERGSETH S., OFTEBRO H., PEDERSEN J. I.Subcellular localization of vitamin D3 25-hydroxylation in human liver.J. Biol. Chem.25919841093610946
    PubMed | ISI | Google Scholar
  • 286 SAITOU M., NARUMIYA S., KAKIZUKA A.Alteration of a single amino acid residue in retinoic acid receptor causes a dominant negative phenotype.J. Biol. Chem.26919941910119107
    PubMed | ISI | Google Scholar
  • 287 SASAKI H., HARADA H., HANADA Y., MORINO H., SUZAWA M., SHIMPO E., KATSUMATA T., MASUHIRO Y., MATSUDA K., EBIHARA K.Transcriptional activity of a fluorinated vitamin D analog on VDR-RXR-mediated gene suppression.Biochemistry341995370377
    Crossref | PubMed | ISI | Google Scholar
  • 288 SATCHELL D. P., NORMAN A. W.Metabolism of the cell differentiating agent 1,25-(OH)2-16-ene-23-yne vitamin D3 by leukemic cells.J. Steroid Biochem. Mol. Biol.571996117124
    Crossref | PubMed | ISI | Google Scholar
  • 289 SCHNEIDER H., FEYEN J. H., SEUWEN K., MOVVA N. R.Cloning and functional expression of a human parathyroid hormone receptor.Eur. J. Pharmacol.2461993149155
    Crossref | PubMed | ISI | Google Scholar
  • 290 SCHRADER M., KAHLEN J. P., CARLBERG C.Functional characterization of a novel type of 1,25-dihydroxyvitamin D3 response element identified in the mouse c-fos promoter.Biochem. Biophys. Res. Commun.2301997646651
    Crossref | PubMed | ISI | Google Scholar
  • 291 SCHRADER M., MULLER K. M., CARLBERG C.Specificity and flexibility of vitamin D signalling.J. Biol. Chem.269199455015504
    PubMed | ISI | Google Scholar
  • 292 SCHRADER M., NAYERI S., KAHLEN J. P., MULLER K. M., CARLBERG C.Natural vitamin D3 response elements formed by inverted palindromes: polarity-directed ligand sensitivity of vitamin D3 receptor-retinoid X receptor heterodimer-mediated transactivation.Mol. Cell. Biol.15199511541161
    Crossref | PubMed | ISI | Google Scholar
  • 293 SEEMAN E., TSALAMANDRIS C., BASS S., PEARCE G.Present and future of osteoporosis therapy.Bone17, Suppl.199523S29S
    Crossref | ISI | Google Scholar
  • 294 SEMMLER E. J., HOLICK M. F., SCHNOES H. K., DeLUCA H. F.The synthesis of 1,25-dihydroxycholecalciferol: a metabolically active form of vitamin D3 .Tetrahedron Lett.40197241474150
    Crossref | Google Scholar
  • 295 SHANKAR V. N., DILWORTH F. J., MAKIN H. L. J., SCHROEDER N. J., TRAFFORD D. A. J., KISSMEYER A.-M., CALVERLEY M. J., BINDERUP E., JONES G.Metabolism of the vitamin D analog EB1089 by cultured human cells: redirection of hydroxylation site to distal carbons of the side-chain.Biochem. Pharmacol.531997783793
    Crossref | PubMed | ISI | Google Scholar
  • 296 SHERWOOD L. M., RUSSELL J.The role of 1,25-(OH)2D3 in regulating parathyroid gland function.Proc. Soc. Exp. Biol. Med.1911989233237
    Crossref | PubMed | ISI | Google Scholar
  • 297 SHINKI T., HIN C. H., NISHIMURA A., NAGAI Y., OHYAMA Y., NOSHIRO M., OKUDA K., SUDA T.Parathyroid hormone inhibits 25-hydroxyvitamin D3-24-hydroxylase mRNA expression stimulated by 1α,25-dihydroxyvitamin D3 in rat kidney but not in intestine.J. Biol. Chem.26719921375713762
    Crossref | PubMed | ISI | Google Scholar
  • 298 SHINKI T., SHIMADA H., WAKINO S., ANAZAWA H., HAYASHI M., SARUTA T., DeLUCA H. F., SUDA T.Cloning and expression of rat 25-hydroxyvitamin D3-1α-hydroxylase cDNA.Proc. Natl. Acad. Sci. USA9419981292012925
    Crossref | ISI | Google Scholar
  • 299 SHULTZ, T. D., F. J. H. HEATH III, and R. KUMAR. Do tissues other than the kidney produce 1,25-dihydroxyvitamin D3 in vivo? A reexamination. Proc. Natl. Acad. Sci. USA 80: 1746–1750, 1983.
    Google Scholar
  • 300 SIEGAL N., WOMGSURAWAT N., ARMBRECHT H. J.Parathyroid hormone stimulates dephosphorylation of the renodoxin component of the 25-hydroxyvitamin D3-1α-hydroxylase from rat renal cortex.J. Biol. Chem.26119861699817003
    PubMed | ISI | Google Scholar
  • 301 SILVER, J., and H. M. KRONENBERG. Parathyroid hormone — molecular biology and regulation. In: Principles of Bone Biology, edited by J. B. Bilezikian, L. G. Raisz, and G. A. Rodan. San Diego, CA: Academic, 1996, chapt. 24, p. 325–337.
    Google Scholar
  • 302 SILVER J., MOALLEM E., KILAV R., EPSTEIN E., SELA A., NAVEH-MANY T.New insights into the regulation of parathyroid hormone synthesis and secretion in chronic renal failure.Nephrol. Dial. Transplant.11, Suppl. 3199625
    Crossref | PubMed | ISI | Google Scholar
  • 303 SILVER, J., and T. NAVEH-MANY. Vitamin D and the parathyroid glands. In: Vitamin D, edited by D. Feldman, F. H. Glorieux, and J. W. Pike. San Diego, CA: Academic, 1997, chapt. 23, p. 353–367.
    Google Scholar
  • 304 SILVER J., NAVEH-MANY T., MAYER H., SCHMEIZER H. J., POPVTZER M. M.Regulation by vitamin D metabolites of parathyroid hormone gene transcription in vivo in the rat.J. Clin. Invest.78198612961301
    Crossref | PubMed | ISI | Google Scholar
  • 305 SIMPSON R. U., DeLUCA H. F.Characterization of a receptor-like protein for 1,25-dihydroxyvitamin D3 in rat skin.Proc. Natl. Acad. Sci. USA77198058225826
    Crossref | PubMed | ISI | Google Scholar
  • 306 SJODEN G., SMITH C., LINDGREN V., DeLUCA H. F.1-Alpha-hydroxyvitamin D2 is less toxic than 1-alpha-hydroxyvitamin D3 in the rat.Proc. Soc. Exp. Biol. Med.1781985432436
    Crossref | PubMed | ISI | Google Scholar
  • 307 SLATOPOLSKY, E., J. FINCH, C. RITTER, M. DENDA, J. MORRISSEY, A. BROWN, and H. F. DeLUCA. A new analog of calcitriol, 19-nor-1,25-dihydroxyvitamin D2 suppresses PTH secretion in uremic rats in the absence of hypercalcemia (Abstract). J. Bone Miner. Res: S167, 1995.
    Google Scholar
  • 308 SLATOPOLSKY E., WEERTS C., THIELAND J., HORST R., HARTER H., MARTIN K.Marked suppression of secondary hyperparathyroidism by I.V. administration of 1,25-dihydroxycholecalciferol in uremic patients.J. Clin. Invest.74198421362141
    Crossref | PubMed | ISI | Google Scholar
  • 309 SMITH E. L., WALWORTH N. C., HOLICK M. F.Effect of 1α,25-dihydroxyvitamin D3 on the morphologic and biochemical differentiation of cultured human epidermal keratinocytes grown in serum-free conditions.J. Invest. Dermatol.861986709714
    Crossref | PubMed | ISI | Google Scholar
  • 310 SONE T., OZONO K., PIKE J. W.A 55-kilodalton accessory factor facilitates vitamin D receptor binding.Mol. Endocrinol.5199115781586
    Crossref | PubMed | Google Scholar
  • 311 SOOY K., SCHERMERHORN T., SHARP G. W., FLEISCHER N., SURANA M., FUSCO-DEMANE D., AIRAKSINEN M., MEYER M., CHRISTAKOS S.Disruption of calbindin-D28k in the mouse (Abstract).J. Bone Miner. Res.12, Suppl.1997S123
    ISI | Google Scholar
  • 312 SPENCER R., CHARMAN M., WILSON P. W., LAWSON D. E. M.The relationship between vitamin D-stimulated calcium transport and intestinal calcium-binding protein in the chicken.Biochem. J.170197893102
    Crossref | PubMed | ISI | Google Scholar
  • 313 ST. ARNAUD R., ARABIAN A., GLORIEUX F. H.Abnormal bone development in mice deficient for the vitamin D 24-hydroxylase gene (Abstract).J. Bone Miner. Res.111996S126
    ISI | Google Scholar
  • 314 ST. ARNAUD R., ARABIAN A., TRAVERS R., GLORIEUX F. H.Partial rescue of abnormal bone formation in 24-hydroxylase knock-out mice supports a role for 24,25-(OH)2D3 in intramembranous ossification (Abstract).J. Bone Miner. Res.121997S111
    PubMed | ISI | Google Scholar
  • 315 ST. ARNAUD, R., A. ARABIAN, R. TRAVERS, and F. H. GLORIEUX. Abnormal intramembranous ossification in mice deficient for the vitamin D 24-hydroxylase gene. In: Vitamin D. Chemistry, Biology and Clinical Applications of the Steroid Hormone, edited by A. W. Norman, R. Bouillon, and M. Thomasset. Riverside: Univ. of California, 1997, p. 635–639.
    Google Scholar
  • 316 ST. ARNAUD R., MESSERLIAN S., MOIR J. M., OMDAHL J. L., GLORIEUX F. H.The 25-hydroxyvitamin D 1α-hydroxylase gene maps to the pseudovitamin D-deficiency rickets (PDDR) disease locus.J. Bone Miner. Res.12199715521559
    Crossref | PubMed | ISI | Google Scholar
  • 317 STEIN, G. S., J. B. LIAN, J. L. STEIN, A. J. VAN WIJNEN, B. FRENKEL, and M. MONTECINO. Mechanisms regulating osteoblast proliferation and differentiation. In: Bone Biology, edited by J. P. Bilezekian, G. Raisz, and G. A. Rodan. San Diego, CA: Academic, 1996, chapt. 5, p. 69–86.
    Google Scholar
  • 318 STERN P.A monolog on analogs.Calcif. Tissue Int.33198114
    Crossref | PubMed | ISI | Google Scholar
  • 319 STERN, P. H. 1,25-Dihydroxyvitamin D3 interactions with local factors in bone remodeling. In: Vitamin D, edited by D. Feldman, F. H. Glorieux, and J. W. Pike. San Diego, CA: Academic, 1997, chapt. 22, p. 341–352.
    Google Scholar
  • 320 STRUGNELL S., BYFORD V., MAKIN H. L. J., MORIARTY R. M., GILARDI R., LEVAN L. W., KNUTSON J. C., BISHOP C. W., JONES G.1,24(S)-dihydroxyvitamin D2: a biologically active product of 1-hydroxyvitamin D2 made in the human hepatoma, Hep3B.Biochem. J.3101995233241
    Crossref | PubMed | ISI | Google Scholar
  • 322 STRUGNELL S. A., DeLUCA H. F.The vitamin D receptor-structure and transcriptional activation.Proc. Soc. Exp. Biol. Med.2151997223228
    Crossref | PubMed | ISI | Google Scholar
  • 323 STRUGNELL S. A., WIEFLING B. A., DeLUCA H. F.Vitamin D receptor ligand-binding domain expressed as a fusion protein (Abstract).J. Bone Miner. Res.101995S396
    ISI | Google Scholar
  • 324 STUMPF, W. E., M. SAR, and H. F. DeLUCA. Sites of action of 1,25(OH)2 vitamin D3 identified by thaw-mount autoradiography. In: Hormonal Control of Calcium Metabolism, edited by D. V. Cohn, R. V. Talmage, and J. L. Matthews. Amsterdam: Excerpta Medica, 1981, p. 222–229.
    Google Scholar
  • 325 STUMPF W. E., SAR M., REID F. A., TANAKA Y., DeLUCA H. F.Target cells for 1,25-dihydroxyvitamin D3 in intestinal tract, stomach, kidney, skin, pituitary and parathyroid.Science206197911881190
    Crossref | PubMed | ISI | Google Scholar
  • 326 SU M.-J., BIKLE D. D., MANCIANTI M. L., PILLAI S.1,25-Dihydroxyvitamin D3 potentiates the keratinocyte response to calcium.J. Biol. Chem.26919941472314729
    PubMed | ISI | Google Scholar
  • 327 SUDA T., DeLUCA H. F., SCHNOES H. F., PONCHON G., TANAKA Y., HOLICK M. F.21,25-Dihydroxycholecalciferol: a metabolite of vitamin D3 preferentially active on bone.Biochemistry9197029172922
    Crossref | PubMed | ISI | Google Scholar
  • 328 SUDA, T., and N. TAKAHASHI. Vitamin D and osteoclastogenesis. In: Vitamin D, edited by D. Feldman, F. H. Glorieux, and J. W. Pike. San Diego, CA: Academic, 1997, chapt. 21, p. 329–340.
    Google Scholar
  • 329 SUDA T., TAKAHASHI N., MARTIN T. J.Modulation of osteoclast differentiation: update 1995.Endocr. Rev.41995266270
    Google Scholar
  • 330 TAKAHASHI Y., SUDA T., YAMADA S., TAKEYAMA H., NISHII Y.Isolation, identification and biological activity of 25-hydroxyvitamin-24-oxovitamin D3: a new metabolite of vitamin D3 generated by in vitro incubation with kidney homogenates.Biochemistry20198116811686
    Crossref | PubMed | ISI | Google Scholar
  • 331 TAKEDA T., ARAKAWA M., KUWANO R.Organization and expression of the mouse spot35/calbindin-D28k gene: identification of the vitamin D-responsive promoter region.Biochem. Biophys. Res. Commun.2041994889897
    Crossref | PubMed | ISI | Google Scholar
  • 332 TAKEYAMA K., KITANAKA S., SATO T., KOBORI M., YANAGISAWA J., KATO S.25-Hydroxyvitamin D3 1α-hydroxylase and vitamin D synthesis.Science277199718271830
    Crossref | PubMed | ISI | Google Scholar
  • 333 TALMAGE R. V.Morphological and physiological consideration in a new concept of calcium transport in bone.Am. J. Anat.1291970467476
    Crossref | PubMed | Google Scholar
  • 334 TANAKA H., ABE E., MIYAURA C., KURIBAYASHI T., KONNO K., NISHII Y., SUDA T.1α,25-Dihydroxycholecalciferol and a human myeloid leukemia cell line (HL-60). The presence of a cytosol receptor and induction of differentiation.Biochem. J.2041982713719
    Crossref | PubMed | ISI | Google Scholar
  • 335 TANAKA Y., CASTILLO L., DeLUCA H. F., IKEKAWA N.The 24-hydroxylation of 1,25-dihydroxyvitamin D3 .J. Biol. Chem.252197714211424
    PubMed | ISI | Google Scholar
  • 336 TANAKA Y., DeLUCA H. F.The control of 25-hydroxyvitamin D metabolism by inorganic phosphorus.Arch. Biochem. Biophys.1541973566574
    Crossref | PubMed | ISI | Google Scholar
  • 337 TANAKA Y., DeLUCA H. F.Stimulation of 24,25-dihydroxyvitamin D3 production by 1,25-dihydroxyvitamin D3 .Science183197411981200
    Crossref | PubMed | ISI | Google Scholar
  • 338 TANAKA Y., DeLUCA H. F., KOBAYASHI Y., TAGUCHI T., IKEKAWA N., MORISAKI M.Biological activity of 24,24-difluoro-25-hydroxyvitamin D3 . Effect of blocking of 24-hydroxylation on the functions of vitamin D.J. Biol. Chem.254197971637167
    PubMed | ISI | Google Scholar
  • 339 TANAKA Y., FRANK H., DeLUCA H. F.Biological activity of 1,25-dihydroxyvitamin D3 in the rat.Endocrinology921973417422
    Crossref | PubMed | ISI | Google Scholar
  • 340 TANAKA Y., LORENC R. S., DeLUCA H. F.The role of 1,25-dihydroxyvitamin D3 and parathyroid hormone in the regulation of chick renal 25-hydroxyvitamin D3-24-hydroxylase.Arch. Biochem. Biophys.1711975521526
    Crossref | PubMed | ISI | Google Scholar
  • 341 THAKKER R. V., FRAHER L. J., ADAMI S., KARMALI R., O'RIORDAN J. L. H.Circulating concentrations of 1,25-dihydroxyvitamin D3 in patients with primary hyperparathyroidism.Bone Miner. Res.11986137142
    Google Scholar
  • 342 THOMASSET, M. Calbindin-D9K . In: Vitamin D, edited by D. Feldman, F. H. Glorieux, and J. W. Pike. San Diego, CA: Academic, 1997, chapt. 14, p. 223–232.
    Google Scholar
  • 343 TILYARD M. W., SPEARS G. F. S., THOMSON J., DOVEY S.Treatment of post-menopausal osteoporosis with calcium.N. Engl. J. Med.3261992357362
    Crossref | PubMed | ISI | Google Scholar
  • 344 TOMON M., TENENHOUSE H. S., JONES G.Expression of 25-hydroxyvitamin D3-24-hydroxylase activity in Caco-2 cells. An in vitro model of intestinal vitamin D catabolism.Endocrinology126199028682875
    Crossref | PubMed | ISI | Google Scholar
  • 345 TORCHIA J., ROSE D. W., INOSTROZA J., KAMEL Y., WESTIN S., GLASS C. K., ROSENFELD M. G.The transcriptional co-activator p/CIP binds CBP and mediates nuclear-receptor function.Nature3871997677684
    Crossref | PubMed | ISI | Google Scholar
  • 346 TUCKER G., GAGNON R. E., HAUSSLER M. R.Vitamin D3-25-hydroxylase: tissue occurrence and lack of regulation.Arch. Biochem. Biophys.15519734757
    Crossref | PubMed | ISI | Google Scholar
  • 347 TURNER R. T., PUZAS J. E., FORTE M. D., LESTER G. E., GRAY T. K., HOWARD G. A., BAYLINK D. J.In vitro synthesis of 1α,25-dihydroxycholecalciferol and 24,25-dihydroxycholecalciferol by isolated calvarial cells.Proc. Natl. Acad. Sci. USA77198057205724
    Crossref | PubMed | ISI | Google Scholar
  • 348 UHLAND-SMITH, A. An Investigation of the Role of Vitamin D in Reproduction, Growth and Development (PhD thesis). Madison: Univ. of Wisconsin, 1992.
    Google Scholar
  • 349 UMESONO K., MURAKAMI K. K., THOMPSON C. C., EVANS R. M.Direct repeats as selective response elements for the thyroid hormone, retinoic acid, and vitamin D3 receptors.Cell65199112551266
    Crossref | PubMed | ISI | Google Scholar
  • 350 USKOKOVIC, M. R., P. S. MANCHAND, S. PELEG, and A. W. NORMAN. Synthesis and preliminary evaluation of the biological properties of a 1,25-dihydroxyvitamin D3 analog with two side-chains. In: Vitamin D. Chemistry, Biology and Clinical Applications of the Steroid Hormone, edited by A. W. Norman, R. Bouillon, and M. Thomasset. Univ. of California, 1997, p. 19–21.
    Google Scholar
  • 351 USUI E., NOSHIRO M., OKUDA K.Molecular cloning of cDNA for vitamin D3 25-hydroxylase from rat liver mitochondria.Fed. Eur. Biol. Sci. Lett.2621990135138
    Crossref | PubMed | ISI | Google Scholar
  • 352 VAN DEM BEMD G. C., POLS H. A. P., BIRKENHAGER J. C., VAN LEEUWEN J. P.Conformational change and enhanced stabilization of the vitamin D receptor by the 1,25-dihydroxy vitamin D3 analog, KH1060.Proc. Natl. Acad. Sci. USA9319961068510690
    Crossref | PubMed | ISI | Google Scholar
  • 353 VAN DEN BEMD G. J. C., DILWORTH F. J., WILLIAMS G. R., POLS H. A. P., JONES G., VAN LEEUWEN J. P. T. M.Metabolites of the 1,25-dihydroxyvitamin D3 analog KH1060 induce similar conformational changes as KH1060 itself (Abstract).J. Bone Miner. Res.121997S453
    ISI | Google Scholar
  • 354 VLAHECEVIC Z. R., JAIRATH S. K., HEUMAN D. M., STRAVITZ R. T., HYLEMON P. B., AVADHANI N. G., PANDAK W. M.Transcriptional regulation of hepatic sterol 27-hydroxylase by bile acids.Am. J. Physiol.270(Gastrintest. Liver Physiol. 33)1996G646G652
    Google Scholar
  • 355 VOM BAUR, E., C. ZECHEL, D. HEERY, M. J. S. HEINE, J. M. GARNIER, V. VIVAT, B. LE DOUARIN, H. GRONEMEYER, P. CHAMBON, and R. LOSSON. Differential ligand-dependent interactions between the AF-2 activating domain of nuclear receptors and the putative transcriptional intermediary factors mSUG1 and TIF1. EMBO J. 15: 110–124, 1996.
    Google Scholar
  • 356 WAGNER R. L., APRILETTI J. W., McGRATH M. E., WEST B. L., BAXTER J. D., FLETTERICK R. J.A structural role for hormone in the thyroid hormone receptor.Nature3781995690697
    Crossref | PubMed | ISI | Google Scholar
  • 357 WASSERMAN, R. H., and J. J. FEHER. Vitamin D-dependent calcium-binding proteins. In: Calcium Binding Proteins and Calcium Function, edited by R. H. Wasserman, R. A. Corradino, E. Carafoli, R. H. Kretsinger, D. H. MacLennan, and S. L. Siegel. New York: Elsevier/North-Holland, 1977, p. 292–302.
    Google Scholar
  • 358 WHEELER D. G., HORSFORD J., MICHALAK M., WHITE J. H., HENDY G. N.Calreticulin inhibits vitamin D3 signal transduction.Nucleic Acids Res.23199532683274
    Crossref | PubMed | ISI | Google Scholar
  • 359 WHITFIELD G. K., HSIEH J. C., NAKAJIMA S., MacDONALD P. N., THOMPSON P. D., JURUTKA P. W., HAUSSLER C. A., HAUSSLER M. R.A highly conserved region in the hormone-binding domain of the human vitamin D receptor contains residues vital for heterodimerization with retinoid X receptor and for transcriptional activation.Mol. Endocrinol.9199511661179
    PubMed | Google Scholar
  • 360 WILLIAMS G. R., BLAND R., SHEPPARD M. C.Retinoids modify regulation of endogenous gene expression by vitamin D3 and thyroid hormone in three osteosarcoma cell lines.Endocrinology136199543044314
    Crossref | PubMed | ISI | Google Scholar
  • 361 WURTZ, J.-M., B. GUILLOT, and D. MORAS. 3-D model of the ligand-binding domain of the vitamin D receptor based on the crystal structure of holo-RAR. In: Vitamin D. Chemistry, Biology and Clinical Applications of the Steroid Hormone, edited by A. W. Norman, R. Bouillon, and M. Thomasset. Riverside: Univ. of California, 1997, p. 165–172.
    Google Scholar
  • 362 YAMAMOTO M., KAWANOKE Y., TAKAHASHI H., SHIMAZAWA E., KIMURA S., OGATA E.Vitamin D deficiency and renal calcium transport in the rat.J. Clin. Invest.741984507513
    Crossref | PubMed | ISI | Google Scholar
  • 363 YANG S., SMITH C., DeLUCA H. F.1α,25-Dihydroxyvitamin D3 and 19-nor-1α,25-dihydroxyvitamin D2 suppress immunoglobulin production and thymic lymphocyte proliferation in vivo.Biochim. Biophys. Acta11581993269286
    Google Scholar
  • 364 YANG S., SMITH C., PRAHL J. M., DeLUCA H. F.Vitamin D deficiency suppresses cell-mediated immunity in vivo.Arch. Biochem. Biophys.303198398106
    Crossref | ISI | Google Scholar
  • 365 YOSHIZAWA T., HANDA Y., UEMATSU Y., TAKEDA S., SEKINE K., YOSHIHARA Y., KAWAKAMI T., ARIOKA K., SATO H., UCHIYAMA Y.Mice lacking the vitamin D receptor exhibit impaired bone formation, uterine hypoplasia and growth retardation after weaning.Nature Genet.161997391396
    Crossref | PubMed | ISI | Google Scholar
  • 366 ZECHEL C., SHEN X. Q., CHEN J. Y., CHEN Z. P., CHAMBON P., GRONEMEYER H.The dimerization interfaces formed between the DNA binding domains of RXR, RAR and TR determine the binding specificity and polarity of the full-length receptors to direct repeats.EMBO J.13199414251433
    Crossref | PubMed | ISI | Google Scholar
  • 367 ZHOU A., ELGORT M. G., ALLEGRETO E. A.Retinoid X receptor (RXR) ligands activate the human 25-hydroxyvitamin D3-24-hydroxylase promoter via RXR heterodimer binding to two vitamin D-response elements and elicit additive effects with 1,25-dihydroxyvitamin D3 .J. Biol. Chem.27219971902719034
    Crossref | PubMed | ISI | Google Scholar
  • 368 ZIEROLD C., DARWISH H. M., DeLUCA H. F.Two vitamin D response elements function in the rat 1,25-dihydroxyvitamin D 24-hydroxylase promoter.J. Biol. Chem.270199516751678
    Crossref | PubMed | ISI | Google Scholar