© 2014 John Wiley & Sons A/S. Published by John Wiley & Sons Ltd doi:10.1111/tra.12177

Review

Binding Domain-Driven Intracellular Trafficking of Sterols for Synthesis of Steroid Hormones, Bile Acids and Oxysterols Andrew Midzak1∗ and Vassilios Papadopoulos1,2,3,4∗ 1 Research

Institute of the McGill University Health Centre, McGill University, Montreal, Quebec, Canada of Medicine, McGill University, Montreal, Quebec, Canada 3 Department of Biochemistry, McGill University, Montreal, Quebec, Canada 4 Department of Pharmacology and Therapeutics, McGill University, Montreal, Quebec, Canada 2 Department



Corresponding authors: Andrew Midzak, [email protected] and Vassilios Papadopoulos, [email protected]

Abstract Steroid hormones, bioactive oxysterols and bile acids are all derived from

cholesterol metabolism to steroid hormones and bile acid, relating it to

the biological metabolism of lipid cholesterol. The enzymatic pathways

both lipid- and protein-based mechanisms facilitating intracellular and

generating these compounds have been an area of intense research for

intraorganellar cholesterol movement and delivery to these pathways. In

almost a century, as cholesterol and its metabolites have substantial

particular, we examine evidence for the involvement of specific protein

impacts on human health. Owing to its high degree of hydrophobicity

domains involved in cholesterol binding, which impact cholesterol move-

and the chemical properties that it confers to biological membranes, the

ment and metabolism in steroidogenesis and bile acid synthesis. A better

distribution of cholesterol in cells is tightly controlled, with subcellular

understanding of the physical mechanisms by which these protein- and

organelles exhibiting highly divergent levels of cholesterol. The manners

lipid-based systems function is of fundamental importance to under-

in which cells maintain such sterol distributions are of great interest in

standing physiological homeostasis and its perturbation.

the study of steroid and bile acid synthesis, as limiting cholesterol sub-

Keywords cholesterol, mitochondria, protein motifs, steroidogenesis,

strate to the enzymatic pathways is the principal mechanism by which

steroids, sterol trafficking

production of steroids and bile acids is regulated. The mechanisms by

Received 25 April 2014, revised and accepted for publication 28 May

which cholesterol moves within cells, however, remain poorly under-

2014, uncorrected manuscript published online 30 May 2014, pub-

stood. In this review, we examine the subcellular machinery involved in

lished online 7 July 2014

The lipid cholesterol has generated some of the greatest research interest regarding the chemicals that make up life. Providing profoundly critical physical properties as a component of the membranes that separate the biotic from the abiotic, cholesterol exerts significant physiological effects in its own right. Moreover, cholesterol is enzymatically metabolized in cells to structurally diverse families of steroids, oxysterols and bile acids, each of which have specific and highly potent signaling capabilities. Our

understanding of the ways through which cells handle cholesterol and respond to cholesterol metabolite signaling has expanded enormously over the past half century. However, the manner in which cells synthesize steroids and other cholesterol metabolites, despite the progress made, has yet to fully integrate the advances of cholesterol biophysics and cellular biology into its theoretical framework. These issues have captured the interest of researchers in the field (1), and in this review, we look to extend on this www.traffic.dk 895

Midzak and Papadopoulos

Figure 1: Cholesterol and its oxidized metabolites. Cholesterol is a cycloperhydropentanophenanthrene-ringed structure that can be chemically modified to yield molecules possessing potent biological activity. Shown in the center is cholesterol, with the carbon atoms indicated by numbers and the rings designated by letters. In vertebrates, cholesterol can be oxidized to form oxysterols by enzymatic reactions or by auto-oxidation in all cells (e.g. 27-hydroxycholesterol and 7α-hydroxycholesterol), or they can be oxidized to bile acids in hepatocytes (e.g. cholic acid and chenodeoxycholate, right ), or oxidized to steroid hormones in steroidogenic cells (e.g. pregnenolone and cortisol, left ). In arthropods, differential enzymatic systems lead to the metabolism of cholesterol to ecdysteroids (e.g. 20-hydroxyecdysone, top). Note that only steroids lack the aliphatic tail of cholesterol; in other metabolites, increased aqueous solubility is achieved by oxidation of the aliphatic tail. Not shown for clarity are additional chemical modifications of cholesterol with biological relevance, such as lipidation and sulfation.

previous work and offer our own critical assessment of the relationship between sterol movement and sterol metabolism. We shall begin with a description of steroidogenesis and the role of mitochondrial cholesterol transport in this process. We then step back and describe the current understanding of general cellular cholesterol transport, placing emphasis on the biophysical behavior of cholesterol in lipid and aqueous media. Finally, we describe the proposed role of lipid-binding domains in steroidogenesis and cellular cholesterol traffic, highlighting both the strengths and weaknesses of our current knowledge of these processes. 896

Steroidogenesis, Oxysterol and Bile Acid Synthesis, and Mitochondrial Cholesterol Metabolism All vertebrate steroids are metabolic products of cholesterol. The structural characteristics of cholesterol and its oxygenated products confer varying chemical properties to the molecules (Figure 1). Cholesterol is characterized by (i) a fused cyclopentanophenanthrene four-ring backbone with a single double bond between carbons 5′ and 6′ ; (ii) a single stereospecific hydroxylation at the 3′ carbon (3β-OH); and (iii) a saturated carbon tail attached to Traffic 2014; 15: 895–914

Protein Motifs in Sterol Traffic and Metabolism

carbon 17′ of the D ring. In contrast, vertebrate steroid hormones (2,3) and bile acids (4) are oxygenated at multiple positions on the cyclopentanophenanthrene backbone. Steroids differ from bile acids by retaining a double bond at either the 4′ or 5′ carbons and, most importantly, steroid hormones are oxygenated at the 17′ carbon rather than possessing an extended aliphatic tail. This last characteristic greatly increases steroid hydrophilicity and makes this class of lipids structurally distinct from the oxysterols and bile acids, which retain the aliphatic tail of cholesterol (though it itself is oxygenated). This class of lipids is also distinct from the compounds that have been referred to as steroids in the lower metazoans [e.g. ecdysteroids in insects (5)], but that more closely resemble oxysterols and bile as they retain the sterol carbon tail. While oxysterols and bile acids are still functionally grouped by their chemical structure, the contemporary definition of each class of steroids is based on the nuclear receptor(s) to which it binds rather than on its chemical structure (6,7). The structures of steroids were biochemically determined in the early 20th century (8) and with extensive work conducted to establish precursor/product relationships in mammalian tissues. This work led to the general mapping of the chemical metabolism of sterols and steroids, and it provided the framework for establishing the pathways of steroidogenesis; however, it was not until the 1980s, with the isolation and cloning of steroidogenic enzymes, that the steroidogenic pathways were defined. This work has been recently, and authoritatively, reviewed by Miller and Auchus (3). Briefly, steroidogenic enzymes fall into two broad families: the cytochrome P450 (CYP) and hydroxysteroid dehydrogenase/ketosteroid reductase (HSD/KSR) enzymes (2,3). CYPs are a group of oxidative enzymes, all of which have about 500 amino acids and contain a prosthetic heme group, and they utilize electrons from NAD(P)H for enzymatic activity (9). Humans contain 57 CYPs, which can be further subclassified into mitochondrial and microsomal CYPs, each of which utilizes different enzymatic sources of electrons. Mitochondrial enzymes utilize a ferrodoxin reductase–ferrodoxin electron transport chain, whereas microsomal enzymes utilize a single P450 oxidoreductase (10). Six P450 enzymes are involved in steroidogenesis (Figure 2): the mitochondrial CYP11A1, Traffic 2014; 15: 895–914

CYP11B1 and CYP11B2, and the endoplasmic reticular (ER) CYP17A1, CYP19A1 and CYP21A1. The HSDs have molecular masses of approximately 40 kDa, do not have prosthetic heme groups and require nicotinamide adenine dinucleotide (phosphates) (NADH/NAD+ or NADPH/NADP+ ) as cofactors to either reduce or oxidize a steroid by two electrons via a hydride transfer mechanism (11). Most HSD reactions involve the conversion of an alcohol to a ketone (or vice versa), though in the case of 3β-HSDs, dehydrogenation is accompanied by the isomerization of the adjacent carbon–carbon double bond from the Δ5 position to the Δ4 position of the steroid backbone (12). Although steroidogenic enzymes are broadly distributed throughout the body, allowing multiple cells and tissues to locally metabolize steroids for particular needs (13), steroidogenic cells are able to synthesize steroids de novo and are defined by their expression of the cytochrome P450 enzyme, CYP11A1. Such steroidogenic cells include both the classical hormone-dependent steroidogenic cells of the adrenal cortex and the male and female gonad- and hormone-independent steroidogenic cells of the female placenta, as well as non-classical steroidogenic tissue, such as steroidogenic cells in the central nervous system (3,14). CYP11A1 is a member of the mitochondrial clade of cytochrome P450 enzymes, which contains six additional CYPs in humans (CYP11B1, CYP11B2, CYP24 and CYP27A1, as well as CYP27B1 and CYP27C1). This clade differs from microsomal cytochrome P450s in their cellular localization – mitochondrial CYPs are peripherally attached to the matrix side of the inner mitochondrial membrane, whereas microsomal CYPs are attached to the lumen face of the ER membranes – and their source of catalytic electrons – mitochondrial CYPs utilize an electron transport chain consisting of ferrodoxin reductase and ferrodoxin, in contrast to the single P450 oxidoreductase for microsomal CYPs. Interestingly, the mitochondrial CYP clade appears to share a preference for sterol-like substrates, as CYP11A1 and CYP27 hydroxylate the aliphatic tail of cholesterol (15,16), CYP11B1 and CYP11B2 hydroxylate the 11′ and 18′ carbons of steroid substrates (17) and CYP24 hydroxylates the secosteroid 1,25-dihydroxyvitamin D3 (18,19). This substrate preference may reflect a deep evolutionary relationship, 897

Midzak and Papadopoulos

Figure 2: Major steroidogenic and bile acid synthetic pathways. A simplified schematic diagram of the transformation of cholesterol to bioactive steroids and bile acids is shown; chemical intermediates have been removed for clarity. The mitochondria are shown centrally, as this is the rate-limiting site of cholesterol metabolism to steroid hormones, and a key site in the alternative pathway of bile acid synthesis. Human nomenclature of enzymes is used. Following scission of the aliphatic tail of cholesterol by the cytochrome P450 CYP11A1, the steroid pregnenolone is metabolized to further steroid products, which are dependent upon the expression of additional steroidogenic enzymes. Note that the majority of these enzymes are localized in the ER, though CYP11B1 and CYP11B2, facilitating the final conversion of glucocorticoid and mineralocorticoid steroid hormones, respectively, are mitochondrial enzymes, which necessitate the return of steroid metabolites. The intracellular trafficking of steroids after the synthesis of pregnenolone is believed to be due to diffusion, as steroids have increased hydrophobicity and no steroid transporters have been identified. In bile acid synthesis, the rate-limiting step in the majority of bile acid synthesis (∼75–90%) is the 7α-hydroxylation of the sterol B ring by cytochrome P450 CYP7A1 localized in the ER of hepatocytes; the 27-hydroxylation of cholesterol by mitochondrial CYP27A1 enzymes throughout the body contributes to circulating levels of 27-hydroxycholesterol, which can subsequently be 7α-hydroxylated by hepatic and extrahepatic CYP7B1. as the mitochondrial CYP family is highly conserved among metazoans (20,21), though any non-mammalian substrates are unknown at this time. As noted above, CYP11A1 is absolutely essential for the synthesis of all vertebrate steroids, and the proposed reaction mechanism of CYP11A1 involves three sequential modifications of cholesterol. First, cholesterol is hydroxylated at carbon 22; second, cholesterol is hydroxylated at carbon 20; and finally, oxidative scission of the C20–C22 bond of 20(3),22R-dihydroxycholesterol yields pregnenolone and isocaproaldehyde (22). These hypotheses have been supported by the crystal structures of bovine and human CYP11A1 (23,24), indicating that the prosthetic heme is in close proximity to the 20′ and 22′ carbons of cholesterol. 898

The transcriptional levels of CYP11A1 determine the cellular steroidogenic capacity of a cell (25), and as the expression of CYP11A1 is slow to rise and fall, it constitutes a key chronic mechanism of steroidogenic regulation. In the classical steroidogenic cells of the adrenal cortex and gonads, circulating pituitary hormones stimulate intracellular signaling cascades, leading to the elevation and maintenance of CYP11A1 expression (26,27) – processes that require the transcription factor, steroidogenic factor 1 (SF-1) (28). CYP11A1 regulation in other tissues is not as well characterized, but it appears to be constitutively maintained at physiological levels by different mechanisms (29). All steroidogenic cells metabolize cholesterol to steroids basally, and CYP11A1 expression dictates the level of basal Traffic 2014; 15: 895–914

Protein Motifs in Sterol Traffic and Metabolism

steroidogenesis. The mechanisms dictating basal steroidogenesis are not clear at this time, though they remain an important area for research, as this form of steroidogenesis is performed by the placenta and steroidogenic cells in the central nervous system, which are not stimulated by known hormones and factors. Although basal steroidogenesis may be a consequence of low levels of sterol diffusion to CYP11A1 in the mitochondrial matrix, sterol-binding proteins discussed below appear critical in facilitating the process, suggesting a role for protein machinery in the process. Basal steroidogenesis is insufficient to maintain the necessary levels of steroid hormones; however, the pituitary hormones that stimulate the adrenal cortex and gonads not only promote chronic expression of steroidogenic enzymes but also acutely stimulate a rapid and dramatic rise in steroid hormone synthesis. This acute phase of steroidogenesis is characterized by significant morphological changes (30) and intramitochondrial cholesterol transport and delivery to CYP11A1 – a process that is rate limiting in the synthesis of steroids (31,32). Given that intramitochondrial cholesterol transport constitutes the primary locus of control for hormone-stimulated steroidogenesis, this acute step has become a key area of research into the synthesis of steroids as well as a proxy for steroidogenesis in general (25). In the rest of this review, we will examine our current knowledge of how sterols move between membranes and then proceed to a discussion of the protein motifs that help facilitate cholesterol movement.

Biophysical Concepts and Sterol Chemical Activity Although the theoretical considerations of mitochondrial cholesterol transport hinge on the large difference in the inner and outer mitochondrial membrane cholesterol levels, it is important to note that this difference is only part of the larger compartmentalization of cholesterol that cells undertake as part of their homeostatic physiology. Cholesterol concentrations are quite divergent throughout the cellular organelles (33,34) and the following sections describe both this distribution and the models that describe lipid’s contribution to this distribution, before transitioning to a discussion on the models that underline the contribution of protein. Traffic 2014; 15: 895–914

The plasma membrane contains the highest levels of cholesterol in the cell, followed by compartments intimately associated with the plasma membrane, such as the Golgi apparatus and the endosomal recycling compartment (ERC). There appears to be a gradient of cholesterol along the secretory pathway of cells, with the highest levels found at the plasma membrane, intermediate levels in the Golgi apparatus and low levels in the ER (35,36). For this reason, the delivery of cholesterol via vesicular transport, utilizing cellular secretory machinery, is proposed to play a major role in cellular cholesterol homeostasis, though mechanisms exist to transfer cholesterol to organelles outside of the secretory pathway, such as the mitochondria (37). The fact that there are low levels of cholesterol in the ER is interesting, as the ER is the site of cholesterol synthesis, and cholesterol that is added to the cell, either directly or in endocytosed lipoproteins, is quickly acetylated by the ER-localized acyl-CoA cholesterol acyltransferase (ACAT) enzyme and delivered to lipid droplets (themselves considered to be lipid-rich buds of the ER) (38). Collectively, these observations indicate that the ER serves as a central cholesterol distribution center for the cell, though it retains low levels of cholesterol. These low levels of cholesterol make the ER an excellent sterol-sensing organelle, and it has been shown that the transcriptional machinery regulating sterol synthesis and metabolism, governed by the transcription factor SREBP, is especially sensitive to sterol concentrations (36,39). This genetic machinery, operating on the time scale of hours to days, does not, however, explain how cells rapidly control cellular sterol distribution within seconds to minutes (40) – mechanisms that we are only recently beginning to understand. Much of the work on the cellular distribution and movement of cholesterol has been performed in non-steroidogenic mammalian and yeast cells. With regard to steroidogenic cells, the source of cholesterol transported to the mitochondria has long been considered to be the lipid droplets, beginning with ultrastructural work indicating that lipid droplet volume decreased upon hormonal stimulation of the adrenocortical cells (41). More recently, non-perturbational imaging of mouse Y-1 adrenocortical tumor cells demonstrated cytoskeletal trafficking of lipid droplets to the mitochondria (42), suggesting the direct delivery of lipid droplet cholesterol for steroidogenesis. Such a system was supported 899

Midzak and Papadopoulos

by knockout of the intermediate filament vimentin, in which reduction of lipid droplet cholesterol delivery along the cytoskeleton disturbed adrenal and ovarian, but not testicular, steroidogenesis (43). Interestingly, work by Freeman utilizing a Leydig tumor cell model suggested that steroidogenic cells were opportunistic, deriving cholesterol both from stores in the plasma membrane as well as diverting newly synthesized cholesterol from the ER (44–46). The precise contribution of the subcellular compartments to steroidogenesis remains to be determined, but it is plausible that steroidogenic cells obtain cholesterol from any source they can, with the actively metabolizing steroidogenic mitochondria serving as a cholesterol ‘drain’, drawing cholesterol from multiple sources along the sterol trafficking pathways (Figure 3). A key theoretical model in understanding sterol distribution within lipid membranes and organelles is the concept of sterol chemical ‘activity’, which refers to the sterol capacity to carry out chemical or physical processes (47). Molecules in a membrane can be viewed as moving laterally within the membrane, traversing the membrane or escaping the membrane (Figure 4A) (47,48). The last mode of molecular movement corresponds to sterol chemical activity, making the sterol accessible to other molecular agents outside of the membrane (49). This accessibility was demonstrated in cellular systems utilizing the cholesterol-modifying cholesterol oxidase enzyme (50) and the cholesterol cyclodextrin polysaccharide (MCD) (51,52), which showed that increasing the local concentrations of cholesterol in membranes promoted the escape of cholesterol from the membranes and increased cholesterol’s association with aqueous protein acceptors (Figure 4B). Two biophysical models of cholesterol interactions with phospholipids provide useful frameworks through which to understand cellular cholesterol activity. Each of these models addresses separate, but not exclusive, aspects of cholesterol interaction with neighboring phospholipids. The condensed complex model of cholesterol–phospholipid interaction concentrates upon cholesterol’s interaction with the fatty acyl chains of phospholipids (53,54). In this model, stoichiometric hydrophobic associations between cholesterol and phospholipids with long saturated acyl chains form compact 900

Figure 3: Cholesterol trafficking pathways in mammalian cells. Cholesterol transport in mammalian cells can proceed along ‘inside-out’ or ‘outside-in’ traffic patterns, based on cellular needs. Sterol can be transported from its site of synthesis to the plasma membrane along the secretory pathway, involving the Golgi apparatus, or directly to the plasma membrane, presumably involving sterol carrier proteins, with the latter comprising the bulk of sterol transfer. Cholesterol from the plasma membrane, or absorbed from circulating lipoproteins, proceeds from the PM or the ERC to the ER, where the cholesterol is rapidly acetylated by the resident ER protein, ACAT. Acetylated cholesterol is incorporated in lipid droplets where it is stored for cellular needs. Cholesterol concentrations exceeding the carrying capacity of the ER activate the SREBP cholesterol genetic regulatory machinery, enacting the cellular homeostatic response to lower cholesterol levels. Mitochondria are commonly cholesterol-poor organelles, primarily due to extremely low inner mitochondrial membrane levels. Hormonal activation of steroidogenesis promotes the intramitochondrial transfer of cholesterol from the outer mitochondrial membrane to the inner mitochondrial membrane, and it appears to traffic cholesterol along multiple cellular pathways to maintain steroidogenesis. Abbreviations: PM, plasma membrane; ERC, endosomal recycling compartment; Golgi, Golgi apparatus; ER, endoplasmic reticulum; LD, lipid droplet; Mito, mitochondria. complexes of low free energy and small lateral areas (Figure 4C). Conversely, cholesterol associates less well with unsaturated and short-chain fatty acyl chains, minimizing the formation of condensed complexes (55). Cholesterol within these complexes has lower accessibility and activity than cholesterol outside of these complexes and, consequently, cholesterol activity within, and escape Traffic 2014; 15: 895–914

Protein Motifs in Sterol Traffic and Metabolism

Figure 4: Schematics of cholesterol behavior within the membranes and their interactions with phospholipids and proteins. A) Schematic depiction of cholesterol lipid’s lateral and transverse movement within a phospholipid membrane. Cholesterol has a high degree of mobility within the lipid environment of the membrane, with lateral diffusion times on the millisecond timescale observed in model and cellular membranes. Transverse movement within the membrane (flip-flop) is also rapid, with theoretical and experimental measurements on the millisecond to minute timescale. However, transverse movement out of the membrane (escape activity) is quite slow, as it takes place on the timescale of hours to days in the absence of proximal acceptors, whether they are proteins or other membranes. B) Stepwise schematic of LBP-mediated sterol transfer. Cholesterol in the membranes (yellow rectangles; red circle denotes 3′ OH group) is recognized (1) by a soluble LBP, which binds to the sterol and (2) extracts it from the donor membrane. The association of the LBP with an acceptor membrane (3) results in the release of the sterol, which begins the cycle anew (4). C) Simplified illustration of a condensed complex model of the cholesterol–phospholipid association and sterol escape activity. At low sterol-to-phospholipid ratios, cholesterol is able to associate stoichiometrically with the acyl chains of the surrounding phospholipids. When local cholesterol concentration exceeds the capacity of the membrane to form stoichiometric complexes, the excess cholesterol looks to move (either laterally or transversely), increasing its activity and possible association with molecules outside of its membrane. D) Simplified illustration of the umbrella model of the cholesterol–phospholipid association in sterol escape activity. Phospholipids with a high degree of acyl chain unsaturation (left , unsaturated acyl chains in orange) will sterically clash with cholesterol in the membranes, which is in contrast to phospholipids with a high degree of acyl chain saturation (right , saturated acyl chains in purple). Moreover, phospholipids with small headgroups (left ) will insufficiently shield the hydrophobic cholesterol molecule from the aqueous media, making the escape of the cholesterol molecule to a more hydrophobic environment energetically favorable. In contrast, phospholipids with large headgroups (right ) shield the smaller cholesterol molecule from the aqueous surroundings, reducing the degree of accessibility and activity of the cholesterol. Traffic 2014; 15: 895–914

901

Midzak and Papadopoulos

from, the membranes is greater when its concentration exceeds the stoichiometric capacity of the membrane (56). This model well addresses the high sterol distribution in the plasma membrane, which contains high levels of phospholipids with long saturated acyl chains (e.g. sphingomyelin), as well as the low sterol distribution in the ER and mitochondria, which contains proportionally greater levels of unsaturated phospholipids (57). The condensed complex model also fits well with cellular work that shows that amphipathic molecules capable of breaking cholesterol–phospholipid interactions sharply facilitate cellular cholesterol movement (58,59). Cholesterol–phospholipid interactions can be influenced not only by the fatty acyl tails of phospholipids but also by the headgroups at the membrane interface. This aspect of sterol chemistry is addressed by the umbrella model (60). In this model, the hydrophilic 3′ hydroxyl group of cholesterol is insufficient for shielding the hydrophobic bulk of the molecule from the aqueous media. Consequently, cholesterol ‘snorkels’ in the phospholipid membrane shielded from water by the amphipathic headgroups of the surrounding phospholipids (Figure 4D). The model predicts that phospholipids with larger headgroups, such as sphingomyelin and phosphatidylcholine, will better shield and reduce the activity of cholesterol than phospholipids with smaller headgroups, such as phosphatidylethanolamine and cardiolipin. This model also addresses acyl chain dynamics, as unsaturated acyl tails sterically clash with cholesterol, whereas saturated chains fit well with the rigid ringed structure of cholesterol. The umbrella model has received theoretical and experimental support in model membranes (61), and has interesting implications for steroidogenesis. The inner mitochondrial membranes in rat liver (62,63) and adrenocortical cells (64,65) contain proportionally greater levels of small-headed than large-headed phospholipids, which is consistent with the low levels of cholesterol observed. However, the ratio of large-headed to small-headed phospholipids in the outer mitochondrial membrane approaches that of the plasma membrane and the ERC, possibly contributing to the lack of equilibration noted between the inner and outer mitochondrial membranes. Moreover, work with CYP11A1 in reconstituted liposomal systems supports the importance of the enzyme’s phospholipid microenvironment and cholesterol’s escape activity 902

in steroidogenesis, as CYP11A1 itself could be viewed as a sterol-acceptor protein akin to cholesterol oxidase (66). A series of articles from Kimura and coworkers well delineated the positive correlation between increasing phospholipid acyl chain unsaturation and increasing CYP11A1 activity (67–70). This work also supported the importance of the effects that phospholipid headgroup exerted in steroidogenesis, as cardiolipin with its minimal headgroup was highly effective at stimulating CYP11A1 activity (69). Our discussion of sterol movement thus far has remained focused upon the lipid membrane environment. However, while cholesterol moves rapidly within lipid environments, its movement in aqueous media is energetically highly unfavorable, and the transfer between lipid membranes across the aqueous space is very slow (71), especially in comparison to the rate at which cholesterol is transported intramitochondrially in steroidogenic cells. For this reason, candidate proteins have been proposed to facilitate intermembrane cholesterol transport (72), extracting cholesterol from donor membranes (much akin to cyclodextrin) and delivering it to acceptor membranes (Figure 4B). Consequently, sterol activity plays a strong theoretical role in the discussion of lipid-binding and transfer proteins. Many of these proteins, which include members of the steroidogenic acute regulatory protein (STAR)-related lipid transfer (START) domain (73) and oxysterol-binding protein-related domain (ORD) (74), among others, contain hydrophobic cavities that facilitate sterol binding. In addition to the large hydrophobic cavities of these proposed lipid transport proteins, smaller sterol-binding domains have been identified which, instead of binding and transporting sterol, may be involved in binding and targeting sterol for cellular processes. In the following sections, we discuss work investigating the mechanisms of these sterol-binding domains in the context of steroid biosynthesis.

The STAR and the START Domain In humans, the START domain family of proteins contains 16 members, of which the founding member, STAR, was initially discovered in the investigation of the acute mitochondrial production of steroids (75). The remaining 15 members of the START family were identified in the human genome by homologous sequence alignment Traffic 2014; 15: 895–914

Protein Motifs in Sterol Traffic and Metabolism

(76), and the proteins can be mapped to several distinct subfamilies – many with a characterized ligand, but several with no known ligand. The protein family has also been shown to be quite ancient, with members distributed across multiple phyla of life (77), although the ligands for these proteins, if any, are largely unknown. The closest mammalian homologs to STAR – STARD3, STARD4, STARD5 and STARD6 – constitute a proposed family of sterol-binding proteins, and the majority of research into this family has aimed at validating this hypothesis. As noted above, STAR was originally identified in the search for a protein factor that mediated the acute mitochondrial transport of cholesterol and steroidogenesis in hormone-responsive cells. Orme-Johnson and coworkers (75,78), using two-dimensional polyacrylamide gel electrophoresis, identified a mitochondrial phosphoprotein that was rapidly upregulated by hormonal stimulation in steroidogenic cells. The subsequent isolation and characterization of this protein, named STAR, revealed a 285-amino acid protein characterized by an approximately 65-amino acid mitochondrial targeting sequence attached to an approximately 220-amino acid START domain (79). Research into the human condition of congenital adrenal lipoid hyperplasia (CAH), characterized by cellular lipid accumulation and the inability to synthesize steroids, with early death due to a lack of glucocorticoids (80), revealed that these patients contained missense and deletion mutations in STAR (81–83). The subsequent knockout of Star in mice confirmed that this protein was essential for adrenal steroidogenesis (84). It is interesting to note that STAR expression appears limited to classical hormone-dependent steroidogenic tissues, and does not seem to contribute to the basal steroid production of these tissues or the steroidogenesis of hormone-independent and non-classical tissues (1), suggesting that it is essential for stimulation of high-flux mitochondrial cholesterol movement, a critical, albeit specialized, form of cholesterol movement. Because STAR was identified as a mitochondrial protein, original models of its action involved its import into the mitochondria, possibly forming contact sites through which cholesterol could traverse from the outer to the inner mitochondrial membrane (85). However, it was found that the expression of STAR lacking the mitochondrial targeting Traffic 2014; 15: 895–914

sequence in steroidogenic cells stimulated steroidogenesis to the same levels as the expression of full-length STAR. This suggests that the import of STAR is not necessary for its action, and that the protein functions on the outside of the mitochondria (86). This hypothesis was confirmed by Miller and coworkers who showed that tethering of STAR to the outer mitochondrial membrane increased its activity, while tethering it to the mitochondrial intermembrane space or the matrix abolished its activity, suggesting that the mitochondrial import of STAR serves as an ‘off’ switch for its steroidogenic activity (87). That the mitochondrial targeting sequence of STAR is unnecessary is questioned, however, by mouse models in which full-length STAR has been replaced by STAR that lacks the mitochondrial targeting sequence, and which stochastically exhibits the CAH phenotype of an inability to synthesize steroids, indicating a more complex scenario (88). The fact that STAR facilitated mitochondrial cholesterol transport led to the hypothesis that STAR directly facilitates cholesterol transport. The crystal structure of the homologous MLN64/STARD3 was shown to consist of a helix–fold–helix motif forming a hydrophobic cavity capable of fitting a cholesterol molecule, and both STAR and MLN64/STARD3 were shown to physically associate with cholesterol (89). This initial START domain structure carries over to subsequently obtained crystal structures of the cholesterol-associated STAR, STARD4 and STARD5 proteins (90,91). Molecular modeling and dynamics studies have been used to support a model of STAR as a sterol transporter, with its C-terminal α-helix serving as a lid over its hydrophobic cavity (92,93). The model of STAR as a cholesterol transporter was initially supported by biochemical work showing the transfer of radiolabeled cholesterol and fluorescent sterol from donor to acceptor membranes by wild-type, but not mutated, STAR (94,95). Furthermore, reconstitution of the CYP11A1 system in a minimal liposomal system revealed that STAR was able to facilitate the transfer of cholesterol from donor vesicles to acceptor vesicles containing CYP11A1, where it was metabolized to pregnenolone (96). The finding that STAR mutants that were incapable of stimulating steroid production also possessed diminished capacity to facilitate sterol transfer in vitro further supported a model for STAR-mediated mitochondrial sterol transport. However, the stoichiometry of STAR 903

Midzak and Papadopoulos

sterol transport in vitro, calculated to be approximately two sterol molecules per minute (96), was several orders of magnitude less than that observed in a cellular system, which was calculated to be ∼400 sterol molecules per minute (97). In addition, the inclusion of cardiolipin in donor vesicles increased the escape activity of cholesterol, which is consistent with the umbrella model of membrane sterol dynamics discussed above, but this had no effect on STAR sterol transfer activity, indicating that STAR was not acting on active cholesterol (96). Moreover, the findings that a CAH STAR mutation reduces STAR activity, but is functional in terms of sterol binding and transport (98), in conjunction with the observation of STAR’s extramitochondrial site of action, suggest that STAR does not serve as a cellular sterol transporter; rather, it acts as a cellular sterol sensor. Moreover, the findings that STAR associates physically and functionally with the outer mitochondrial membrane proteins, such as the voltage-dependent anion channel (VDAC) (99), indicate that additional molecular interactions contribute to STAR’s action on the outer mitochondrial membrane and that reductionist models of STAR acting in isolation are overly simplistic. Moreover, a large body of work illustrates that STAR requires conformational flexibility for its action, which is consistent with both transporter and sensor hypotheses; as such, further work will be required to reconcile models of STAR action (100–102).

The ERC: Niemann–Pick Type C Proteins and STARD3 The ERC contains among the highest cholesterol content in the cell, in many cases rivaling that of the plasma membrane. Moreover, interest in ERC cholesterol trafficking has been motivated by the study of the human disease of Niemann–Pick type C (NPC), a condition in which the ERC pathologically collects cholesterol and glycosphingolipids leading to progressive neurodegeneration and death (103). Genetic linkage analysis has revealed mutations in two genes, termed NPC1 and NPC2, that underlie the pathology of NPC disease, with NPC1 accounting for approximately 95% of cases (104) and NPC2 accounting for the remaining 5% (105). NPC1 is a 1278-residue endosomal integral membrane glycoprotein containing 13 transmembrane domains, and NPC2 is a soluble 151-residue glycoprotein found in the lysosomal lumen. Both proteins 904

are cholesterol binding (106,107), with NPC2 enveloping the aliphatic end of the molecule, leaving the 3β-OH group exposed, while NPC1 binds the 3β-OH end of the molecule at its N-terminus located in the lysosomal lumen. The reverse orientation of cholesterol binding by these two molecules has led to the model that NPC1 and NPC2 act in tandem transporting cholesterol out the ERC (108). In this model, NPC2, which appears to function as a soluble lipid-binding and transfer protein (109), delivers cholesterol from the endosomal lipoproteins to NPC1, which promotes cholesterol egress from the organelle. Interestingly, the BALB/c mouse characterized by a spontaneous mutation that assisted in identifying NPC mutations in humans (110) was identified as hypogonadal (though still fertile), with diminished capacity to synthesize testosterone (111). Whether NPC-mediated endosomal cholesterol directly contributes to steroidogenesis, or whether these observations were due to non-specific toxicity, remains unclear. In addition to the NPC proteins, the sterol-associated STARD3 protein has been linked to ERC cholesterol efflux. STARD3 is the most closely related sterol-associated START domain protein to STAR, and it was originally termed MLN64 for metastatic lymph node, clone 64. STARD3 was initially identified in a screen of clones prepared from the metastatic lymph nodes of women with ductal breast carcinoma (112). In humans, the STARD3 gene is located on chromosome 17q12-21, a region termed the HER2 amplicon, containing ERBB2, TOP2 and MED1, among others, and heterogeneously amplified in approximately 15% of breast cancers (113,114). Whether STARD3 plays a key role in the initiation and/or progression of breast cancer remains unclear, but co-silencing HER2 and STARD3 led to the additive inhibition of cell viability, and also induced apoptosis in breast cancer cells, suggesting that STARD3 may be a potential pharmacological target for breast cancer therapy (115,116). Human STARD3 contains 445-amino acid residues and is divided into two regions. The N-terminus of the protein contains four predicted transmembrane helices and is homologous to the 234-amino acid STARD3NL/MENTHO protein located on chromosome 7 (117). The C-terminus of STARD3 contains the START domain, which is structurally similar to STAR (89,118), and similar to the START domain of STAR, it Traffic 2014; 15: 895–914

Protein Motifs in Sterol Traffic and Metabolism

is able to stimulate mitochondrial cholesterol movement (119). STARD3 is integrated in the endosomal membrane through its STARD3NL-like motif (120), with its START domain pointing into the cytoplasm. This cytoplasmically facing topology, in conjunction with the finding that the STAR3NL-like motif is able to bind cholesterol (121), makes STARD3 an attractive candidate for cholesterol efflux from the ERC. Indeed, there is evidence that STARD3 is able to mediate cholesterol egress from the endosomes to the mitochondria (122), partnering with NPC2 in the absence of NPC1 (123). Moreover, STARD3 is proposed to play a role in steroidogenic cells that do not express STAR, such as in the placenta (124), because STARD3 START domain fragments have been found in these cells (118). Moreover, the STARD3 START domain possesses approximately 75% of the steroidogenic activity as the STAR START domain (125). However, the role that STARD3 plays in mitochondrial cholesterol delivery or in intramitochondrial cholesterol transport remains controversial, as STARD3 expression in hepatocytes negates the steroidogenic activity of exogenously expressed STAR (126). Furthermore, the targeted deletion of the STARD3 START domain in mice causes little to no alterations in markers of cellular and physiological sterol metabolism (127). It is plausible that STARD3 does not play a role as a cholesterol transporter per se, but rather as a sterol sensor facilitating sterol movement, as it is enriched in late endosomes (128) that are also enriched in cholesterol (129). In addition, the STARD3NL-like domain facilitates a tethering interaction with the resident ER vesicle-associated protein (130). Interestingly, STARD3 was recently identified as a high-affinity lutein-binding protein in primate retina (131,132). The STARD3 homolog in the moth Bombyx mori was also discovered to bind lutein, (133) suggesting that STARD3 may be evolutionarily promiscuous in the binding of hydrophobic molecules.

The Cytosolic START Domain Proteins, STARD4, STARD5 and STARD6 Three additional START domain proteins are associated with cholesterol, though their role in steroidogenesis, if any, is unclear. These proteins, STARD4, STARD5 and STARD6, lack any domains outside of the START domain (134), and as they lack any targeting sequences, they are generally Traffic 2014; 15: 895–914

considered cytosolic (135). Evidence for STARD4 from hepatocytes and macrophages supports this hypothesis, though the minor association of STARD4 with the ER has been observed (136). STARD5 has also been observed in a number of compartments, including the Golgi apparatus and the nucleus (137), though the role of these localizations is unclear. Of these proteins, STARD4’s role in cellular cholesterol trafficking is the best supported. As noted above, the crystal structure of STARD4 consists of the basic helix–loop–helix domain, which is characteristic of the START domain (89). STARD4 expression is regulated by the sterol regulatory element-binding protein-2 (SREBP-2) sterol synthetic machinery (138), the only putative cellular sterol transporter to be regulated by the SREBP-2 genetic network (139). Moreover, STARD4 was found to bind cholesterol using a cholesterol overlay assay, and it increased cholesteryl ester formation in mouse hepatocytes when overexpressed, suggesting that STARD4 plays a role in ER cholesterol delivery and in ACAT sterol esterification (140). Conversely, STARD4 knockdown in the hepatocyte HepG2 cell line resulted in the retention of sterol in the plasma membrane, with reduced trafficking to the ERC and ER (140). Elegant work by Mesmin et al. provided strong evidence that STARD4 functions as a theoretical lipid-binding and transport protein, as STARD4 knockdown in U2OS osteosarcoma cells was able to be compensated for by the polysaccharide MCD (141). However, similar to other START proteins, e.g. STARD3, knockout animals deleted for STARD4 show minimal phenotype (142). Comparatively less is known about the START domain proteins, STARD5 and STARD6. Both consist solely of a START domain with no other features, much akin to STARD4 (134). STARD5 has been shown to be expressed in macrophages (138), testicular Sertoli cells (143) and in the kidney (137) and liver (144), though its role in these tissues is unknown. STARD5 has been shown to biochemically bind both radiolabeled cholesterol and 25-hydroxycholesterol, but not 27-hydroxycholesterol (145), suggesting that STARD5 is involved in intracellular cholesterol and oxysterol homeostasis, possibly in the context of ER stress (146). However, a structural analysis of STARD5 using circular dichroism and nuclear magnetic resonance indicated that STARD5 is not able to bind cholesterol or 25-hydroxycholesterol; rather, it binds the bile acids, cholate and chenocholate, indicating that it plays 905

Midzak and Papadopoulos

a role in bile acid signaling and transport (147,148). The discordant results observed between the methodologies used remain to be resolved. STARD6 has been observed in rat male germ cells (149) as well as in the rat brain (150), though again its exact function is unclear. Most recently, STARD6 expression was detected in the steroidogenic cells of the porcine corpus luteum, peaking during the mid-luteal phase (151). The observation that STARD6 is capable of promoting intramitochondrial cholesterol movement and steroidogenesis in a COS-1 cell system transfected with the CYP11A1 enzymatic system, and it does so to the same levels as STAR (125,151), suggests that STARD6 may be involved in ovarian cholesterol trafficking and steroidogenesis.

ORD Family of Sterol Transporters The ORD family of proteins is another family of soluble protein factors implicated in cellular sterol movement. The founding member of the family, oxysterol-binding protein (OSBP), was originally identified during a search for the mechanisms that provide acute control of sterol synthesis by oxysterols (152), though OSBP does not appear to be involved in this process (153). In humans, there are 12 members of the ORD family (154,155), and in addition to the ORD, many of the proteins contain additional proteinand lipid-binding domains that have been hypothesized to target them to subcellular spaces. ORD proteins have attracted a significant amount of research into cellular cholesterol trafficking, especially in yeast, where they comprise the largest family of sterol-binding proteins, as yeast lacks the START domain family (156). However, there is no evidence that any of the ORD proteins play any role in steroidogenesis or in the trafficking of cholesterol to or within the mitochondria, and so they will not be reviewed here; more general recent reviews are available (74,157,158). It should be noted that one recent finding is of interest to the present discussion, as molecular promiscuity has arisen when discussing STARD3 and STARD5. In this context, it has been observed that the ORD of the yeast Osh4p, as well as the mammalian OSBP, is able to bind and transport the lipid phosphatidylinositol 4-phosphate using their ORDs – a process that competes with their cholesterol-binding and transport capabilities at the same domains (159,160). This ligand promiscuity has been 906

tied to cellular function by observing that these proteins localize to the ER–Golgi interface, where they appear to regulate sterol distribution between the organelles, as well as in ER cholesterol sensing. It is interesting to speculate whether such a mechanism translates to the START domain proteins discussed above.

The Translocator Protein (18 kDa) and the Cholesterol Recognition Amino Acid Consensus Motif The finding that benzodiazepines were able to stimulate steroidogenesis in a variety of steroidogenic systems, including classical and non-classical steroidogenic cells (161,162), led to the proposal that a peripheral benzodiazepine receptor, later renamed the 18-kDa translocator protein (TSPO) (163), is involved in mitochondrial cholesterol transport and steroidogenesis. This hypothesis was supported by findings that other pharmacological TSPO ligands, as well as the endogenous protein diazepam-binding inhibitor (164) [also known as acyl-CoA-binding domain protein 1 (ACBD1)], were able to stimulate steroid production in a TSPO-associated manner. The involvement of TSPO in natural steroidogenesis was further supported by pharmacological work showing that the TSPO ligand, flunitrazepam, was able to inhibit steroidogenesis in Leydig and adrenal cell lines (165), and in genetic work in the transformed rat R2C Leydig cell line. Homologous recombination-based TSPO deletion in R2C cells eliminated the ability of cells to synthesize steroids, and this lesion was corrected by transfection of a plasmid containing wild-type TSPO and by the subsequent expression of the protein (166). The absolute necessity of TSPO has been recently questioned; however, Testis-specific knockout of TSPO in a mouse model did not exhibit any steroidogenic phenotype (167). Whether this lack of phenotype was due to TSPO not being a component of the natural steroidogenic machinery, or due to compensation by another factor, remains to be determined (168). Regardless, TSPO’s ability to stimulate steroidogenesis and mitochondrial cholesterol movement in both classical steroidogenic cells as well as in non-classical tissues makes it an attractive target of study, especially as it is a proven druggable target (169–174). Traffic 2014; 15: 895–914

Protein Motifs in Sterol Traffic and Metabolism

TSPO’s involvement in drug-mediated steroidogenesis and the associated mitochondrial cholesterol transport has led to an investigation of this protein’s association with cholesterol. Deletion mutation analysis of TSPO revealed that it was able to increase the uptake of cholesterol in Escherichia coli, which naturally lacks cholesterol, and that the deletion of the C-terminus of the protein abolished this cholesterol uptake. Comparison of this region of TSPO with other proteins associated with cholesterol revealed a consensus sequence consisting of -L/V-X(1–5) -Y-X(1–5) -R/K-, where X(1–5) corresponds to one to five of any amino acid (175). Cholesterol inhibited the cross-linking of radioactive promegestone to a recombinant peptide containing the cholesterol recognition amino acid consensus (CRAC) motif of TSPO (176), and recombinant mouse TSPO reconstituted in liposomes was found to bind radioactive cholesterol with nanomolar affinity (177). Furthermore, the mutation of the central tyrosine in the recombinant protein abolished cholesterol binding (178). The involvement of TSPO and its CRAC motif was further supported by work showing that the pharmacological targeting of CRAC with the steroid 5-androsten-3β,17,19-triol was able to inhibit acute hormone-mediated steroidogenesis (179). Recently, the crystal structure of mouse TSPO was solved in complex with the TSPO drug ligand, PK 11195, showing that drug ligand binding to TSPO distorts the CRAC motif, possibly explaining how TSPO ligands stimulate steroid synthesis by releasing TSPO-bound cholesterol (180). The definition of the CRAC motif has also been recently expanded to include its palindromic sequence following the discovery that the left-to-right CRAC motif was insufficient in detecting cholesterol-binding portions of the human nicotinic acetylcholine receptor, and it was further delineated as a CARC motif (181,182). The cholesterol specificity of the CRAC motif is also of great interest, in light of observations that TSPO’s CRAC motif is also able to bind the steroids promegestone (175), 5-androsten-3β,17,19-triol and 5-androsten-3β,19-diol-17-one (but not 5-androsten3β,7β,19-triol or 5-androsten-3β,11β,19-triol; 183), and steroids are able to induce structural changes (184). The role of such promiscuity remains to be determined. Although TSPO has been hypothesized to be a cholesterol translocator (185), whether it facilitates cholesterol movement on its own, or whether it facilitates cholesterol Traffic 2014; 15: 895–914

movement through its association with other proteins remains unclear. TSPO has been known to closely interact with mitochondrial VDAC (186). VDAC itself is a potential regulator of mitochondrial cellular cholesterol movement (187) and a potential functional partner for STAR (99). TSPO and STAR appear to be functionally complementary in steroidogenic model systems (188), though the physical association between these proteins has been less than conclusive, with FRET and protein cross-linking (189,190) suggesting that a physical association exists, but bioluminescence resonance energy transfer (BRET) analysis remains inconclusive (191). Moreover, the hormonal stimulation of steroidogenic cells triggers the assembly of a multiprotein complex at the outer mitochondrial membrane, termed the transduceosome (190), containing VDAC, protein kinase A (PKA), the PKA-anchoring protein ACBD3 (192) and the inner mitochondrial membrane/membrane contact site protein ATAD3 (193). In this complex, TSPO may serve as a cholesterol reservoir, specifically targeting cholesterol to contact sites for delivery to CYP11A1.

Conclusions and Perspectives We have described the current understanding of steroid biosynthesis in relation to cholesterol’s physical properties and its transport via cellular protein machinery. Biophysical models have begun to explain the molecular details of cholesterol’s dynamics within phospholipid membranes. Genome sequencing has revealed the numerous families of lipid-binding proteins (LBPs) that have been conserved throughout many domains of life. Biochemical and cell biology experiments have begun to demonstrate the roles of these LBPs in cellular cholesterol homeostasis. Proteomic technologies have begun to elucidate the protein interactions that drive cellular processes. With regard to steroidogenesis, the convergence of technologies and investigations has painted a picture of macromolecular protein complexes with components that work in concert to transport cholesterol into the mitochondria. However, the molecular details of cholesterol movement between the lipid and protein environments involved are still lacking. Further work incorporating the models and concepts outlined here will be required to better understand this fundamental process of steroid biosynthesis. 907

Midzak and Papadopoulos

Acknowledgments This work was supported by grants from the Canadian Institutes of Health Research (MOP102647 and MOP125983), a Canada Research Chair in Biochemical Pharmacology to V. P. and a postdoctoral fellowship from the Fonds de recherche du Québec – Santé to A. M. The authors have no conflict of interest to declare.

References 1. Miller WL, Bose HS. Early steps in steroidogenesis: intracellular cholesterol trafficking. J Lipid Res 2011;52:2111–2135. 2. Payne AH, Hales DB. Overview of steroidogenic enzymes in the pathway from cholesterol to active steroid hormones. Endocr Rev 2004;25:947–970. 3. Miller WL, Auchus RJ. The molecular biology, biochemistry, and physiology of human steroidogenesis and its disorders. Endocr Rev 2011;32:81–151. 4. Hoffmann AF, Hagey LR, Krasowski MD. Bile salts of vertebrates: structural variation and possible evolutionary significance. J Lipid Res 2010;51:226–246. 5. Brown MR, Sieglaff DH, Rees HH. Gonadal ecdysteroidogenesis in Arthropoda: occurrence and regulation. Annu Rev Entomol 2009;54:105–125. 6. Chambon P. The nuclear receptor superfamily: a personal retrospect on the first two decades. Mol Endocrinol 2005;19:1418–1428. 7. Evans RM. The nuclear receptor superfamily: a Rosetta stone for physiology. Mol Endocrinol 2005;19:1429–1438. 8. Miller WL. A brief history of adrenal research. Steroidogenesis – the soul of the adrenal. Mol Cell Endocrinol 2013;371:5–14. 9. Gonzalez FJ. The molecular biology of cytochrome P450s. Pharmacol Rev 1988;40:243–288. 10. Miller WL. Minireview: regulation of steroidogenesis by electron transfer. Endocrinology 2005;146:2544–2550. 11. Agarwal AK, Auchus RJ. Cellular redox state regulates hydroxysteroid dehydrogenase activity and intracellular hormone potency. Endocrinology 2005;146:2531–2538. 12. Thomas JL, Duax WL, Addlagatta A, Brandt S, Fuller RR, Norris W. Structure/function relationships responsible for coenzyme specificity and the isomerase activity of human type 1 β-hydroxysteroid dehydrogenase/isomerase. J Biol Chem 2003;278:35483–35490. 13. Lathe R, Kotelevsky Y. Steroid signaling: ligand-binding promiscuity, molecular symmetry, and the need for gating. Steroids 2014;82C:14–22. 14. Papadopoulos V, Lecanu L, Brown RC, Han Z, Yao ZX. Peripheral-type benzodiazepine receptor in neurosteroid biosynthesis, neuropathology and neurological disorders. Neuroscience 2006;138:749–756. 15. Shikita M, Hall PF. Cytochrome P-450 from bovine adrenocortical mitochondria: an enzyme for the side chain cleavage of cholesterol. I. Purification and properties. J Biol Chem 1973;248:5598–5604. 16. Norlin M, Wikvall K. Enzymes in the conversion of cholesterol into bile acids. Curr Mol Med 2007;7:199–218.

908

17. Kawamoto T, Mitsuuchi Y, Toda K, Yokoyama Y, Miyahara K, Miura S, Ohnishi T, Ichikawa Y, Nakaok K, Imura H, Ulick S, Shizuta Y. Role of steroid 11β-hydroxylase and steroid 18-hydroxylase in the biosynthesis of glucocorticoids and mineralocorticoids in humans. Proc Natl Acad Sci U S A 1992;89:1458–1462. 18. Takeyama K, Kitanaka S, Sato T, Kobori M, Yanagisawa J, Kato S. 25-Hydroxyvitamin D3 1a-hydroxylase and vitamin D synthesis. Science 1997;277:1827–1830. 19. Pikuleva IA, Waterman MR. Cytochromes p450: roles in diseases. J Biol Chem 2013;288:17091–17098. 20. Markov GV, Tavares R, Dauphin-Villemant C, Demeneix BA, Baker ME, Laudet V. Independent elaboration of steroid hormone signaling pathways in metazoans. Proc Natl Acad Sci U S A 2009;106:11913–11918. 21. Fan J, Papadopoulos V. Evolutionary origin of the mitochondrial cholesterol transport machinery reveals a universal mechanism of steroid hormone biosynthesis in animals. PLoS One 2013;8:e76701. 22. Burstein S, Gut M. Intermediates in the conversion of cholesterol to pregnenolone: kinetics and mechanism. Steroids 1976;28:115–131. 23. Mast N, Annalora AJ, Lodowski DT, Palczewski K, Stout CD, Pikuleva IA. Structural basis for three-step sequential catalysis by the cholesterol side chain cleavage enzyme CYP11A1. J Biol Chem 2011;286:5607–5613. 24. Strushkevich N, MacKenzie F, Cherkesova T, Grabovec I, Usanov S, Park HW. Structural basis for pregnenolone biosynthesis by the mitochondrial monooxygenase system. Proc Natl Acad Sci U S A 2011;108:10139–10143. 25. Miller WL. Steroid hormone synthesis in mitochondria. Mol Cell Endocrinol 2013;379:62–73. 26. Alfano J, Brownie AC, Orme-Johnson WH, Beinert H. Adrenal mitochondrial cytochrome P-450 and cholesterol side chain cleavage activity. Differences in the response of the zona glomerulosa and zona fasciculata-reticularis to adrenocorticotropic hormone and its withdrawal. J Biol Chem 1973;248:7860–7864. 27. Payne AH. Hormonal regulation of cytochrome P450 enzymes, cholesterol side-chain cleavage and 17 alpha-hydroxylase/C17-20 lyase in Leydig cells. Biol Reprod 1990;42:399–404. 28. Schimmer BP, White PC. Minireview: steroidogenic factor 1: its roles in differentiation, development, and disease. Mol Endocrinol 2010;24:1322–1337. 29. Shih MC, Chiu YN, Hu MC, Guo IC, Chung BC. Regulation of steroid production: analysis of Cyp11a1 promoter. Mol Cell Endocrinol 2011;336:80–84. 30. Stevens VL, Tribble DL, Lambeth JD. Regulation of mitochondrial compartment volumes in rat adrenal cortex by ether stress. Arch Biochem Biophys 1985;242:324–327. 31. Privalle CT, Crivello JF, Jefcoate CR. Regulation of intramitochondrial cholesterol transfer to side-chain cleavage cytochrome P-450 in rat adrenal gland. Proc Natl Acad Sci U S A 1983;80:702–706. 32. Privalle CT, McNamara BC, Dhariwal MS, Jefcoate CR. ACTH control of cholesterol side-chain cleavage at adrenal mitochondrial

Traffic 2014; 15: 895–914

Protein Motifs in Sterol Traffic and Metabolism

33.

34. 35. 36.

37. 38. 39. 40.

41.

42.

43.

44. 45.

46.

47.

48. 49. 50.

51.

cytochrome P-450scc. Regulation of intramitochondrial cholesterol transfer. Mol Cell Endocrinol 1987;53:87–101. van Meer G, Voelker DR, Feigenson GW. Membrane lipids: where they are and how they behave. Nat Rev Mol Cell Biol 2008;9:112–2411. Ikonen E. Cellular cholesterol trafficking and compartmentalization. Nat Rev Mol Cell Biol 2008;9:125–138. Mesmin B, Maxfield FR. Intracellular sterol dynamics. Biochim Biophys Acta 2009;1791:636–645. Radhakrishnan A, Goldstein JL, McDonald JG, Brown MS. Switch-like control of SREBP-2 transport triggered by small changes in ER cholesterol: a delicate balance. Cell Metab 2008;8:512–521. Prinz WA. Lipid trafficking sans vesicles: where, why, how? Cell 2010;143:870–874. Sturley SL, Hussain MM. Lipid droplet formation on opposing sides of the endoplasmic reticulum. J Lipid Res 2012;53:1800–1810. Brown MS, Goldstein JL. Cholesterol feedback: from Schoenheimer’s bottle to Scap’s MELADL. J Lipid Res 2009;50(Suppl):S15–S27. Baumann NA, Sullivan DP, Ohvo-Rekilä H, Simonot C, Pottekat A, Klaassen Z, Beh CT, Menon AK. Transport of newly synthesized sterol to the sterol-enriched plasma membrane occurs via nonvesicular equilibration. Biochemistry 2005;44:5816–5826. Zoller LC, Malamed S. Acute effects of ACTH on dissociated adrenocortical cells: quantitative changes in mitochondria and lipid droplets. Anat Rec 1975;182:473–478. Nan X, Potma EO, Xie XS. Nonperturbative chemical imaging of organelle transport in living cells with coherent anti-stokes Raman scattering microscopy. Biophys J 2006;91:728–735. Shen WJ, Zaidi SK, Patel S, Cortez Y, Ueno M, Azhar R, Azhar S, Kraemer FB. Ablation of vimentin results in defective steroidogenesis. Endocrinology 2012;153:3249–3257. Freeman DA. Cyclic AMP mediated modification of cholesterol traffic in Leydig tumor cells. J Biol Chem 1987;262:13061–13068. Freeman DA. Plasma membrane cholesterol: removal and insertion into the membrane and utilization as substrate for steroidogenesis. Endocrinology 1989;124:2527–2534. Nagy L, Freeman DA. Effect of cholesterol transport inhibitors on steroidogenesis and plasma membrane cholesterol transport in cultured MA-10 Leydig tumor cells. Endocrinology 1990;126:2267–2276. Lange Y, Steck TL. Cholesterol homeostasis and the escape tendency (activity) of plasma membrane cholesterol. Prog Lipid Res 2008;47:319–332. Lindblom G, Orädd G. Lipid lateral diffusion and membrane heterogeneity. Biochim Biophys Acta 2009;1788:234–244. Radhakrishnan A, McConnell HM. Chemical activity of cholesterol in membranes. Biochemistry 2000;39:8119–8124. Lange Y, Cutler HB, Steck TL. The effect of cholesterol and other intercalated amphipaths on the contour and stability of the isolated red cell membrane. J Biol Chem 1980;255:9331–9337. Steck TL, Ye J, Lange Y. Probing red cell membrane cholesterol movement with cyclodextrin. Biophys J 2002;83:2118–2125.

Traffic 2014; 15: 895–914

52. Leventis R, Silvius JR. Use of cyclodextrins to monitor transbilayer movement and differential lipid affinities of cholesterol. Biophys J 2001;81:2257–2267. 53. Radhakrishnan A, McConnell HM. Condensed complexes of cholesterol and phospholipids. Biophys J 1999;77:1507–1517. 54. Radhakrishnan A, McConnell H. Condensed complexes in vesicles containing cholesterol and phospholipids. Proc Natl Acad Sci U S A 2005;102:12662–12666. 55. McConnell HM, Radhakrishnan A. Condensed complexes of cholesterol and phospholipids. Biochim Biophys Acta 2003;1610:159–173. 56. Lange Y, Ye J, Steck TL. How cholesterol homeostasis is regulated by plasma membrane cholesterol in excess of phospholipids. Proc Natl Acad Sci U S A 2004;101:11664–11667. 57. Schneiter R, Brügger B, Sandhoff R, Zellnig G, Leber A, Lampl M, Athenstaedt K, Hrastnik C, Eder S, Daum G, Paltauf F, Wieland FT, Kohlwein SD. Electrospray ionization tandem mass spectrometry (ESI-MS/MS) analysis of the lipid molecular species composition of yeast subcellular membranes reveals acyl chain-based sorting/remodeling of distinct molecular species en route to the plasma membrane. J Cell Biol 1999;146:741–754. 58. Lange Y, Ye J, Steck TL. Activation of membrane cholesterol by displacement from phospholipids. J Biol Chem 2005;280:36126–36131. 59. Ratajczak MK, Ko YT, Lange Y, Steck TL, Lee KY. Cholesterol displacement from membrane phospholipids by hexadecanol. Biophys J 2007;93:2038–2047. 60. Huang J, Feigenson GW. A microscopic interaction model of maximum solubility of cholesterol in lipid bilayers. Biophys J 1999;76:2142–2157. 61. Ali MR, Cheng KH, Huang J. Assess the nature of cholesterol-lipid interactions through the chemical potential of cholesterol in phosphatidylcholine bilayers. Proc Natl Acad Sci U S A 2007;104:5372–5377. 62. Daum G, Vance JE. Import of lipids into mitochondria. Prog Lipid Res 1997;36:103–130. 63. Horvath SE, Daum G. Lipids of mitochondria. Prog Lipid Res 2013;52:590–614. 64. Wang HP, Pfeiffer DR, Kimura T, Tchen TT. Phospholipids of adrenal cortex mitochondria and the steroid hydroxylases: the lipid-environment of cytochrome P-450. Biochem Biophys Res Commun 1974;57:93–99. 65. Cheng B, Kimura T. The distribution of cholesterol and phospholipid composition in submitochondrial membranes from bovine adrenal cortex: fundamental studies of steroidogenic mitochondria. Lipids 1983;18:577–584. 66. Stevens VL, Lambeth JD, Merrill AH Jr. Use of cytochrome P-450scc to measure cholesterol-lipid interactions. Biochemistry 1986;25:4287–4292. 67. Kido T, Yamakura F, Kimura T. The transfer of cholesterol and hydroxycholesterol derivatives from liposome to soluble cytochrome P-450 scc. Biochim Biophys Acta 1981;666:370–381.

909

Midzak and Papadopoulos

68. Kido T, Kimura T. Stimulation of cholesterol binding to steroid-free cytochrome P-450scc by poly(L-lysine). The implication in functions of labile protein factor for adrenocortical steroidogenesis. J Biol Chem 1981;256:8561–8568. 69. Igarashi Y, Kimura T. Importance of the unsaturated fatty acyl group of phospholipids in their stimulatory role on rat adrenal mitochondrial steroidogenesis. Biochemistry 1986;25:6461–6466. 70. Igarashi Y, Kimura T. Adrenic acid content in rat adrenal mitochondrial phosphatidylethanolamine and its relation to ACTH-mediated stimulation of cholesterol side chain cleavage reaction. J Biol Chem 1986;261:14118–14124. 71. Phillips MC, Johnson WJ, Rothblat GH. Mechanisms and consequences of cellular cholesterol exchange and transfer. Biochim Biophys Acta 1987;906:223–276. 72. Lev S. Non-vesicular lipid transport by lipid-transfer proteins and beyond. Nat Rev Mol Cell Biol 2010;11:739–750. 73. Clark BJ. The mammalian START domain protein family in lipid transport in health and disease. J Endocrinol 2012;212:257–275. 74. Raychaudhuri S, Prinz WA. The diverse functions of oxysterol-binding proteins. Annu Rev Cell Dev Biol 2010;26:157–177. 75. Pon LA, Orme-Johnson NR. Acute stimulation of steroidogenesis in corpus luteum and adrenal cortex by peptide hormones. Rapid induction of a similar protein in both tissues. J Biol Chem 1986;261:6594–6599. 76. Soccio RE, Breslow JL. StAR-related lipid transfer (START) proteins: mediators of intracellular lipid metabolism. J Biol Chem 2003;278:22183–22186. 77. Iyer LM, Koonin EV, Aravind L. Adaptations of the helix-grip fold for ligand binding and catalysis in the START domain superfamily. Proteins 2001;43:134–144. 78. Pon LA, Hartigan JA, Orme-Johnson NR. Acute ACTH regulation of adrenal corticosteroid biosynthesis. Rapid accumulation of a phosphoprotein. J Biol Chem 1986;261:13309–13316. 79. Clark BJ, Wells J, King SR, Stocco DM. The purification, cloning, and expression of a novel luteinizing hormone-induced mitochondrial protein in MA-10 mouse Leydig tumor cells. Characterization of the steroidogenic acute regulatory protein (StAR). J Biol Chem 1994;269:28314–28322. 80. Bose HS, Sugawara T, Strauss JF 3rd, Miller WL, International Congenital Lipoid Adrenal Hyperplasia Consortium. The pathophysiology and genetics of congenital lipoid adrenal hyperplasia. N Engl J Med 1996;335:1870–1878. 81. Lin D, Sugawara T, Strauss JF 3rd, Clark BJ, Stocco DM, Saenger P, Rogol A, Miller WL. Role of steroidogenic acute regulatory protein in adrenal and gonadal steroidogenesis. Science 1995;267:1828–1831. 82. Bose HS, Sato S, Aisenberg J, Shalev SA, Matsuo N, Miller WL. Mutations in the steroidogenic acute regulatory protein (StAR) in six patients with congenital lipoid adrenal hyperplasia. J Clin Endocrinol Metab 2000;85:3636–3639. 83. Sahakitrungruang T, Soccio RE, Lang-Muritano M, Walker JM, Achermann JC, Miller WL. Clinical, genetic, and functional characterization of four patients carrying partial loss-of-function

910

84.

85. 86.

87.

88.

89. 90.

91.

92.

93.

94.

95.

96.

97.

mutations in the steroidogenic acute regulatory protein (StAR). J Clin Endocrinol Metab 2010;95:3352–3359. Caron KM, Soo SC, Wetsel WC, Stocco DM, Clark BJ, Parker KL. Targeted disruption of the mouse gene encoding steroidogenic acute regulatory protein provides insights into congenital lipoid adrenal hyperplasia. Proc Natl Acad Sci U S A 1997;94:11540–11545. Stocco DM, Clark BJ. Regulation of the acute production of steroids in steroidogenic cells. Endocr Rev 1996;17:221–244. Arakane F, Kallen CB, Watari H, Foster JA, Sepuri NB, Pain D, Stayrook SE, Lewis M, Gerton GL, Strauss JF 3rd.. The mechanism of action of steroidogenic acute regulatory protein (StAR). StAR acts on the outside of mitochondria to stimulate steroidogenesis. J Biol Chem 1998;273:16339–16345. Bose HS, Lingappa VR, Miller WL. Rapid regulation of steroidogenesis by mitochondrial protein import. Nature 2002;417:87–91. Sasaki G, Ishii T, Jeyasuria P, Jo Y, Bahat A, Orly J, Hasegawa T, Parker KL. Complex role of the mitochondrial targeting signal in the function of steroidogenic acute regulatory protein revealed by bacterial artificial chromosome transgenesis in vivo. Mol Endocrinol 2008;22:951–964. Tsujishita Y, Hurley JH. Structure and lipid transport mechanism of a StAR-related domain. Nat Struct Biol 2000;7:408–414. Romanowski MJ, Soccio RE, Breslow JL, Burley SK. Crystal structure of the Mus musculus cholesterol-regulated START protein 4 (StarD4) containing a StAR-related lipid transfer domain. Proc Natl Acad Sci U S A 2002;99:6949–6954. Thorsell AG, Lee WH, Persson C, Siponen MI, Nilsson M, Busam RD, Kotenyova T, Schüler H, Lehtiö L. Comparative structural analysis of lipid binding START domains. PLoS One 2011;6:e19521. Murcia M, Faráldo-Gómez JD, Maxfield FR, Roux B. Modeling the structure of the StART domains of MLN64 and StAR proteins in complex with cholesterol. J Lipid Res 2006;47:2614–2630. Barbar E, Lavigne P, Lehoux JG. Validation of the mechanism of cholesterol binding by StAR using short molecular dynamics simulations. J Steroid Biochem Mol Biol 2009;113:92–97. Kallen CB, Billheimer JT, Summers SA, Stayrook SE, Lewis M, Strauss JF 3rd.. Steroidogenic acute regulatory protein (StAR) is a sterol transfer protein. J Biol Chem 1998;273:26285–26288. Petrescu AD, Gallegos AM, Okamura Y, Strauss JF 3rd, Schroeder F. Steroidogenic acute regulatory protein binds cholesterol and modulates mitochondrial membrane sterol domain dynamics. J Biol Chem 2001;276:36970–36982. Tuckey RC, Headlam MJ, Bose HS, Miller WL. Transfer of cholesterol between phospholipid vesicles mediated by the steroidogenic acute regulatory protein (StAR). J Biol Chem 2002;277:47123–47128. Artemenko IP, Zhao D, Hales DB, Hales KH, Jefcoate CR. Mitochondrial processing of newly synthesized steroidogenic acute regulatory protein (StAR), but not total StAR, mediates cholesterol transfer to cytochrome P450 side chain cleavage enzyme in adrenal cells. J Biol Chem 2001;276:46583–46596.

Traffic 2014; 15: 895–914

Protein Motifs in Sterol Traffic and Metabolism

98. Baker BY, Epand RF, Epand RM, Miller WL. Cholesterol binding does not predict activity of the steroidogenic acute regulatory protein, StAR. J Biol Chem 2007;282:10223–10232. 99. Bose M, Whittal RM, Miller WL, Bose HS. Steroidogenic activity of StAR requires contact with mitochondrial VDAC1 and phosphate carrier protein. J Biol Chem 2008;283:8837–8845. 100. Bose HS, Whittal RM, Baldwin MA, Miller WL. The active form of the steroidogenic acute regulatory protein, StAR, appears to be a molten globule. Proc Natl Acad Sci U S A 1999;96:7250–7255. 101. Baker BY, Yaworsky DC, Miller WL. A pH-dependent molten globule transition is required for activity of the steroidogenic acute regulatory protein, StAR. J Biol Chem 2005;280:41753–41760. 102. Yaworsky DC, Baker BY, Bose HS, Best KB, Jensen LB, Bell JD, Baldwin MA, Miller WL. pH-dependent Interactions of the carboxyl-terminal helix of steroidogenic acute regulatory protein with synthetic membranes. J Biol Chem 2005;280:2045–2054. 103. Vanier MT, Millat G. Niemann-Pick disease type C. Clin Genet 2003;64:269–281. 104. Park WD, O’Brien JF, Lundquist PA, Kraft DL, Vockley CW, Karnes PS, Patterson MC, Snow K. Identification of 58 novel mutations in Niemann-Pick disease type C: correlation with biochemical phenotype and importance of PTC1-like domains in NPC1. Hum Mutat 2003;22:313–325. 105. Naureckiene S, Sleat DE, Lackland H, Fensom A, Vanier MT, Wattiaux R, Jadot M, Lobel P. Identification of HE1 as the second gene of Niemann-Pick C disease. Science 2000;290:2298–2301. 106. Kwon HJ, Abi-Mosleh L, Wang ML, Deisenhofer J, Goldstein JL, Brown MS, Infante RE. Structure of N-terminal domain of NPC1 reveals distinct subdomains for binding and transfer of cholesterol. Cell 2009 Jun 26;137:1213–1224. 107. Xu S, Benoff B, Liou HL, Lobel P, Stock AM. Structural basis of sterol binding by NPC2, a lysosomal protein deficient in Niemann-Pick type C2 disease. J Biol Chem 2007;282:23525–23531. 108. Wang ML, Motamed M, Infante RE, Abi-Mosleh L, Kwon HJ, Brown MS, Goldstein JL. Identification of surface residues on Niemann-Pick C2 essential for hydrophobic handoff of cholesterol to NPC1 in lysosomes. Cell Metab 2010;12:166–173. 109. Infante RE, Wang ML, Radhakrishnan A, Kwon HJ, Brown MS, Goldstein JL. NPC2 facilitates bidirectional transfer of cholesterol between NPC1 and lipid bilayers, a step in cholesterol egress from lysosomes. Proc Natl Acad Sci U S A 2008;105:15287–15292. 110. Pentchev PG, Comly ME, Kruth HS, Patel S, Proestel M, Weintroub H. The cholesterol storage disorder of the mutant BALB/c mouse. A primary genetic lesion closely linked to defective esterification of exogenously derived cholesterol and its relationship to human type C Niemann-Pick disease. J Biol Chem 1986;261:2772–2777. 111. Roff CF, Strauss JF 3rd, Goldin E, Jaffe H, Patterson MC, Agritellis GC, Hibbs AM, Garfield M, Brady RO, Pentchev PG. The murine Niemann-Pick type C lesion affects testosterone production. Endocrinology 1993;133:2913–2923. 112. Tomasetto C, Régnier C, Moog-Lutz C, Mattei MG, Chenard MP, Lidereau R, Basset P, Rio MC. Identification of four novel human genes amplified and overexpressed in breast carcinoma and localized

Traffic 2014; 15: 895–914

113.

114.

115.

116.

117.

118.

119.

120.

121.

122.

123.

124.

125.

126.

to the q11-q21.3 region of chromosome 17. Genomics 1995;28:367–376. Jacot W, Fiche M, Zaman K, Wolfer A, Lamy PJ. The HER2 amplicon in breast cancer: topoisomerase IIA and beyond. Biochim Biophys Acta 2013;1836:146–157. Sahlberg KK, Hongisto V, Edgren H, Mäkelä R, Hellström K, Due EU, Moen Vollan HK, Sahlberg N, Wolf M, Børresen-Dale AL, Perälä M, Kallioniemi O. The HER2 amplicon includes several genes required for the growth and survival of HER2 positive breast cancer cells. Mol Oncol 2013;7:392–401. Akula N, Midzak A, Lecanu L, Papadopoulos V. Identification of small-molecule inhibitors of the steroidogenic acute regulatory protein (STARD1) by structure-based design. Bioorg Med Chem Lett 2012;22:4139–4143. Chitrala KN, Yeguvapalli S. Ligand-based virtual screening to predict inhibitors against metastatic lymph node 64. J Recept Signal Transduct Res 2014;34:92–96. Alpy F, Wendling C, Rio MC, Tomasetto C. MENTHO, a MLN64 homologue devoid of the START domain. J Biol Chem 2002;277:50780–50787. Bose HS, Whittal RM, Huang MC, Baldwin MA, Miller WL. N-218 MLN64, a protein with StAR-like steroidogenic activity, is folded and cleaved similarly to StAR. Biochemistry 2000;39:11722–11731. Watari H, Arakane F, Moog-Lutz C, Kallen CB, Tomasetto C, Gerton GL, Rio MC, Baker ME, Strauss JF 3rd.. MLN64 contains a domain with homology to the steroidogenic acute regulatory protein (StAR) that stimulates steroidogenesis. Proc Natl Acad Sci U S A 1997;94:8462–8467. Liapis A, Chen FW, Davies JP, Wang R, Ioannou YA. MLN64 transport to the late endosome is regulated by binding to 14-3-3 via a non-canonical binding site. PLoS One 2012;7:e34424. Alpy F, Latchumanan VK, Kedinger V, Janoshazi A, Thiele C, Wendling C, Rio MC, Tomasetto C. Functional characterization of the MENTAL domain. J Biol Chem 2005;280:17945–17952. Zhang M, Liu P, Dwyer NK, Christenson LK, Fujimoto T, Martinez F, Comly M, Hanover JA, Blanchette-Mackie EJ, Strauss JF 3rd.. MLN64 mediates mobilization of lysosomal cholesterol to steroidogenic mitochondria. J Biol Chem 2002;277:33300–33310. Charman M, Kennedy BE, Osborne N, Karten B. MLN64 mediates egress of cholesterol from endosomes to mitochondria in the absence of functional Niemann-Pick Type C1 protein. J Lipid Res 2010;51:1023–1034. Olvera-Sanchez S, Espinosa-Garcia MT, Monreal J, Flores-Herrera O, Martinez F. Mitochondrial heat shock protein participates in placental steroidogenesis. Placenta 2011;32:222–229. Bose HS, Whittal RM, Ran Y, Bose M, Baker BY, Miller WL. StAR-like activity and molten globule behavior of StARD6, a male germ-line protein. Biochemistry 2008;47:2277–2288. Ren S, Hylemon P, Marques D, Hall E, Redford K, Gil G, Pandak WM. Effect of increasing the expression of cholesterol transporters (StAR, MLN64, and SCP-2) on bile acid synthesis. J Lipid Res 2004;45:2123–2131.

911

Midzak and Papadopoulos

127. Kishida T, Kostetskii I, Zhang Z, Martinez F, Liu P, Walkley SU, Dwyer NK, Blanchette-Mackie EJ, Radice GL, Strauss JF 3rd. Targeted mutation of the MLN64 START domain causes only modest alterations in cellular sterol metabolism. J Biol Chem 2004;279:19276–19285. 128. Hölttä-Vuori M, Alpy F, Tanhuanpää K, Jokitalo E, Mutka AL, Ikonen E. MLN64 is involved in actin-mediated dynamics of late endocytic organelles. Mol Biol Cell 2005;16:3873–3886. 129. van der Kant R, Zondervan I, Janssen L, Neefjes J. Cholesterol-binding molecules MLN64 and ORP1L mark distinct late endosomes with transporters ABCA3 and NPC1. J Lipid Res 2013;54:2153–2165. 130. Alpy F, Rousseau A, Schwab Y, Legueux F, Stoll I, Wendling C, Spiegelhalter C, Kessler P, Mathelin C, Rio MC, Levine TP, Tomasetto C. STARD3 or STARD3NL and VAP form a novel molecular tether between late endosomes and the ER. J Cell Sci 2013;126:5500–5512. 131. Li B, Vachali P, Frederick JM, Bernstein PS. Identification of StARD3 as a lutein-binding protein in the macula of the primate retina. Biochemistry 2011;50:2541–2549. 132. Vachali P, Li B, Nelson K, Bernstein PS. Surface plasmon resonance (SPR) studies on the interactions of carotenoids and their binding proteins. Arch Biochem Biophys 2012;519:32–37. 133. Tabunoki H, Sugiyama H, Tanaka Y, Fujii H, Banno Y, Jouni ZE, Kobayashi M, Sato R, Maekawa H, Tsuchida K. Isolation, characterization, and cDNA sequence of a carotenoid binding protein from the silk gland of Bombyx mori larvae. J Biol Chem 2002;277:32133–32140. 134. Soccio RE, Adams RM, Romanowski MJ, Sehayek E, Burley SK, Breslow JL. The cholesterol-regulated StarD4 gene encodes a StAR-related lipid transfer protein with two closely related homologues, StarD5 and StarD6. Proc Natl Acad Sci U S A 2002;99:6943–6948. 135. Alpy F, Tomasetto C. Give lipids a START: the StAR-related lipid transfer (START) domain in mammals. J Cell Sci 2005;118:2791–2801. 136. Rodriguez-Agudo D, Calderon-Dominguez M, Ren S, Marques D, Redford K, Medina-Torres MA, Hylemon P, Gil G, Pandak WM. Subcellular localization and regulation of StarD4 protein in macrophages and fibroblasts. Biochim Biophys Acta 2011;1811:597–606. 137. Chen YC, Meier RK, Zheng S, Khundmiri SJ, Tseng MT, Lederer ED, Epstein PN, Clark BJ. Steroidogenic acute regulatory-related lipid transfer domain protein 5 localization and regulation in renal tubules. Am J Physiol Renal Physiol 2009;297:F380–F388. 138. Soccio RE, Adams RM, Maxwell KN, Breslow JL. Differential gene regulation of StarD4 and StarD5 cholesterol transfer proteins. Activation of StarD4 by sterol regulatory element-binding protein-2 and StarD5 by endoplasmic reticulum stress. J Biol Chem 2005;280:19410–19418. 139. Horton JD, Shah NA, Warrington JA, Anderson NN, Park SW, Brown MS, Goldstein JL. Combined analysis of oligonucleotide microarray

912

140.

141.

142.

143.

144.

145.

146.

147.

148.

149.

150.

151.

152.

153.

data from transgenic and knockout mice identifies direct SREBP target genes. Proc Natl Acad Sci U S A 2003;100:12027–12032. Rodriguez-Agudo D, Ren S, Wong E, Marques D, Redford K, Gil G, Hylemon P, Pandak WM. Intracellular cholesterol transporter StarD4 binds free cholesterol and increases cholesteryl ester formation. J Lipid Res 2008;49:1409–1419. Mesmin B, Pipalia NH, Lund FW, Ramlall TF, Sokolov A, Eliezer D, Maxfield FR. STARD4 abundance regulates sterol transport and sensing. Mol Biol Cell 2011;22:4004–4015. Riegelhaupt JJ, Waase MP, Garbarino J, Cruz DE, Breslow JL. Targeted disruption of steroidogenic acute regulatory protein D4 leads to modest weight reduction and minor alterations in lipid metabolism. J Lipid Res 2010;51:1134–1143. Ishikawa T, Morris PL. Interleukin-1beta signals through a c-Jun N-terminal kinase-dependent inducible nitric oxide synthase and nitric oxide production pathway in Sertoli epithelial cells. Endocrinology 2006;147:5424–5430. Rodriguez-Agudo D, Ren S, Hylemon PB, Montañez R, Redford K, Natarajan R, Medina MA, Gil G, Pandak WM. Localization of StarD5 cholesterol binding protein. J Lipid Res 2006;47:1168–1175. Rodriguez-Agudo D, Ren S, Hylemon PB, Redford K, Natarajan R, Del Castillo A, Gil G, Pandak WM. Human StarD5, a cytosolic StAR-related lipid binding protein. J Lipid Res 2005;46:1615–1623. Rodriguez-Agudo D, Calderon-Dominguez M, Medina MA, Ren S, Gil G, Pandak WM. ER stress increases StarD5 expression by stabilizing its mRNA and leads to relocalization of its protein from the nucleus to the membranes. J Lipid Res 2012;53:2708–2715. Létourneau D, Lorin A, Lefebvre A, Frappier V, Gaudreault F, Najmanovich R, Lavigne P, LeHoux JG. StAR-related lipid transfer domain protein 5 binds primary bile acids. J Lipid Res 2012;53:2677–2689. Létourneau D, Lorin A, Lefebvre A, Cabana J, Lavigne P, LeHoux JG. Thermodynamic and solution state NMR characterization of the binding of secondary and conjugated bile acids to STARD5. Biochim Biophys Acta 2013;1831:1589–1599. Gomes C, Oh SD, Kim JW, Chun SY, Lee K, Kwon HB, Soh J. Expression of the putative sterol binding protein Stard6 gene is male germ cell specific. Biol Reprod 2005;72:651–658. Chang IY, Kim JH, Cho KW, Yoon SP. Acute responses of DNA repair proteins and StarD6 in rat hippocampus after domoic acid-induced excitotoxicity. Acta Histochem 2013;115:234–239. Lavoie HA, Whitfield NE, Shi B, King SR, Bose HS, Hui YY. STARD6 is expressed in steroidogenic cells of the ovary and can enhance de novo steroidogenesis. Exp Biol Med (Maywood) 2014;239:430–435. Taylor FR, Saucier SE, Shown EP, Parish EJ, Kandutsch AA. Correlation between oxysterol binding to a cytosolic binding protein and potency in the repression of hydroxymethylglutaryl coenzyme A reductase. J Biol Chem 1984;259:12382–12387. Nishimura T, Inoue T, Shibata N, Sekine A, Takabe W, Noguchi N, Arai H. Inhibition of cholesterol biosynthesis by 25-hydroxycholesterol is independent of OSBP. Genes Cells 2005;10:793–801.

Traffic 2014; 15: 895–914

Protein Motifs in Sterol Traffic and Metabolism

154. Jaworski CJ, Moreira E, Li A, Lee R, Rodriguez IR. A family of 12 human genes containing oxysterol-binding domains. Genomics 2001;78:185–196. 155. Lehto M, Laitinen S, Chinetti G, Johansson M, Ehnholm C, Staels B, Ikonen E, Olkkonen VM. The OSBP-related protein family in humans. J Lipid Res 2001;42:1203–1213. 156. Mesmin B, Antonny B, Drin G. Insights into the mechanisms of sterol transport between organelles. Cell Mol Life Sci 2013;70:3405–3421. 157. Ngo MH, Colbourne TR, Ridgway ND. Functional implications of sterol transport by the oxysterol-binding protein gene family. Biochem J 2010;429:13–24. 158. Olkkonen VM, Li S. Oxysterol-binding proteins: sterol and phosphoinositide sensors coordinating transport, signaling and metabolism. Prog Lipid Res 2013;52:529–538. 159. de Saint-Jean M, Delfosse V, Douguet D, Chicanne G, Payrastre B, Bourguet W, Antonny B, Drin G. Osh4p exchanges sterols for phosphatidylinositol 4-phosphate between lipid bilayers. J Cell Biol 2011;195:965–978. 160. Mesmin B, Bigay J, Moser von Filseck J, Lacas-Gervais S, Drin G, Antonny B. A four-step cycle driven by PI(4)P hydrolysis directs sterol/PI(4)P exchange by the ER-Golgi tether OSBP. Cell 2013;155:830–843. 161. Krueger KE, Papadopoulos V. Peripheral-type benzodiazepine receptors mediate translocation of cholesterol from outer to inner mitochondrial membranes in adrenocortical cells. J Biol Chem 1990;265:15015–15022. 162. Papadopoulos V, Mukhin AG, Costa E, Krueger KE. The peripheral-type benzodiazepine receptor is functionally linked to Leydig cell steroidogenesis. J Biol Chem 1990;265:3772–3779. 163. Papadopoulos V, Baraldi M, Guilarte TR, Knudsen TB, Lacapère JJ, Lindemann P, Norenberg MD, Nutt D, Weizman A, Zhang MR, Gavish M. Translocator protein (18kDa): new nomenclature for the peripheral-type benzodiazepine receptor based on its structure and molecular function. Trends Pharmacol Sci 2006;27:402–409. 164. Papadopoulos V, Berkovich A, Krueger KE, Costa E, Guidotti A. Diazepam binding inhibitor and its processing products stimulate mitochondrial steroid biosynthesis via an interaction with mitochondrial benzodiazepine receptors. Endocrinology 1991;129:1481–1488. 165. Papadopoulos V, Nowzari FB, Krueger KE. Hormone-stimulated steroidogenesis is coupled to mitochondrial benzodiazepine receptors. Tropic hormone action on steroid biosynthesis is inhibited by flunitrazepam. J Biol Chem 1991;266:3682–3687. 166. Papadopoulos V, Amri H, Li H, Boujrad N, Vidic B, Garnier M. Targeted disruption of the peripheral-type benzodiazepine receptor gene inhibits steroidogenesis in the R2C Leydig tumor cell line. J Biol Chem 1997;272:32129–32135. 167. Morohaku K, Pelton SH, Daugherty DJ, Butler WR, Deng W, Selvaraj V. Translocator protein/peripheral benzodiazepine receptor is not required for steroid hormone biosynthesis. Endocrinology 2014;155:89–97.

Traffic 2014; 15: 895–914

168. Papadopoulos V. On the role of the translocator protein (18-kDa) TSPO in steroid hormone biosynthesis. Endocrinology 2014;155:15–20. 169. Austin CJ, Kahlert J, Kassiou M, Rendina LM. The translocator protein (TSPO): a novel target for cancer chemotherapy. Int J Biochem Cell Biol 2013;45:1212–1216. 170. Qi X, Xu J, Wang F, Xiao J. Translocator protein (18 kDa): a promising therapeutic target and diagnostic tool for cardiovascular diseases. Oxid Med Cell Longev 2012;2012:162934. 171. Gatliff J, Campanella M. The 18 kDa translocator protein (TSPO): a new perspective in mitochondrial biology. Curr Mol Med 2012;12:356–368. 172. Scarf AM, Kassiou M. The translocator protein. J Nucl Med 2011;52:677–680. 173. Rupprecht R, Papadopoulos V, Rammes G, Baghai TC, Fan J, Akula N, Groyer G, Adams D, Schumacher M. Translocator protein (18 kDa) (TSPO) as a therapeutic target for neurological and psychiatric disorders. Nat Rev Drug Discov 2010;9:971–988. 174. Papadopoulos V, Lecanu L. Translocator protein (18 kDa) TSPO: an emerging therapeutic target in neurotrauma. Exp Neurol 2009;219:53–57. 175. Li H, Papadopoulos V. Peripheral-type benzodiazepine receptor function in cholesterol transport. Identification of a putative cholesterol recognition/interaction amino acid sequence and consensus pattern. Endocrinology 1998;139:4991–4997. 176. Li H, Yao Z, Degenhardt B, Teper G, Papadopoulos V. Cholesterol binding at the cholesterol recognition/interaction amino acid consensus (CRAC) of the peripheral-type benzodiazepine receptor and inhibition of steroidogenesis by an HIV TAT-CRAC peptide. Proc Natl Acad Sci U S A 2001;98:1267–1272. 177. Lacapère JJ, Delavoie F, Li H, Péranzi G, Maccario J, Papadopoulos V, Vidic B. Structural and functional study of reconstituted peripheral benzodiazepine receptor. Biochem Biophys Res Commun 2001;284:536–541. 178. Jamin N, Neumann JM, Ostuni MA, Vu TK, Yao ZX, Murail S, Robert JC, Giatzakis C, Papadopoulos V, Lacapère JJ. Characterization of the cholesterol recognition amino acid consensus sequence of the peripheral-type benzodiazepine receptor. Mol Endocrinol 2005;19:588–594. 179. Midzak A, Akula N, Lecanu L, Papadopoulos V. Novel androstenetriol interacts with the mitochondrial translocator protein and controls steroidogenesis. J Biol Chem 2011;286:9875–9887. 180. Jaremko L, Jaremko M, Giller K, Becker S, Zweckstetter M. Structure of the mitochondrial translocator protein in complex with a diagnostic ligand. Science 2014;343:1363–1366. 181. Baier CJ, Fantini J, Barrantes FJ. Disclosure of cholesterol recognition motifs in transmembrane domains of the human nicotinic acetylcholine receptor. Sci Rep 2011;1:69. 182. Fantini J, Barrantes FJ. How cholesterol interacts with membrane proteins: an exploration of cholesterol-binding sites including CRAC, CARC, and tilted domains. Front Physiol 2013;4:31.

913

Midzak and Papadopoulos

183. Midzak A, Rammouz G, Papadopoulos V. Structure-activity relationship (SAR) analysis of a family of steroids acutely controlling steroidogenesis. Steroids 2012;77:1327–1334. 184. Delavoie F, Li H, Hardwick M, Robert JC, Giatzakis C, Péranzi G, Yao ZX, Maccario J, Lacapère JJ, Papadopoulos V. In vivo and in vitro peripheral-type benzodiazepine receptor polymerization: functional significance in drug ligand and cholesterol binding. Biochemistry 2003;42:4506–4519. 185. Culty M, Li H, Boujrad N, Amri H, Vidic B, Bernassau JM, Reversat JL, Papadopoulos V. J Steroid Biochem Mol Biol 1999;69:123–130. 186. Garnier M, Dimchev AB, Boujrad N, Price JM, Musto NA, Papadopoulos V. In vitro reconstitution of a functional peripheral-type benzodiazepine receptor from mouse Leydig tumor cells. Mol Pharmacol 1994;45:201–211. 187. Campbell AM, Chan SH. The voltage dependent anion channel affects mitochondrial cholesterol distribution and function. Arch Biochem Biophys 2007;466:203–210. 188. Hauet T, Yao ZX, Bose HS, Wall CT, Han Z, Li W, Hales DB, Miller WL, Culty M, Papadopoulos V. Peripheral-type benzodiazepine receptor-mediated action of steroidogenic acute regulatory protein on cholesterol entry into leydig cell mitochondria. Mol Endocrinol 2005;19:540–554.

914

189. West LA, Horvat RD, Roess DA, Barisas BG, Juengel JL, Niswender GD. Steroidogenic acute regulatory protein and peripheral-type benzodiazepine receptor associate at the mitochondrial membrane. Endocrinology 2001;142:502–505. 190. Liu J, Rone MB, Papadopoulos V. Protein-protein interactions mediate mitochondrial cholesterol transport and steroid biosynthesis. J Biol Chem 2006;281:38879–38893. 191. Bogan RL, Davis TL, Niswender GD. Peripheral-type benzodiazepine receptor (PBR) aggregation and absence of steroidogenic acute regulatory protein (StAR)/PBR association in the mitochondrial membrane as determined by bioluminescence resonance energy transfer (BRET). J Steroid Biochem Mol Biol 2007;104:61–67. 192. Li H, Degenhardt B, Tobin D, Yao ZX, Tasken K, Papadopoulos V. Identification, localization, and function in steroidogenesis of PAP7: a peripheral-type benzodiazepine receptor- and PKA (RIalpha)-associated protein. Mol Endocrinol 2001;15:2211–2228. 193. Rone MB, Midzak AS, Issop L, Rammouz G, Jagannathan S, Fan J, Ye X, Blonder J, Veenstra T, Papadopoulos V. Identification of a dynamic mitochondrial protein complex driving cholesterol import, trafficking, and metabolism to steroid hormones. Mol Endocrinol 2012;26:1868–1882.

Traffic 2014; 15: 895–914

Binding domain-driven intracellular trafficking of sterols for synthesis of steroid hormones, bile acids and oxysterols.

Steroid hormones, bioactive oxysterols and bile acids are all derived from the biological metabolism of lipid cholesterol. The enzymatic pathways gene...
15MB Sizes 0 Downloads 3 Views