Abstract

How rapid induction of steroid hormone biosynthesis occurs in response to trophic hormone stimulation of steroidogenic cells has been a subject of intensive investigation for approximately six decades. A key observation made very early was that acute regulation of steroid biosynthesis required swift and timely synthesis of a new protein whose role appeared to be involved in the delivery of the substrate for all steroid hormones, cholesterol, from the outer to the inner mitochondrial membrane where the process of steroidogenesis begins. It was quickly learned that this transfer of cholesterol to the inner mitochondrial membrane was the regulated and rate-limiting step in steroidogenesis. Following this observation, the quest for this putative regulator protein(s) began in earnest in the late 1950s. This review provides a history of this quest, the candidate proteins that arose over the years and facts surrounding their rise or decline. Only two have persisted—translocator protein (TSPO) and the steroidogenic acute regulatory protein (StAR). We present a detailed summary of the work that has been published for each of these two proteins, the specific data that has appeared in support of their role in cholesterol transport and steroidogenesis, and the ensuing observations that have arisen in recent years that have refuted the role of TSPO in this process. We believe that the only viable candidate that has been shown to be indispensable is the StAR protein. Lastly, we provide our view on what may be the most important questions concerning the acute regulation of steroidogenesis that need to be asked in future.

Introduction: the early history

When we were asked to write this review article for the special issue of Biology of Reproduction commemorating the 50th Anniversary of the founding of the Society for the Study of Reproduction, a suggested topic for this review was “Steroidogenesis.” As steroidogenesis encompasses a very large field and would be far beyond the scope of the parameters outlined for this issue, we have rather confined the topic to a problem that the three of us have spent a considerable amount of time pursuing over the past years, namely, the acute regulation of steroid hormone biosynthesis. This is an interesting and important topic that spans a period of approximately six decades and has been the subject of intense investigation. The steroid hormones represent an important group of compounds that are synthesized mainly in the adrenal glands and the gonads, but are also synthesized in other tissues in the body. The adrenal steroids, the glucocorticoids and mineralocorticoids, are responsible for regulating carbohydrate metabolism, stress management, and salt balance. Testicular androgenic steroids are responsible for maintaining reproductive function and male secondary sex characteristics, while the ovarian progestins and estrogens are indispensable in maintaining female secondary sex characteristics and reproductive capacity. An additional class of steroids, the neurosteroids, are synthesized in the central nervous system and have specialized functions that appear to be limited to those tissues.

The common characteristic for the synthesis of steroid hormones in all steroidogenic tissues is the utilization of cholesterol as the substrate for steroid formation [1]. The first reaction is the conversion of cholesterol to pregnenolone and is a result of the action of the cytochrome P450 side-chain cleavage enzyme (P450scc; CYP11A1), that is part of the cholesterol side chain cleavage system that resides on the matrix side of the inner mitochondrial membrane (IMM) [2]. Pregnenolone then exits the mitochondria and is converted to progesterone and other steroids in the microsomal compartment of the cell. Also, in some steroidogenic cells, downstream steroids can re-enter the mitochondria and be converted to the final product dependent upon the complement of steroidogenic enzymes present within those cells [3]. These processes are regulated by the pituitary trophic hormones adrenocorticotropin (ACTH), luteinizing hormone (LH), and follicle-stimulating hormone (FSH). In the majority of steroidogenic tissues in the body these responses arise as a result of stimulation by the appropriate trophic hormone acting via the cAMP-protein kinase A (PKA) second messenger pathway. In addition, the mineralocorticoid-producing zona glomerulosa cells of the adrenal gland also respond to stimuli that transduce their signals through the Ca2+- signaling and protein kinase C (PKC) pathways. The rapid or acute effects of hormone stimulation occur within minutes and are inhibited by protein synthesis inhibitors such as cycloheximide or puromycin. The acute effects can be distinguished temporally from slower, chronic effects that occur on the order of several hours and involve mechanisms that increase gene transcription and translation of the steroid hydroxylase cytochrome P450 and the steroid dehydrogenase enzymes involved in the biosynthesis of these steroids. The characteristics of the differences between the acute and chronic regulation of steroid hormone biosynthesis have been dealt with in greater detail in several previous review articles [1, 4–7]. The current review will focus only on the mechanisms involved in the acute phase of steroid hormone biosynthesis. The steroidogenic pathway, like many metabolic synthetic pathways, has a rate-limiting step and it was experimentally determined that the regulated and rate-limiting step in steroidogenesis was the delivery of cholesterol from the outer mitochondrial membrane (OMM) to the IMM where the CYP11A1 enzyme is located [2, 8]. A significant period of investigation ensued to determine the nature of the acutely regulated step. Early investigations indicated that ACTH could stimulate the biosynthesis of steroids [9] and most importantly, that acute steroid production had an absolute requirement for the synthesis of new proteins, as mentioned above [10–14]. Further studies demonstrated that this putative regulator protein appeared to function at the level of the delivery of cholesterol to the CYP11A1 enzyme in the IMM [15–19]. Thus, the characterization of the regulated step had a specific target on which to focus, namely, a newly synthesized protein.

Over the next few decades candidate proteins for the acute regulator emerged and included the sterol carrier protein 2 (SCP2) (reviewed in [20]), the steroidogenesis-activator polypeptide (SAP) [21, 22], the peripheral benzodiazepine receptor/translocator protein (PBR/TSPO), and the steroidogenic acute regulatory protein (StAR). Each of these proteins appeared to have characteristics that rendered them as viable candidates. However, subsequent research on SCP2 knockout mice ruled out an involvement for this protein in steroidogenesis [23]. Also, while originally characterized as a small molecular weight peptide [21, 22], SAP was later identified as a fragment of the larger glucose-regulated protein 78 (GRP78) [24]. GRP78 is a regulator of the unfolded protein response [25], and it was demonstrated that GRP78 knockout mice died at an early embryonic stage [26]. Later studies examining GRP78 conditional knockouts indicated a phenotype in oncogenic signaling [27]. Shortly after these observations, studies on SCP2 and SAP significantly waned and, in contrast, StAR and TSPO continued to be studied in great detail over the next three decades.

TSPO and StAR as candidate proteins

TSPO

The peripheral benzodiazepine receptor (PBR), now called the translocator protein (hereafter referred to as TSPO), was first described in the late 1970s (see review [28]). TSPO has a high binding affinity for benzodiazepines but was a distinctly different receptor from the central benzodiazepine receptor, the γ-aminobutyric acid type A receptor/GABAA receptor. There have been many studies attempting to characterize the pharmacological and physiological functions of TSPO utilizing in vitro steroidogenic cell lines [29, 30]. TSPO has been found to be present in many different tissues, but it appeared to be most abundant in steroidogenic cells and was localized to the mitochondria in those cells [31, 32]. In 1989–1990, studies demonstrated that treatment of adrenocortical cells and testicular Leydig cells with the TSPO ligands PK11195 and Ro5–4864 could stimulate steroid synthesis [29, 33], and expansion of these studies over the next two decades indicated that TSPO ligands could increase steroid hormone production in several different steroidogenic cell types [34–36]. These observations were further corroborated when it was reported that an intracellular TSPO-binding protein, the diazepam-binding inhibitor increased steroid synthesis in steroidogenic cells [36–39]. Knockdown of TSPO expression indicated that the presence of this protein was an absolute requirement for steroid synthesis [40, 41]. Inhibition of TSPO expression could, by itself, inhibit steroid biosynthesis thus indicating its absence could prevent steroidogenesis even in the presence of other proteins (e.g. StAR). This observation was interpreted as demonstrating that TSPO acted downstream from StAR, possibly functioning as a mitochondrial membrane cholesterol channel that received intracellular cholesterol from StAR to be transported to the IMM [34, 42]. At that time, it was not possible to confirm these observations in vivo as attempts to produce TSPO knockout mice were reported to result in early embryonic lethality [43]. Another study that appeared to strongly support the role of TSPO in steroidogenesis demonstrated that steroid production in the constitutively steroid synthesizing cell line, the R2C rat Leydig tumor cell line, occurred as a result of a higher affinity TSPO ligand-binding site in these cells [44]. In later experiments, disruption of Tspo gene in R2C cells almost completely inhibited their ability to produce steroid hormones [41]. It was also reported that TSPO contained a cholesterol-binding amino acid consensus motif that was described by the authors as providing a mechanism for the binding and transport of cholesterol into the mitochondria [45]. Consideration of all these reported characteristics resulted in the authors concluding that TSPO played an indispensable role in cholesterol transfer to the IMM and thus, steroidogenesis [41, 43, 46]. This conclusion has persisted for over two decades with a number of models having been proposed as to how TSPO functioned in cholesterol transport into the mitochondria.

A more recent model that has been proposed for the role of TSPO in cholesterol transport to the IMM involves its participation in an 800 kDa protein complex [47]. Proteins in this putative complex, that was named the “transduceosome,” included TSPO, the voltage-dependent anion channel 1 (VDAC1), P450scc, the ATPase family AAA domain-containing protein 3A (ATAD3A), and optic atrophy type 1 proteins. Knockdown of ATAD3A or VDAC, considered to be part of this complex, resulted in an inhibition of steroid synthesis. This model also indicated that StAR could increase steroid synthesis by mobilizing cholesterol that is bound to TSPO “polymers” present in the putative complex. Problems with this model are that native PAGE is highly prone to artifacts when dealing with transmembrane proteins, as this would include membrane vesicles that contain proteins without any rationale for associations. By providing only a short selected list of proteins without disclosing the full proteomics results [47], the study lacks necessary rigor for meaningful differentiation fact from artifact. Therefore, the existence of such a “transduceosome” complex is highly questionable. Moreover, the decrease in steroid biosynthesis with knockdown of ATAD3A and VDAC1 could be secondary to overall mitochondrial and cellular health, as transmission electron microscopy images in the same study showed severe structural disruption to mitochondria after ATAD3A and VDAC1 knockdown in MA-10 Leydig cells [47]. Unfortunately, no knockdown of TSPO was included in these experiments so assessing its absolute requirement or its function in this putative complex mediating cholesterol transfer and steroidogenesis could not be determined. Nevertheless, this model continued to receive attention and has been summarized in some recent reviews [48–50].

StAR

Orme-Johnson and colleagues originally described several rapidly induced ACTH and LH 30 kDa phosphoproteins in adrenocortical cells, and in rat corpus luteum and mouse Leydig cells, respectively [51–58]. These studies demonstrated a close relationship between the appearance of the 30 kDa proteins and steroid hormone biosynthesis and that their synthesis was cycloheximide sensitive. Our laboratory was engaged in similar studies in hormone-stimulated MA-10 mouse Leydig tumor cells and observed proteins which were identical to those described by Orme-Johnson [59–65]. These proteins were localized to the mitochondria and consisted of several different isoforms of a 30 kDa protein processed from a 37 kDa precursor protein that contained an N-terminal mitochondrial signaling sequence [53, 61]. Subsequent studies demonstrated close correlations between the synthesis of steroids and the synthesis of the 30 kDa mitochondrial proteins and thus, they represented reasonable candidates for the regulatory protein [51–65]. Purification of the 30 kDa protein, cloning of the cDNA for the precursor, and its sequencing were successfully accomplished in 1994 [66]. The nucleic acid sequence of the cDNA and the amino acid sequence of the 37 kDa protein were unique, indicating that it represented a previously undescribed protein. Expression of the 37 kDa protein in MA-10 mouse Leydig tumor cells resulted in a significant increase in steroid production in the absence of hormone stimulation. Also, expression of the 37 kDa protein in nonsteroidogenic COS-1 cells resulted in a several fold increase in pregnenolone production [67–70]. These results produced a “cause and effect” relationship between the 37–30 kDa proteins and steroid production and indicated a direct role for these proteins in hormone-regulated steroid production. As a result of these observations, this protein was named the steroidogenic acute regulatory protein or StAR [66].

StAR is composed of a single functional domain with an α/β helix-grip fold structure containing a nine-stranded anti-parallel β-sheet forming a long hydrophobic cleft that binds cholesterol. The amino acid sequence for the StAR functional domain was used to identify a protein family, called the START domain family (START for StAR-related lipid-transfer domain) [72]. START domains bind hydrophobic lipids and StAR belongs to the mammalian START subfamily STARD1. The STARD1 subfamily has two members: StAR (STARD1) and MLN64 (STARD3). STARD3 was identified as an amplified transcript in breast cancer-derived metastatic axillary lymph nodes (MLN), and named MLN64 reflecting it was clone #64 in the screen [71]. MLN64 was the first reported protein to have a domain that shared 33% sequence identity and 53% sequence similarity with human StAR [72, 73]. STARD3 has been implicated to function in a StAR-independent mechanism to deliver cholesterol to mitochondria for steroidogenesis. Specifically, STARD3 is proposed to stimulate progesterone production by the human placenta, the one steroidogenic tissue that lacks StAR. STARD3 is a multidomain START protein with amino terminal MENTAL (MLN64 N-terminal) domain followed by a FFAT (diphenylalanine (FF)–acidic track (AT)) motif, and the carboxyl-terminal START domain. The MENTAL domain is composed of four transmembrane helices that localize STARD3 to the late endosomes and orients the START domain toward the cytoplasm. STARD3 overexpression is capable of enriching cholesterol in the mitochondria [74] or plasma membrane [75], depending on the cell type studied. The expression of only the START domain of MLN64/STARD3 is capable of stimulating progesterone synthesis in transfected COS-1 cells [76]. Again, the significance of this finding may be linked to placental steroidogenesis where high levels of progesterone are synthesized from cholesterol in the absence of StAR. In human and rodent tissues, MLN64 was detected in placenta and the START domain is processed from the larger protein, providing a possible mechanism for cholesterol transport into mitochondria in this tissue [77]. The START domain of STARD3 may function with a mitochondrial heat-shock protein to promote cholesterol transfer into mitochondria [78]. However, mice that express a STARD3 form that lacks the START domain are viable and fertile [79]; therefore, additional studies are warranted to determine the mechanism controlling cholesterol delivery into mitochondria for placental steroidogenesis.

Following the cloning of the StAR cDNA, collaborative studies with Dr W Miller and Dr J Strauss III demonstrated that mutations in the StAR gene resulted in congenital lipoid adrenal hyperplasia (lipoid CAH) [68]. Lipoid CAH is a lethal condition characterized by a nearly complete inability of the newborn to synthesize steroids. These patients have enlarged adrenals containing high levels of cholesterol and cholesterol esters and an increase in the amount of lipid accumulation in testicular Leydig cells indicating an inability to convert cholesterol to pregnenolone. In females, the fetal ovary makes little or no steroids after the first trimester [80]. Therefore, the ovary appears to be normal until the time of puberty when it is stimulated with LH for the first time. Without StAR action, the amount of estrogen synthesized by StAR-independent action in the ovary is greatly reduced while cholesterol begins to accumulate. The suboptimal level of estrogen does result in a marginal amount of feminization that includes some breast development and vaginal bleeding [81–83]. However, the continued accumulation of cholesterol in the latter part of the cycle in the ovary eventually impairs the synthesis of progesterone and results in anovulatory cycles [83]. A recent review provides an excellent summary of the mutations in StAR that have been uncovered as of the present time [84]. These observations added compelling evidence for the essential role of this protein in the regulation of steroidogenesis since, in essence, lipoid CAH provides a human StAR knockout and the phenotype has the expected characteristics. Collaborative studies of our laboratory with Dr K. Parker and colleagues demonstrated StAR mRNA expression was tightly correlated with the appearance of steroidogenic cells and the timing of steroidogenesis in the adrenal glands and testis during embryonic development in the mouse [85]. These studies clearly demonstrated the presence of StAR transcripts in the adrenal cortex and testicular Leydig cells, but as expected, there were no StAR transcripts in the ovary during development, as estrogen is not produced in this organ until puberty, as also seen in the human female. In another study, also in collaboration with the Parker laboratory, targeted disruption of the StAR gene in mice was used to successfully produce StAR null mice [86]. StAR knockout mice had female external genitalia (regardless of sex), failed to grow normally, and died within a short period of time, presumably as a result of adrenocortical insufficiency. Serum levels of corticosterone and aldosterone were depressed, while levels of ACTH and CRH were elevated indicating impaired production of adrenal steroids with an accompanying loss of feedback regulation at the level of the hypothalamus or pituitary. The adrenal glands had an abnormal cortex that displayed elevated lipid deposits in the knockout mouse as did the Leydig cells in the testis. As seen in the human, the ovaries in the StAR knockout mice were unaffected at the time of birth, but continued to accumulate lipids as the animals entered puberty [87]. Thus, the StAR knockout mouse demonstrated characteristics similar to human lipoid CAH and further substantiated the necessity for StAR action in steroid biosynthesis.

With the availability of StAR reagents, results obtained with methodologies such as western and northern analysis, in situ hybridization, immunocytochemistry, RNAse protection assays, and RT PCR have all been used to demonstrate that StAR expression is essentially confined to steroidogenic tissues. Importantly, these studies were conducted in many different laboratories and successfully demonstrated that StAR was present in the steroidogenic cells of steroidogenic tissues. StAR expression has been demonstrated in adrenal cortical layers, adrenal tumors, ovarian theca cells, ovarian granulosa cells, ovarian corpora lutea cells, fetal mouse giant trophoblast cells, and testicular Leydig cells [88–94]. Using more sensitive technologies StAR protein has also been detected in a wider variety of tissues where its roles are, so far, not completely known [95, 96].

Characterizing the role of StAR in cholesterol transport to the IMM has proven to be refractory to elucidation even after more than two decades of effort. We originally hypothesized that import of StAR protein into the mitochondrial matrix temporarily formed contact sites between the OMM and IMM allowing the hydrophobic cholesterol to transfer between the membranes [67]. Based on the first report for a crystal structure of a START domain for STARD3/MLN64 that has a conserved structure with StAR and also binds cholesterol, Tsujishita and Hurley proposed that StAR acted as an intramitochondrial membrane shuttle to mediate cholesterol transport to the IMM by binding a one molecular of cholesterol at a time [97]. It was also proposed that StAR might alter the molecular conformation of the mitochondrial membranes to allow for the passage of cholesterol to the IMM, before it is transported to the mitochondrial matrix (reviewed in [98]). However, StAR action had to be reconsidered when it was shown that StAR need not enter the mitochondria to be active and that interactions at the OMM were sufficient for cholesterol transfer. The report that deletion of 62 amino acids at the N-terminus of StAR prevented mitochondrial import, but did not affect cholesterol transfer and steroid production [99, 100], was supported by studies that showed a TOM20-StAR fusion protein that was located at the OMM due to TOM20 and could not enter the matrix but could still induce steroidogenesis [101]. In an effort to explain what might be occurring at the level of the OMM, Dr Walter Miller and colleagues performed a series of biophysical studies, which demonstrated that StAR undergoes a conformational alteration when it binds to cholesterol [102]. This alteration, caused by an acid-induced breaking of hydrogen bonds, appears to be required for StAR activity and is a result of StAR’s transition to a molten globule form [102–104]. Biochemical studies suggested that StAR interacts with a protein complex that promotes cholesterol transfer from the OMM to the IMM. Recent studies have revealed possible specific involvement of the VDAC2 and TOM22 for StAR activity at the OMM membrane [105, 106]. However, there has yet to arise a model that adequately explains the mechanism of StAR in the transfer of cholesterol to the IMM. And future studies addressing StAR’s mechanism of action should include the role of StAR phosphorylation. A PKA-dependent phosphorylation at S194 (mouse) or S195 (human) is critical for StAR’s function [107, 108], reviewed in [109]). This was best demonstrated using an in vivo approach wherein re-expression of wild-type StAR in StAR knockout mice restored adrenal and testis steroid production but mice expressing the StAR-S194A transgenic displayed lipid accumulation in the adrenal and testis, and low corticosterone and testosterone levels similar to StAR knockout mice [110]. Furthermore, it is likely that StAR transcription, translation, and phosphorylation are coordinated processes in the acute response to move cholesterol into the mitochondria. . In particular, data support StAR mRNA is associated with mitochondria and the mitochondrial A-kinase anchoring protein 121 (AKAP121) and type II PKA (PKAR2) [111–114]. Newly transcribed StAR mRNA is proposed to associate with AKAP121 based on the observation that a StAR mRNA-AKAP121 interaction occurred only after Bt2cAMP treatment of H295R human adrenocortical cells to increase StAR mRNA levels [114]. StAR mRNA levels but not protein levels are increased after Bt2cAMP treatment of MA-10 cells that lack AKAP121, suggesting that StAR translation is diminished when the mRNA is not localized to the mitochondria. The StAR mRNA-AKAP121-PKAR2 complex at the mitochondria warrants further examination to determine whether rapid phosphorylation of StAR S194/195 at the OMM occurs, and whether this phosphorylation promotes specific interactions with OMM/IMM proteins such as VDAC2 or TOM22.

Using the TOM20-StAR fusion protein, work performed by Dr Papadopoulos and colleagues demonstrated that knockdown of TSPO using antisense oligonucleotides resulted in an inhibition of steroid synthesis [46]. These findings suggested cooperation between StAR and TSPO mediated mitochondrial cholesterol transport and proposed a model in which StAR carried cholesterol from cellular stores to the OMM where TSPO acted as a protein tunnel for the import of cholesterol to the IMM [42].

As further investigations added additional corroborations to the roles of TSPO and StAR in the transfer of cholesterol to the IMM in support of steroidogenesis, the mechanism whereby this occurred remained a mystery and still does. However, a major shift in this field occurred in 2014 and beyond with several reports that seriously questioned the role of TSPO in the acute regulation of steroidogenesis. The following section will summarize those studies.

TSPO is not involved in mitochondrial cholesterol import for steroidogenesis

In this section, rather than describe events in chronology, we present the most compelling pieces of evidence from precise genetic models that indicate that TSPO is not involved in steroidogenesis. In MA-10 Leydig cells, complete disruption of the Tspo gene using CRISPR/Cas9-mediated targeting had no effect on the extent of their ability to synthesize steroid hormones [115]. This means that in the complete absence of TSPO protein, mitochondrial cholesterol import progressed with absolutely no impediments. Of the three TSPO-knockout subclones generated, one even produced significantly higher levels of steroids compared to TSPO-intact controls. Without observing any loss of function, this result was in direct contrast to previous reports that seemed to demonstrate an indispensable role for TSPO in mitochondrial cholesterol import [41, 43, 46]. Two fundamental observations reported by this group were as follows: (1) there occurs an almost complete elimination of steroid biosynthesis in R2C Leydig cells after Tspo gene disruption [41]; (2) knockdown of TSPO in MA-10 cells not only inhibited their ability to synthesize steroids but also prevented StAR import into mitochondria [46], suggesting that TSPO was essential for StAR function. Not only did we find it extremely hard to reconcile recent findings with these polar opposite reports, but also recent in vivo observations in Tspo-deleted mice clearly supported recent in vitro results that TSPO was not involved in mitochondrial cholesterol import for steroid hormone biosynthesis.

Conditional Tspo deletion in testicular Leydig cells did not affect testosterone levels [116]. Global Tspo deletion resulted in viable fertile mice with no effects on testosterone, estrogen, progesterone, corticosterone, and aldosterone levels [117]; these TSPO knockout mice had no apparent abnormalities. A second independently generated global Tspo deleted mouse model confirmed this viable phenotype, and also demonstrated that mitochondria isolated from mice lacking TSPO had no defects in cholesterol import function [118]. A third independently generated global Tspo deleted mouse model reconfirmed this viable phenotype with no apparent abnormalities [119]. These three recent independent reports dispute the phenotypic outcome previously reported in a review article that, “Efforts to generate a PBR-negative gene knockout mouse model failed, as the animal died at an early embryonic stage suggesting that PBR is involved in basic functions necessary for embryonic development” [43]. Similarly, at the cellular level, the absolute need for TSPO in maintaining cell viability was again emphasized in review articles with claims that >70% knockdown of TSPO in cultured cells lines resulted in cell death [120, 121]. The basis for these assertions that were provided as precluding the study of steroidogenesis using gene-deleted models remains open to question, particularly because the experimental details were never published as research articles. We have carefully reviewed the history of TSPO research that includes all of the results that underlie the sustained consideration of TSPO in the mitochondrial cholesterol import model [122–124].

The Papadopoulos group subsequently generated the same Amhr2cre/+TspocΔ/Δ mouse and reported that TSPO was crucial for viability and hormone-dependent adrenal steroid formation [125]. We respectfully disagree with this interpretation [126]; the low rate of Amhr2cre/+TspocΔ/Δ reported as being due to embryonic lethality in this manuscript is in fact due to linkage between the Tspo and Cre loci (18.18 centimorgans apart in chromosome 15) [126]. Due to this linkage, independent assortment is not possible and there needs to be chromosomal crossover between the two loci to generate Amhr2cre/+TspocΔ/Δ mice. As this can occur only at a low rate (calculated: 7.6%; observed: 4.4%), it is not indicative of embryonic mortality as classical Mendelian principles are not applicable in this context [126]. Recently, this same group extended efforts to revive a mitochondrial cholesterol transport function for TSPO in rat and human steroidogenic cells [127]; it is essential that this and similar studies receive independent validation.

Another facet to the TSPO story can be found in the field of drug development targeting this protein. This continues to be quite extensive due to significant commercial interest in small molecules that bind TSPO, mainly for diagnostic imaging of inflammatory lesions that overexpress TSPO, and testing its effects on immunomodulation [24 clinical trials in progress or completed in the USA and EU combined (Source: clinicaltrials.gov and clinicaltrialsregister.eu)]. The core mechanism of TSPO-binding drug effects on immunomodulation has been believed to be due to its putative role in steroid biosynthesis. In the existing >400 publications on TSPO-binding drug development, most studies have used the steroidogenesis model to explain outcomes without direct examination of molecular aspects of function resulting from TSPO binding. Studies in adrenocortical and Leydig cells that report effects for TSPO-binding drugs in steroidogenesis [29, 33] are often referenced; however, the pharmacology upon which this phenomenon is based is difficult to interpret (reviewed in [123]). Rigorous target validation was not considered in these studies as TSPO knockout cells were incorrectly considered to be nonviable. Recently, by using TSPO knockout Leydig cells, it was demonstrated that the effect of one prototypical TSPO-binding drug PK11195 in transient steroidogenesis, as previously demonstrated [33], persisted in the complete absence of TSPO [115]. This result is highly significant because PK11195 was used to suggest that TSPO mediates mitochondrial cholesterol import [34]; PK11195 has been used consistently to demonstrate an induction of steroidogenesis in both Leydig [33] and adrenocortical cells [29]. The finding that PK11195-induced steroidogenesis is independent of TSPO [115] provided the first evidence that the entire pharmacological basis for TSPO and steroidogenesis is likely a misinterpretation. Moving forward, it is reassuring that TSPO knockout models will allow for rigorous target validation for the putative TSPO ligands and drug discovery.

Although there are more than 200 publications on TSPO, most of these studies do not mechanistically address mitochondrial cholesterol import but rather explain the multiple effects of TSPO pharmacology by the supposition that it is involved in steroidogenesis. Aside from pharmacology, it is accurate to state that all mechanistic results specific for TSPO involvement in mitochondrial cholesterol import were reported by a single research group [37, 39–41, 44–46, 128]. The only independent study on this topic was performed by Gavish in 1998, where it was observed that results connecting TSPO knockdown and steroidogenesis were inconclusive; no acute effects were observed [129]. Gavish and colleagues concluded that “further studies are needed to confirm the involvement of the 18-kDa PBR subunit in MA-10 Leydig cell steroid biosynthesis” [129].

For complete point-by-point examination of the history and evolution of the TSPO-steroidogenesis model and its links to physiological expression, pharmacology, binding proteins, cholesterol binding/transport, structure, genetic deletion, and human TSPO gene polymorphisms, see the extensive examination of literature presented in other recent reviews [122–124, 130]. Regrettably, the sheer volume of TSPO publications referencing steroidogenesis will continue to present a challenge to researchers, particularly because the physiological function of TSPO remains unestablished [131–133]. Nevertheless, it is encouraging that progress in the TSPO field, distinct from steroidogenesis research, has begun to indicate functions for TSPO particularly in fatty acid metabolism [134] and other aspects of cellular energy production [135]. These may indeed be connected to indirect effects on steroidogenic homeostasis independent of mitochondrial cholesterol import in vivo as reported in a recent study using global Tspo/ mice when all steroidogenic intermediates were profiled [136].

Regulation of STAR/Star gene expression

STAR/Star gene expression in the adrenal cortex and gonads is under the control of the tropic hormones ACTH and the gonadotropins LH and FSH, respectively. ACTH, LH, and FSH bind to their cognate 7-transmembrane G-protein coupled receptors leading to activation of cAMP-dependent PKA, phospholipase C–diacylglycerol-PKC-inositol 1,4,5 trisphosphate (PLC-DAG-PKC-IP3), mitogen activated kinase (MAPK), and calcium signaling pathways. Activation of the PLC-DAG-PKC-IP3 and calcium signaling pathways is also the mechanism of action for angiotensin II (Ang II) and K+-stimulated aldosterone synthesis in the adrenal zona glomerulosa. Ang II activates a G protein-coupled receptor, while K+ activates voltage-gated calcium channels and subsequent calcium-PKC-signaling pathways. Maximal increases in steroidogenesis are observed when there is cross-talk between these pathways, yet activation of the cAMP-PKA signal transduction pathway is the major player in tropic hormone-dependent activation of steroid output. Over the past two decades the studies on STAR/Star gene regulation have uncovered that species-specific and cell-type-specific mechanisms control StAR expression due to differences in promoter sequence and differences in transcription factor activation and recruitment to the cAMP responsive region the STAR/Star proximal promoter [137–139]. Herein we focus on the sequence differences between the rodent Star and human STAR promoter to highlight how multiple trans-acting factors have been identified that work through a common promoter region to increase transcription in a species- and tissue-specific manner.

STAR/Star gene regulation is characteristic of the immediate early response genes [140]: StAR mRNA levels are increased acutely (within minutes) following tropic hormone stimulation, transcription is independent of protein synthesis (cycloheximide-insensitive), and expression is transient (transcription terminates upon removal of the stimulus). In addition, the STAR/Star promoter contains a TATA-box, GATA-4-, and cAMP responsive cis-acting elements, which are commonly found in highly regulated genes. The region of the STAR/Star proximal promoter 150 bp region immediately upstream of the transcription start site is highly conserved and is sufficient for maximal transcriptional activation of mouse, human, and rat StAR genes in the adrenal and gonads [139]. This region contains the cAMP-responsive element (CRE) which is a nonconsensus sequence that binds cAMP response element binding protein (CREB)/CREB-modulator (CREM) family members and activator protein-1 (AP-1) family members. Therefore, this element is referred to as the CRE/AP-1 site [Figure 1A] [137, 141–144]. CREB/CREM/AP-1 are members of the basic leucine zipper (bZIP) family of transcription factors due to the bZIP motif that drives homo- and heterodimerization among family members. AP-1 is composed of members of the Fos family (c-Fos, FosB, Fra-1, and Fra-2) that heterodimerize with members of the Jun family (c-Jun, JunB, and JunD) as well as with some members of the CREB/ATF family, while Jun members can also function as homodimers [145]. The multiple heterodimer partners and the relative levels of CREB v. CREM v. AP-1 proteins contribute to the cAMP-dependent regulation of STAR/Star gene expression in a species or cell-type dependent mechanism. For example, CREB phosphorylation and recruitment to the Star CRE/AP-1 site in mouse Leydig cell cultures mediates the cAMP response yet in the adrenal CREB is not expressed and CREM isoforms bind to the CRE/AP-1 element and activate StAR transcription [146–148]. c-Jun was also shown to be a potent trans-activator of Star transcription in MA-10 mouse Leydig cells via binding to the CRE/AP-1 site. It is likely that c-Jun/CREB heterodimer formation explains the recruitment of these two factors to the same element in response to cAMP-PKA signaling. AP-1 family members JunB-Fos/FosB or JunB-ATF3 also are capable of stimulating STAR promoter-reporter gene expression in H295R human adrenocortical cells, most likely via binding to the CRE/AP-1 element [149]. Importantly, PKA-dependent phosphorylation of GATA-4, CREB/CREM, and/or c-Fos/c-Jun (AP-1) enhances both the recruitment of these factors to the STAR promoter and protein–protein interactions between these factors ([150], reviewed in [138, 151]). The result is enhanced recruitment of the coactivator CREB-binding protein (CBP) and RNAPII to the STAR promoter. However, the cAMP response is dependent on the function of SF-1 and GATA-4 (Figure 1B).

Transcriptional and post-transcriptional mechanisms that control the cAMP-PKA-dependent regulation of STAR expression. (A) Shown is the sequence comparison for the cAMP-responsive region of the mouse and human StAR promoter with the SF-1, CRE/AP-1, and GATA-4 elements are shown. The CRE/AP-1 is a nonconsensus CRE sequence that binds cAMP response element binding protein (CREB)/CREB-modulator (CREM) family members and activator protein-1 (AP-1) family members. (B) The left panel shows a model for maintaining low basal StAR steady-state mRNA levels. Transcriptional repression may be mediated by one or all of the following: DAX-1 and COUP-TF blocking SF-1 action, low interaction of CREB/AP-1/GATA-4 with the StAR promoter, and diminished CREB action due to lack of TORC2 interactions. Post-transcriptional repression may be mediated by promoting StAR mRNA degradation by either TIS11b or miRNAs. The right panel shows the mechanisms stimulated by cAMP-PKA signaling. Tropic hormones ACTH, LH, and FSH bind their cognate G-protein coupled receptors and activate adenylyl cyclase (AC) thereby increasing cAMP levels and promoting PKA-dependent downstream effects. Depicted is the phosphorylation of CREB/CREM, c-Fos/c-Jun (AP-1), and GATA-4 and recruitment of these transcription factors to their respective DNA elements in the StAR proximal promoter. SF-1 may be constitutively bound to the promoter with increased association after cAMP-PKA activation. Protein–protein interactions between GATA-4 and SF-1, AP-1, and CREB are proposed to stabilize the transcription complex and enhance recruitment of the coactivator CBP and RNAPII. Additional activation is mediated by loss of SIK activity and increased TORC2 association with CBP as described in the text. TIS11b levels are induced by cAMP-PKA signaling and may play a role in attenuating the acute response. The role of miRNAs in controlling StAR mRNA levels in hormone-dependent manner remains to be examined.
Figure 1.

Transcriptional and post-transcriptional mechanisms that control the cAMP-PKA-dependent regulation of STAR expression. (A) Shown is the sequence comparison for the cAMP-responsive region of the mouse and human StAR promoter with the SF-1, CRE/AP-1, and GATA-4 elements are shown. The CRE/AP-1 is a nonconsensus CRE sequence that binds cAMP response element binding protein (CREB)/CREB-modulator (CREM) family members and activator protein-1 (AP-1) family members. (B) The left panel shows a model for maintaining low basal StAR steady-state mRNA levels. Transcriptional repression may be mediated by one or all of the following: DAX-1 and COUP-TF blocking SF-1 action, low interaction of CREB/AP-1/GATA-4 with the StAR promoter, and diminished CREB action due to lack of TORC2 interactions. Post-transcriptional repression may be mediated by promoting StAR mRNA degradation by either TIS11b or miRNAs. The right panel shows the mechanisms stimulated by cAMP-PKA signaling. Tropic hormones ACTH, LH, and FSH bind their cognate G-protein coupled receptors and activate adenylyl cyclase (AC) thereby increasing cAMP levels and promoting PKA-dependent downstream effects. Depicted is the phosphorylation of CREB/CREM, c-Fos/c-Jun (AP-1), and GATA-4 and recruitment of these transcription factors to their respective DNA elements in the StAR proximal promoter. SF-1 may be constitutively bound to the promoter with increased association after cAMP-PKA activation. Protein–protein interactions between GATA-4 and SF-1, AP-1, and CREB are proposed to stabilize the transcription complex and enhance recruitment of the coactivator CBP and RNAPII. Additional activation is mediated by loss of SIK activity and increased TORC2 association with CBP as described in the text. TIS11b levels are induced by cAMP-PKA signaling and may play a role in attenuating the acute response. The role of miRNAs in controlling StAR mRNA levels in hormone-dependent manner remains to be examined.

Steroidogenic factor-1 (SF-1) and GATA-4 elements flank the CRE/AP-1 site, and both SF-1 and GATA-4 binding are important for full promoter activity. The minor sequence differences between mouse and human SF-1 and CRE/AP-1 DNA elements influence the preference for transcription factor binding (Figure 1A). Mutation of the SF-1 element to block SF-1 binding located directly upstream of the CRE/AP-1 element resulted in loss of promoter activity for human STAR gene but not for mouse Star gene. This may be explained by the differences in the flanking DNA sequences between the species; the mouse Star promoter contains an overlapping sequence that binds NUR77 (NGFI-B (nerve growth factor induced-B) while human STAR promoter lacks this binding site [152–154]. Thus, SF-1 binding to this proximal element is essential for human promoter activity but not required for mouse promoter activity, likely due to NUR77 compensating for SF-1 function in the rodent [146, 155, 156]. The Star proximal promoter has also a highly conserved CCAAT box element that overlaps the CRE/AP-1 site and C/EBPβ can bind to this site as demonstrated by electrophoretic mobility shift assays [150, 157, 158]. In mouse Leydig and granulosa cells, activation of cAMP-PKA signaling was shown to increase C/EBPβ expression [159, 160], and C/EBPβ binding to the Star proximal promoter was shown to be important for transcription in mouse granulosa-luteal cells [161]. In summary, SF-1-pGATA-4 binding appears to be a common mechanism for mouse and human STAR/Star gene activation in multiple cell types while the homo- or heterodimer partners between CREB/CREM and AP-1 family members are dependent upon the species, cell type, and stimulus (reviewed in [137]).

Cross-talk between signaling pathways provides a fine control for StAR expression and steroidogenesis. Pharmacological inhibition of JAK2 signaling was shown to inhibit angiotensin II-stimulated StAR expression and steroid output in H295R adrenocortical cells and ACTH-stimulated increased corticosterone synthesis in the mouse adrenocortical cell line (ATC-1) or primary cultures of rat adrenocortical cells [162, 163]. STAR-promoter-reporter gene assays showed that STAT5 did not activate STAR transcription, rather JAK2 activity was necessary for increasing pCREB stability that led to trans-activation of the STAR gene [162]. This finding highlights the central role that pCREB plays in STAR/Star transcription. ACTH can also act indirectly to enhance StAR expression and cortisol production. In H295R human adrenocortical cells, ACTH stimulated sphingosine-1-phosphate (S1P) secretion that works in an autocrine manner to promote cortisol synthesis via increased expression of STAR as well as other genes involved in steroidogenesis [164, 165]. S1P stimulation was mediated via binding to a GPCR membrane receptor linked to activating the Ca2+-CamKII-ERK1/2 signaling pathway. MAPK/ERK activation promotes StAR phosphorylation so it is likely that S1P may enhance both expression and function of StAR.

Transcriptional repression also controls STAR/Star gene expression, although fewer studies have been devoted to this mechanism. Repressor proteins work via direct and indirect mechanisms: they bind DNA and recruit corepressor complexes (direct), or they block a trans-activator protein from binding DNA (indirect), thereby blocking assembly of a transcription initiation complex. Dax-1 (dosage sensitive sex reversal-adrenal hypoplasia congenital gene on the X chromosome gene 1), a member of the nuclear receptor superfamily, was the first reported repressor of STAR transcription. DAX-1 contains a strong C-terminal repressor domain yet lacks a DNA-binding domain so mediates its repressive action through indirect mechanisms [166]. For example, DAX-1 binds to a hairpin loop formed within promoters, including the STAR promoter [167, 168], which prevents SF-1 and/or NUR77 binding. In addition, direct SF-1-DAX-1 interactions that tether DAX-1 to the promoter via DNA-bound SF-1 results in recruitment of corepressor proteins and transcriptional repression of StAR [166, 168–170]. COUP-TF (chicken ovalbumin upstream promoter transcription factor I and II) is also capable of blocking SF-1 function through competitive binding to SF-1 elements in the STAR promoter [171–174]. Thus, both DAX-1 and COUP-TF can function to interfere with SF-1 activation of the Star promoter and represses transcription. One possible mechanism for switching from repression to activation is hormone-dependent activation of PKA resulting in decreased DAX-1 or COUP-TF expression, thereby allowing SF-1 and other factors to activate STAR/Star gene expression (Figure 1B). In several studies an inverse relationship between repressor protein(s) and StAR expression levels was observed [171, 175–177]. An interesting example of loss of repressor function involves an indirect mechanism that leads to enhancing CREB-mediated Star gene expression in the adrenal. In brief, CREB regulated transcription coactivator 2 (CRTC2), also known as co-activator transducer of regulated CREB activity (TORC2), promotes the recruitment of CBP to pCREB on target gene promoters (60). TORC2 is phosphorylated by salt-inducible kinase-1 (SIK1), an AMP-kinase family member. Phosphorylated TORC2 (pTORC2) translocates to the cytoplasm; therefore, active SIK1 represses CREB-mediated transcription via diminishing nuclear TORC2. SIK1 was first characterized as a high-salt diet-induced transcript in rat adrenals [178] and later shown to be an ACTH-responsive transcript in Y1 mouse adrenocortical cells [179]. In Y1 mouse adrenocortical cells, low basal StAR expression is proposed to be maintained, in part, by steady-state levels of active nuclear SIK1 and cytoplasmic pTORC2 that contribute to suppressed Star gene expression [63, 67]. To this end, ACTH stimulation results in SIK1 phosphorylation and inactivation. In vivo studies support a temporal correlation between detection of TORC2 and pCREB in the nucleus that precedes an increase in detection of StAR transcription. Thus, one arm of the ACTH-cAMP-PKA pathway leading to acute control of Star transcription is the loss of SIK1 repressor action that allows for maximal pCREB-mediated transcription in the adrenal (Figure 1B) [63, 67].

In addition to transcriptional mechanisms, post-transcriptional regulation of StAR expression has been reported. It was established in early studies by northern blot analysis that two StAR transcripts are generated in a similar temporal time frame in response to cAMP-PKA signaling. A long 3.5 kb transcript and a short 1.6 kb transcript arise due to processing at different polyadenylation sites in the 3΄-UTR. The short transcript is expressed at much lower levels relative to the longer transcript, although the longer transcript is subject to faster degradation. The RNA-binding protein TIS11b (zinc finger protein, Znf36L1, also known as BRF1) binds to an AU-rich destabilizing element present only in the 3.5 kb StAR 3΄-UTR [180] (reviewed in [181]). The significance for targeting this particular transcript for degradation remains to be determined, yet it may represent a mechanism to attenuate the hormonal response [182]. Similar to SIK1, TIS11b is also induced by cAMP-PKA signaling in adrenal cells and an increase in both SIK1 and TIS11b would suppress StAR expression at both the transcriptional and post-transcriptional level. In this manner, the cAMP-PKA signaling remains central to modulating StAR expression and controlling tropic hormone-mediated steroidogenesis.

The role for miRNA-mediated post-transcriptional control of StAR expression in a hormone-dependent manner is a relatively new area of study. One of the challenges for future investigations of miRNAs in StAR regulation is identifying and validating functional miRNA-StAR interactions. Searching TargetScan7, miRDB, miRanda, and miRTarBase databases to identify predicted miRNA binding sites for mouse and human StAR reveals many poorly conserved sites. Studies that have characterized the differentially expressed miRNAs in Leydig cells and the adrenal have not yet revealed regulatory miRNAs that target StAR. In TM3 mouse Leydig cells after LH treatment, a total 2904 miRNAs were detected by deep sequencing, with 27 miRNAs upregulated and 2 downregulated by LH treatment [183]. No list was provided for the miRNAs that were detected in the TM3 Leydig cells, although Let-7b-3p was validated as an upregulated (threefold) miRNA. In contrast, Let-7b-3p was identified using a miRNA microarray approach as a downregulated miRNA in rat adrenal following ACTH treatment [184]. Let-7b is one of the first miRNAs proposed to regulate StAR. A putative miR-Let-7b binding site was identified in the human StAR 3΄-UTR and overexpression of miR-let-7 in HEK-293 cells reduced the activity of a luciferase reporter gene containing a short segment of the human StAR 3΄-UTR [185]. The authors also show that miR-let-7 overexpression in MLTC-1 mouse Leydig tumor cells decreased endogenous StAR mRNA ∼20%, although the mouse StAR 3΄-UTR was never tested. The focus on miR-let-7 in this study was an attempt to explain the observation that overexpression of the long noncoding RNA H19 promotes increased StAR mRNA and protein expression in both human KGN cells and mouse MLTC-1 cells. H19 was previously shown by this group to bind miR-let-7 and block its action [186]. However, StAR was never validated as a bona fide target of miR-let-7b in this study, which requires additional studies to demonstrate that mutation of 3΄-UTR of human (and mouse) StAR abolishes the miRNA-mediated repression. The finding that Let-7b-3p is regulated by cAMP-PKA signaling in two steroidogenic cell types supports a more robust testing for this miRNA in controlling StAR and steroidogenesis in cell-specific context is needed. Another exciting possibility is H19 might function as a positive regulator of StAR transcription and efforts directed at H19 action in ovarian steroidogenesis seems warranted.

Lastly, the most robust study to date has demonstrated that StAR is a bona fide target for miR-150 [187]. miR-150 was shown by both RT-qPCR and in situ hybridization to be highly expressed in mouse Leydig cells relative to germ cells or Sertoli cells. Transfecting either primary Leydig cells or MA-10 mouse Leydig tumor cells with miR-150 antagomir resulted in decreased miR-150 levels, increased StAR mRNA and protein expression, and increased steroidogenesis. These results were complemented by demonstrating that a miR-150 agomir reduced StAR levels and steroidogenesis. Furthermore, modulating miR-150 levels in vivo by intratesticular injection of the agomir or antagomir recapitulated the in vitro results. Full-length mouse StAR 3΄-UTR and a mutant that disrupted the miR-150 seed sequence were tested using Luciferase reporter assays in transfected MA-10 cells. The presence of miR-150 agomir decreased Luciferase activity 3-fold for the wild-type 3΄-UTR but had no effect on the mutant 3΄-UTR. Although the rationale for picking miR-150 to study in this context is not clear, this miRNA should get further attention as it may play a role in modulating steroidogenesis in Leydig cells via targeting StAR. miR-150 was not listed as a differentially expressed miRNA in LH-treated TM3 cells and it may reflect a mechanism to maintain low basal StAR mRNA levels.

Potential indirect mechanisms may also contribute to miRNA-mediated changes in StAR expression. One example is miR-133a targeting FOXL2, a member of the forkhead/hepatocyte nuclear factor 3 (FKH/HNF3) gene family, contributing to control of ovarian steroidogenesis. FOXL2 binds to the STAR proximal promoter (–42 bp) and represses StAR activity in the mouse ovary [188]. Foxl2 mRNA was shown to be a direct target of miRNA-133a and FSH stimulation of KGN human granulosa cells and isolated mouse granulosa cells resulted in increased miR-133a and decreased Foxl2 expression. These studies indicate the miRNA-dependent loss of the FOXL2 repressor contributes to the FSH-dependent increase in StAR expression [161, 189].

Important questions for the future

In the evolution of mitochondrial cholesterol import models over the years, the core element that has stood the test of time is StAR in that it is imperative for cholesterol import to the IMM in somatic steroidogenic cells (Figure 2). Nevertheless, precise knowledge on how this process occurs is something that remains a topic of ongoing investigations. When TSPO is removed from the steroidogenesis equation, the next model that surfaces for explaining StAR mechanism of action is the proposal of its existence as a “molten globule” in the OMM which transitions to take up cholesterol and deliver it to the IMM [103]. In support, recent research has postulated that StAR may interact with components of the mitochondria-associated endoplasmic reticulum membranes (MAMs) specific to the OMM proteins to regulate aspects of its structural folding [190] and subsequent translocation to the mitochondrial matrix [105]. However, considering that mitochondrial import of StAR may not be essential for cholesterol translocation [99, 100], this model fails to provide direct evidence for the mechanism of cholesterol import to the IMM. Moreover, the relationship between StAR and other proteins of the MAMs lacks clarity as disrupting MAMs and/or their regulators could interfere with other essential cellular functions that may indirectly influence steroidogenesis. One area that deserves more directed studies is whether StAR phosphorylation influences MAM interactions, if this is indeed the mechanism for cholesterol transfer. In summary, it would seem to follow that the most important future investigations need to focus on uncovering the exact mechanisms involved in transporting cholesterol to the IMM and the role of StAR in mediating this transport. Providing this information will undoubtedly represent the next biggest advance in this area.

Current iteration for the StAR-mediated mitochondrial cholesterol import model. In previous models, it was considered that at the outer mitochondrial membrane (OMM), StAR delivered cholesterol and TSPO functioned as a channel that mediates cholesterol transport to the inner mitochondrial membrane (IMM). Function of StAR was restricted to the mitochondrial outer membrane, as deletion of its mitochondrial targeting sequence (ΔN62-StAR) or anchoring it to the OMM by fusion to TOM20 (TOM20-StAR) did not affect its function. The transport mechanism was thought to involve TSPO polymerization and association with other proteins such as VDAC1 and ATAD3A. However, these association studies dealing with transmembrane proteins were highly prone to artifacts, and recent results have definitively demonstrated that TSPO is not involved in mitochondrial cholesterol import for steroidogenesis. This fundamental shift has resulted in the current model that mitochondrial cholesterol import is entirely mediated by StAR. It is known that the active form of StAR is a 37 kDa protein that enters the mitochondrial matrix to be processed to a 30 kDa inactive form. The basis for StAR function is partially explained by evidence that StAR exists as a molten globule at the OMM, with a folded N-terminal region and less folded C-terminal region. This state allows StAR to pause as it traverses the OMM and IMM, increasing its activity in importing cholesterol during this transition. At the OMM, VDAC2 has been shown to control StAR processing and activity. Cholesterol imported to the IMM is converted to pregnenolone by CYP11A1 that resides at the matrix side of the IMM, the first enzymatic step in steroid hormone biosynthesis.
Figure 2.

Current iteration for the StAR-mediated mitochondrial cholesterol import model. In previous models, it was considered that at the outer mitochondrial membrane (OMM), StAR delivered cholesterol and TSPO functioned as a channel that mediates cholesterol transport to the inner mitochondrial membrane (IMM). Function of StAR was restricted to the mitochondrial outer membrane, as deletion of its mitochondrial targeting sequence (ΔN62-StAR) or anchoring it to the OMM by fusion to TOM20 (TOM20-StAR) did not affect its function. The transport mechanism was thought to involve TSPO polymerization and association with other proteins such as VDAC1 and ATAD3A. However, these association studies dealing with transmembrane proteins were highly prone to artifacts, and recent results have definitively demonstrated that TSPO is not involved in mitochondrial cholesterol import for steroidogenesis. This fundamental shift has resulted in the current model that mitochondrial cholesterol import is entirely mediated by StAR. It is known that the active form of StAR is a 37 kDa protein that enters the mitochondrial matrix to be processed to a 30 kDa inactive form. The basis for StAR function is partially explained by evidence that StAR exists as a molten globule at the OMM, with a folded N-terminal region and less folded C-terminal region. This state allows StAR to pause as it traverses the OMM and IMM, increasing its activity in importing cholesterol during this transition. At the OMM, VDAC2 has been shown to control StAR processing and activity. Cholesterol imported to the IMM is converted to pregnenolone by CYP11A1 that resides at the matrix side of the IMM, the first enzymatic step in steroid hormone biosynthesis.

The question of what regulates the regulator of steroidogenesis is moving beyond transcriptional control to post-transcriptional mechanisms. The few studies to date in this area highlight that mechanisms that control steady-state StAR mRNA levels contribute to the acute tropic hormone response of steroidogenic cells to increase steroid output. The challenge is identifying the biologically relevant miRNAs that directly target StAR given that hypothesis-driven approach is limited by the lack of strong binding sites predicted for the 3΄-UTR. miRNA-150 provides a good example for how a case-by-case approach to this question might work.

In extraembryonic tissues like the human placenta that do not express StAR but are still capable of producing steroids, alternate mechanisms that facilitate mitochondrial cholesterol import appear to exist. STARD3 (also known as MLN64) was proposed to mediate mitochondrial cholesterol import in the placenta [76, 191]. However, MLN64 steroidogenic function has not been directly addressed in placental cells, and its precise mechanism responsible for the claim of mitochondrial cholesterol import remains far from clear. Recent studies on this protein seem to suggest that it plays a role in endoplasmic reticulum to endosome cholesterol transport [192]. Therefore, the basis of mitochondrial cholesterol import for steroidogenesis in the human placenta is also a topic that requires fundamental exploration.

Footnotes

Grant Support: Many of the studies referenced in this review were performed with the financial support of the National Institutes of Health (Grant Number HD17481) and the Robert A. Welch Foundation (Grant Number B1–0028) to DMS, Startup funds from Cornell University and National Institutes of Health (Grant Number DK110059) to VS, and National Institutes of Health (Grant Number DK51656) and American Heart Association (0555174B) to BJC.

References

1.

Miller
WL
.
Molecular biology of steroid hormone synthesis
.
Endocr Rev
1988
;
9
:
295
318
.

2.

Farkash
Y
,
Timberg
R
,
Orly
J
.
Preparation of antiserum to rat cytochrome P-450 cholesterol side chain cleavage, and its use for ultrastructural localization of the immunoreactive enzyme by protein A-gold technique
.
Endocrinology
1986
;
118
:
1353
1365
.

3.

Miller
WL
.
Early steps in androgen biosynthesis: from cholesterol to DHEA
.
Baillières Clin Endocrinol Metab
1998
;
12
:
67
81
.

4.

Waterman
MR
,
Simpson
ER
.
Regulation of the biosynthesis of cytochromes P-450 involved in steroid hormone synthesis
.
Mol Cell Endocrinol
1985
;
39
:
81
89
.

5.

Simpson
ER
,
Waterman
MR
.
Regulation of the synthesis of steroidogenic enzymes in adrenal cortical cells by ACTH
.
Annu Rev Physiol
1988
;
50
:
427
440
.

6.

Simpson
E
,
Lauber
M
,
Demeter
M
,
Means
G
,
Mahendroo
M
,
Kilgore
M
,
Mendelson
C
,
Waterman
M
.
Regulation of expression of the genes encoding steroidogenic enzymes in the ovary
.
J Steroid Biochem Mol Biol
1992
;
41
:
409
413
.

7.

Miller
WL
,
Auchus
RJ
.
The molecular biology, biochemistry, and physiology of human steroidogenesis and its disorders
.
Endocr Rev
2011
;
32
:
81
151
.

8.

Black
SM
,
Harikrishna
JA
,
Szklarz
GD
,
Miller
WL
.
The mitochondrial environment is required for activity of the cholesterol side-chain cleavage enzyme, cytochrome P450scc.
Proc Natl Acad Sci USA
1994
;
91
:
7247
7251
.

9.

Stone
D
,
Hechter
O
.
Studies on ACTH action in perfused bovine adrenals: the site of action of ACTH in corticosteroidogenesis
.
Arch Biochem Biophys
1954
;
51
:
457
469
.

10.

Ferguson
JJ Jr
.
Puromycin and adrenal responsiveness to adrenocorticotropic hormone
.
Biochim Biophys Acta
1962
;
57
:
616
617
.

11.

Ferguson
JJ Jr
.
Protein synthesis and adrenocorticotropin responsiveness
.
J Biol Chem
1963
;
238
:
2754
2759
.

12.

Garren
LD
,
Ney
RL
,
Davis
WW
.
Studies on the role of protein synthesis in the regulation of corticosterone production by adrenocorticotropic hormone in vivo.
Proc Natl Acad Sci USA
1965
;
53
:
1443
1450
.

13.

Garren
LD
,
Davis
WW
,
Crocco
RM
,
Ney
RL
.
Puromycin analogs: action of adrenocorticotropic hormone and the role of glycogen
.
Science
1966
;
152
:
1386
1388
.

14.

Garren
LD
.
The mechanism of action of adrenocorticotropic hormone
.
Vitam Horm
1968
;
26
:
119
145
.

15.

Davis
WW
,
Garren
LD
.
On the mechanism of action of adrenocorticotropic hormone. The inhibitory site of cycloheximide in the pathway of steroid biosynthesis
.
J Biol Chem
1968
;
243
:
5153
5157
.

16.

Simpson
ER
,
Boyd
GS
.
The cholesterol side-chain cleavage system of the adrenal cortex: a mixed-function oxidase
.
Biochem Biophys Res Commun
1966
;
24
:
10
17
.

17.

Arthur
JR
,
Boyd
GS
.
The effect of inhibitors of protein synthesis on cholesterol side-chain cleavage in the mitochondria of luteinized rat ovaries
.
Eur J Biochem
1974
;
49
:
117
127
.

18.

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 USA
1983
;
80
:
702
706
.

19.

Ohno
Y
,
Yanagibashi
K
,
Yonezawa
Y
,
Ishiwatari
S
,
Matsuba
M
.
A possible role of ``steroidogenic factor'' in the corticoidogenic response to ACTH; Effect of ACTH, cycloheximide and aminoglutethimide on the content of cholesterol in the outer and inner mitochondrial membrane of rat adrenal cortex.
Endocrinol Japon
1983
;
30
:
335
338
.

20.

Vahouny
GV
,
Chanderbhan
R
,
Kharroubi
A
,
Noland
BJ
,
Pastuszyn
A
,
Scallen
TJ
.
Sterol carrier and lipid transfer proteins
.
Adv Lipid Res
1987
;
22
:
83
113
.

21.

Pedersen
RC
,
Brownie
AC
.
Cholesterol side-chain cleavage in the rat adrenal cortex: isolation of a cycloheximide-sensitive activator peptide.
Proc Natl Acad Sci USA
1983
;
80
:
1882
1886
.

22.

Pedersen
RC
,
Brownie
AC
.
Steroidogenesis-activator polypeptide isolated from a rat Leydig cell tumor
.
Science
1987
;
236
:
188
190
.

23.

Seedorf
U
,
Raabe
M
,
Ellinghaus
P
,
Kannenberg
F
,
Fobker
M
,
Engel
T
,
Denis
S
,
Wouters
F
,
Wirtz
KW
,
Wanders
RJ
,
Maeda
N
,
Assmann
G
.
Defective peroxisomal catabolism of branched fatty acyl coenzyme A in mice lacking the sterol carrier protein-2/sterol carrier protein-x gene function
.
Genes Dev
1998
;
12
:
1189
1201
.

24.

Li
XA
,
Warren
DW
,
Gregoire
J
,
Pedersen
RC
,
Lee
AS
.
The rat 78,000 dalton glucose-regulated protein (GRP78) as a precursor for the rat steroidogenesis-activator polypeptide (SAP): the SAP coding sequence is homologous with the terminal end of GRP78
.
Mol Endocrinol
1989
;
3
:
1944
1952
.

25.

Wang
M
,
Wey
S
,
Zhang
Y
,
Ye
R
,
Lee
AS
.
Role of the unfolded protein response regulator GRP78/BiP in development, cancer, and neurological disorders
.
Antioxid Redox Signal
2009
;
11
:
2307
2316
.

26.

Luo
S
,
Mao
C
,
Lee
B
,
Lee
AS
.
GRP78/BiP is required for cell proliferation and protecting the inner cell mass from apoptosis during early mouse embryonic development
.
Mol Cell Biol
2006
;
26
:
5688
5697
.

27.

Wey
S
,
Luo
B
,
Tseng
CC
,
Ni
M
,
Zhou
H
,
Fu
Y
,
Bhojwani
D
,
Carroll
WL
,
Lee
AS
.
Inducible knockout of GRP78/BiP in the hematopoietic system suppresses Pten-null leukemogenesis and AKT oncogenic signaling
.
Blood
2012
;
119
:
817
825
.

28.

Gavish
M
,
Bachman
I
,
Shoukrun
R
,
Katz
Y
,
Veenman
L
,
Weisinger
G
,
Weizman
A
.
Enigma of the peripheral benzodiazepine receptor
.
Pharmacol Rev
1999
;
51
:
629
650
.

29.

Mukhin
AG
,
Papadopoulos
V
,
Costa
E
,
Krueger
KE
.
Mitochondrial benzodiazepine receptors regulate steroid biosynthesis.
Proc Natl Acad Sci USA
1989
;
86
:
9813
9816
.

30.

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
.

31.

Anholt
RR
,
De Souza
EB
,
Oster-Granite
,
ML
,
Snyder
SH
.
Peripheral-type benzodiazepine receptors: autoradiographic localization in whole-body sections of neonatal rats
.
J Pharmacol Exp Ther
1985
;
233
:
517
526
.

32.

Anholt
RR
,
Pedersen
PL
,
De Souza
EB
Snyder SH.
The peripheral-type benzodiazepine receptor. Localization to the mitochondrial outer membrane
.
J Biol Chem
1986
;
261
:
576
583
.

33.

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
.

34.

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
.

35.

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
.

36.

Papadopoulos
V
,
Berkovich
A
,
Krueger
KE
.
The role of diazepam binding inhibitor and its processing products at mitochondrial benzodiazepine receptors: regulation of steroid biosynthesis
.
Neuropharmacology
1991
;
30
:
1417
1423
.

37.

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
.

38.

Papadopoulos
V
,
Guarneri
P
,
Kreuger
KE
,
Guidotti
A
,
Costa
E
.
Pregnenolone biosynthesis in C6-2B glioma cell mitochondria: regulation by a mitochondrial diazepam binding inhibitor receptor.
Proc Natl Acad Sci USA
1992
;
89
:
5113
5117
.

39.

Garnier
M
,
Boujrad
N
,
Oke
BO
,
Brown
AS
,
Riond
J
,
Ferrara
P
,
Shoyab
M
,
Suarez-Quian
CA
,
Papadopoulos
V
.
Diazepam binding inhibitor is a paracrine/autocrine regulator of Leydig cell proliferation and steroidogenesis: action via peripheral-type benzodiazepine receptor and independent mechanisms.
Endocrinology
1993
;
132
:
444
458
.

40.

Boujrad
N
,
Hudson
JR
Jr
,
Papadopoulos
V
.
Inhibition of hormone-stimulated steroidogenesis in cultured Leydig tumor cells by a cholesterol-linked phosphorothioate oligodeoxynucleotide antisense to diazepam-binding inhibitor.
Proc Natl Acad Sci USA
1993
;
90
:
5728
5731
.

41.

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
.

42.

Papadopoulos
V
,
Miller
WL
.
Role of mitochondria in steroidogenesis. Best practice & research
.
Clin Endocrinol Metab
2012
;
26
:
771
790
.

43.

Papadopoulos
V
,
Amri
H
,
Boujrad
N
,
Cascio
C
,
Culty
M
,
Garnier
M
,
Hardwick
M
,
Li
H
,
Vidic
B
,
Brown
AS
,
Reversa
JL
,
Bernassau
JM
et al.
Peripheral benzodiazepine receptor in cholesterol transport and steroidogenesis
.
Steroids
1997
;
62
:
21
28
.

44.

Garnier
M
,
Boujrad
N
,
Ogwuegbu
SO
,
Hudson
JR
Jr
,
Papadopoulos
V
.
The polypeptide diazepam-binding inhibitor and a higher affinity mitochondrial peripheral-type benzodiazepine receptor sustain constitutive steroidogenesis in the R2C Leydig tumor cell line
.
J Biol Chem
1994
;
269
:
22105
22112
.

45.

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 USA
2001
;
98
:
1267
1272
.

46.

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
.

47.

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
.

48.

Midzak
A
,
Zirkin
B
,
Papadopoulos
V
.
Translocator protein: pharmacology and steroidogenesis
.
Biochem Soc Trans
2015
;
43
:
572
578
.

49.

Papadopoulos
V
,
Aghazadeh
Y
,
Fan
J
,
Campioli
E
,
Zirkin
B
,
Midzak
A
.
Translocator protein-mediated pharmacology of cholesterol transport and steroidogenesis
.
Mol Cell Endocrinol
2015
;
408
:
90
98
.

50.

Aghazadeh
Y
,
Zirkin
BR
,
Papadopoulos
V
.
Pharmacological regulation of the cholesterol transport machinery in steroidogenic cells of the testis
.
Vitam Horm
2015
;
98
:
189
227
.

51.

Krueger
RJ
,
Orme-Johnson
NR
.
Acute adrenocorticotropic hormone stimulation of adrenal corticosteroidogenesis. Discovery of a rapidly induced protein
.
J Biol Chem
1983
;
258
:
10159
10167
.

52.

Epstein
LF
,
Orme-Johnson
NR
.
Acute action of luteinizing hormone on mouse Leydig cells: accumulation of mitochondrial phosphoproteins and stimulation of testosterone synthesis
.
Mol Cell Endocrinol
1991
;
81
:
113
126
.

53.

Epstein
LF
,
Orme-Johnson
NR
.
Regulation of steroid hormone biosynthesis. Identification of precursors of a phosphoprotein targeted to the mitochondrion in stimulated rat adrenal cortex cells
.
J Biol Chem
1991
;
266
:
19739
19745
.

54.

Pon
LA
,
Orme-Johnson
NR
.
Acute stimulation of corpus luteum cells by gonadotrophin or adenosine 3΄,5΄-monophosphate causes accumulation of a phosphoprotein concurrent with acceleration of steroid synthesis
.
Endocrinology
1988
;
123
:
1942
1948
.

55.

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
.

56.

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
.

57.

Pon
LA
,
Epstein
LF
,
Orme-Johnson
NR
.
Acute cAMP stimulation in Leydig cells: rapid accumulation of a protein similar to that detected in adrenal cortex and corpus luteum
.
Endocr Res
1986
;
12
:
429
446
.

58.

Alberta
JA
,
Epstein
LF
,
Pon
LA
,
Orme-Johnson
NR
.
Mitochondrial localization of a phosphoprotein that rapidly accumulates in adrenal cortex cells exposed to adrenocorticotropic hormone or to cAMP
.
J Biol Chem
1989
;
264
:
2368
2372
.

59.

Stocco
DM
,
Kilgore
MW
.
Induction of mitochondrial proteins in MA-10 Leydig tumour cells with human choriogonadotropin
.
Biochem J
1988
;
249
:
95
103
.

60.

Stocco
DM
,
Chaudhary
LR
.
Evidence for the functional coupling of cyclic AMP in MA-10 mouse Leydig tumour cells
.
Cell Signal
1990
;
2
:
161
170
.

61.

Stocco
DM
,
Sodeman
TC
.
The 30-kDa mitochondrial proteins induced by hormone stimulation in MA-10 mouse Leydig tumor cells are processed from larger precursors
.
J Biol Chem
1991
;
266
:
19731
19738
.

62.

Stocco
DM
,
Chen
W
.
Presence of identical mitochondrial proteins in unstimulated constitutive steroid-producing R2C rat Leydig tumor and stimulated nonconstitutive steroid-producing MA-10 mouse Leydig tumor cells
.
Endocrinology
1991
;
128
:
1918
1926
.

63.

Stocco
DM
.
Further evidence that the mitochondrial proteins induced by hormone stimulation in MA-10 mouse Leydig tumor cells are involved in the acute regulation of steroidogenesis
.
J Steroid Biochem Mol Biol
1992
;
43
:
319
333
.

64.

Stocco
DM
,
Ascoli
M
.
The use of genetic manipulation of MA-10 Leydig tumor cells to demonstrate the role of mitochondrial proteins in the acute regulation of steroidogenesis.
Endocrinology
1993
;
132
:
959
967
.

65.

Stocco
DM
,
King
S
,
Clark
BJ
.
Differential effects of dimethylsulfoxide on steroidogenesis in mouse MA-10 and rat R2C Leydig tumor cells.
Endocrinology
1995
;
136
:
2993
2999
.

66.

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
.

67.

Stocco
DM
,
Clark
BJ
.
Regulation of the acute production of steroids in steroidogenic cells
.
Endocr Rev
1996
;
17
:
221
244
.

68.

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
.

69.

Sugawara
T
,
Lin
D
,
Holt
JA
,
Martin
KO
,
Javitt
NB
,
Miller
WL
,
Strauss
JF
3rd
.
Structure of the human steroidogenic acute regulatory (StAR) protein gene: StAR stimulates mitochondrial cholesterol 27-hydroxylase
.
Biochemistry
1995
;
34
:
12506
12512
.

70.

Sugawara
T
,
Holt
JA
,
Driscoll
D
,
Strauss
JF
3rd
,
Lin D
,
Miller
WL
,
Patterson
D
,
Clancy
KP
,
Hart
IM
,
Clark
BJ
et al.
Human steroidogenic acute regulatory protein: functional activity in COS-1 cells, tissue-specific expression, and mapping of the structural gene to 8p11.2 and a pseudogene to chromosome 13.
Proc Natl Acad Sci USA
1995
;
92
:
4778
4782
.

71.

Tomasetto
C
,
Regnier
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 to the q11-q21.3 region of chromosome 17
.
Genomics
1995
;
28
:
367
376
.

72.

Ponting
CP
,
Aravind
L
.
START: a lipid-binding domain in StAR, HD-ZIP and signalling proteins
.
Trends Biochem Sci
1999
;
24
:
130
132
.

73.

Moog-Lutz
C
,
Tomasetto
C
,
Regnier
CH
,
Wendling
C
,
Lutz
Y
,
Muller
D
,
Chenard
MP
,
Basset
P
,
Rio
MC
.
MLN64 exhibits homology with the steroidogenic acute regulatory protein (STAR) and is over-expressed in human breast carcinomas
.
Int J Cancer
1997
;
71
:
183
191
.

74.

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
.

75.

Borthwick
F
,
Allen
AM
,
Taylor
JM
,
Graham
A
.
Overexpression of STARD3 in human monocyte/macrophages induces an anti-atherogenic lipid phenotype
.
Clin Sci
2010
;
119
:
265
272
.

76.

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 USA
1997
;
94
:
8462
8467
.

77.

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
.

78.

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
.

79.

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
.

80.

Voutilainen
R
,
Miller
WL
.
Developmental expression of genes for the stereoidogenic enzymes P450scc (20,22-desmolase), P450c17 (17 alpha-hydroxylase/17,20-lyase), and P450c21 (21-hydroxylase) in the human fetus
.
J Clin Endocrinol Metab
1986
;
63
:
1145
1150
.

81.

Bose
HS
,
Sugawara
T
,
Strauss
JF 3rd
,
Miller
WL
,
International Congenital Lipoid Adrenal Hyperplasia C
.
The pathophysiology and genetics of congenital lipoid adrenal hyperplasia
.
N Engl J Med
1996
;
335
:
1870
1879
.

82.

Bose
HS
,
Pescovitz
OH
,
Miller
WL
.
Spontaneous feminization in a 46,XX female patient with congenital lipoid adrenal hyperplasia due to a homozygous frameshift mutation in the steroidogenic acute regulatory protein
.
J Clin Endocrinol Metab
1997
;
82
:
1511
1515
.

83.

Fujieda
K
,
Tajima
T
,
Nakae
J
,
Sageshima
S
,
Tachibana
K
,
Suwa
S
,
Sugawara
T
,
Strauss
JF
3rd
.
Spontaneous puberty in 46,XX subjects with congenital lipoid adrenal hyperplasia. Ovarian steroidogenesis is spared to some extent despite inactivating mutations in the steroidogenic acute regulatory protein (StAR) gene.
J Clin Invest
1997
;
99
:
1265
1271
.

84.

Miller
WL
.
Disorders in the initial steps of steroid hormone synthesis
.
J Steroid Biochem Mol Biol
2017
;
165
:
18
37
.

85.

Clark
BJ
,
Soo
SC
,
Caron
KM
,
Ikeda
Y
,
Parker
KL
,
Stocco
DM
.
Hormonal and developmental regulation of the steroidogenic acute regulatory protein
.
Mol Endocrinol
1995
;
9
:
1346
1355
.

86.

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 USA
1997
;
94
:
11540
11545
.

87.

Hasegawa
T
,
Zhao
L
,
Caron
KM
,
Majdic
G
,
Suzuki
T
,
Shizawa
S
,
Sasano
H
,
Parker
KL
.
Developmental roles of the steroidogenic acute regulatory protein (StAR) as revealed by StAR knockout mice
.
Mol Endocrinol
2000
;
14
:
1462
1471
.

88.

Kiriakidou
M
,
McAllister
JM
,
Sugawara
T
,
Strauss
JF
3rd
.
Expression of steroidogenic acute regulatory protein (StAR) in the human ovary
.
J Clin Endocrinol Metab
1996
;
81
:
4122
4128
.

89.

Liu
J
,
Heikkila
P
,
Kahri
AI
,
Voutilainen
R
.
Expression of the steroidogenic acute regulatory protein mRNA in adrenal tumors and cultured adrenal cells
.
J Endocrinol
1996
;
150
:
43
50
.

90.

Pollack
SE
,
Furth
EE
,
Kallen
CB
,
Arakane
F
,
Kiriakidou
M
,
Kozarsky
KF
,
Strauss
JF
3rd
.
Localization of the steroidogenic acute regulatory protein in human tissues
.
J Clin Endocrinol Metab
1997
;
82
:
4243
4251
.

91.

Pilon
N
,
Daneau
I
,
Brisson
C
,
Ethier
JF
,
Lussier
JG
,
Silversides
DW
.
Porcine and bovine steroidogenic acute regulatory protein (StAR) gene expression during gestation
.
Endocrinology
1997
;
138
:
1085
1091
.

92.

Lin
T
,
Wang
D
,
Hu
J
,
Stocco
DM
.
Upregulation of human chorionic gonadotrophin-induced steroidogenic acute regulatory protein by insulin-like growth factor-I in rat Leydig cells
.
Endocrine
1998
;
8
:
73
78
.

93.

Ronen-Fuhrmann
T
,
Timberg
R
,
King
SR
,
Hales
KH
,
Hales
DB
,
Stocco
DM
,
Orly
J
.
Spatio-temporal expression patterns of steroidogenic acute regulatory protein (StAR) during follicular development in the rat ovary
.
Endocrinology
1998
;
139
:
303
315
.

94.

Arensburg
J
,
Payne
AH
,
Orly
J
.
Expression of steroidogenic genes in maternal and extraembryonic cells during early pregnancy in mice
.
Endocrinology
1999
;
140
:
5220
5232
.

95.

Anuka
E
,
Yivgi-Ohana
N
,
Eimerl
S
,
Garfinkel
B
,
Melamed-Book
N
,
Chepurkol
E
,
Aravot
D
,
Zinman
T
,
Shainberg
A
,
Hochhauser
E
,
Orly
J
.
Infarct-induced steroidogenic acute regulatory protein: a survival role in cardiac fibroblasts
.
Mol Endocrinol
2013
;
27
:
1502
1517
.

96.

Anuka
E
,
Gal
M
,
Stocco
DM
,
Orly
J
.
Expression and roles of steroidogenic acute regulatory (StAR) protein in 'non-classical', extra-adrenal and extra-gonadal cells and tissues
.
Mol Cell Endocrinol
2013
;
371
:
47
61
.

97.

Tsujishita
Y
,
Hurley
JH
.
Structure and lipid transport mechanism of a StAR-related domain
.
Nat Struct Biol
2000
;
7
:
408
414
.

98.

Stocco
DM
.
StAR protein and the regulation of steroid hormone biosynthesis
.
Annu Rev Physiol
2001
;
63
:
193
213
.

99.

Arakane
F
,
Sugawara
T
,
Nishino
H
,
Liu
Z
,
Holt
JA
,
Pain
D
,
Stocco
DM
,
Miller
WL
,
Strauss
JF
3rd
.
Steroidogenic acute regulatory protein (StAR) retains activity in the absence of its mitochondrial import sequence: implications for the mechanism of StAR action
.
Proc Natl Acad Sci USA
1996
;
93
:
13731
13736
.

100.

Wang
X
,
Liu
Z
,
Eimerl
S
,
Timberg
R
,
Weiss
AM
,
Orly
J
,
Stocco
DM
.
Effect of truncated forms of the steroidogenic acute regulatory protein on intramitochondrial cholesterol transfer
.
Endocrinology
1998
;
139
:
3903
3912
.

101.

Bose
HS
,
Lingappa
VR
,
Miller
WL
.
Rapid regulation of steroidogenesis by mitochondrial protein import
.
Nature
2002
;
417
:
87
91
.

102.

Rajapaksha
M
,
Kaur
J
,
Bose
M
,
Whittal
RM
,
Bose
HS
.
Cholesterol-mediated conformational changes in the steroidogenic acute regulatory protein are essential for steroidogenesis
.
Biochemistry
2013
;
52
:
7242
7253
.

103.

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 USA
1999
;
96
:
7250
7255
.

104.

Christensen
K
,
Bose
HS
,
Harris
FM
,
Miller
WL
,
Bell
JD
.
Binding of steroidogenic acute regulatory protein to synthetic membranes suggests an active molten globule
.
J Biol Chem
2001
;
276
:
17044
17051
.

105.

Prasad
M
,
Kaur
J
,
Pawlak
KJ
,
Bose
M
,
Whittal
RM
,
Bose
HS
.
Mitochondria-associated endoplasmic reticulum membrane (MAM) regulates steroidogenic activity via steroidogenic acute regulatory protein (StAR)-Voltage-dependent anion channel 2 (VDAC2) interaction
.
J Biol Chem
2015
;
290
:
2604
2616
.

106.

Rajapaksha
M
,
Kaur
J
,
Prasad
M
,
Pawlak
KJ
,
Marshall
B
,
Perry
EW
,
Whittal
RM
,
Bose
HS
.
An outer mitochondrial translocase, Tom22, Is crucial for inner mitochondrial steroidogenic regulation in adrenal and gonadal tissues
.
Mol Cell Biol
2016
;
36
:
1032
1047
.

107.

Arakane
F
,
Sugawara
T
,
Nishino
H
,
Liu
Z
,
Holt
JA
,
Pain
D
,
Stocco
DM
,
Miller
WL
,
Strauss
JF
3rd
.
Steroidogenic acute regulatory protein (StAR) retains activity in the absence of its mitochondrial import sequence: implications for the mechanism of StAR action
.
Proc Natl Acad Sci USA
1996
;
93
:
13731
13736
.

108.

Jo
Y
,
King
SR
,
Khan
SA
,
Stocco
DM
.
Involvement of protein kinase C and cyclic adenosine 3΄,5΄-monophosphate-dependent kinase in steroidogenic acute regulatory protein expression and steroid biosynthesis in Leydig cells
.
Biol Reprod
2005
;
73
:
244
255
.

109.

Clark
BJ
.
ACTH action on StAR biology
.
Front Neurosci
2016
;
10
:
547
.

110.

Sasaki
G
,
Zubair
M
,
Ishii
T
,
Mitsui
T
,
Hasegawa
T
,
Auchus
RJ
.
The contribution of serine 194 phosphorylation to steroidogenic acute regulatory protein function
.
Mol Endocrinol
2014
;
28
:
1088
1096
.

111.

Dyson
MT
,
Jones
JK
,
Kowalewski
MP
,
Manna
PR
,
Alonso
M
,
Gottesman
ME
,
Stocco
DM
.
Mitochondrial A-kinase anchoring protein 121 binds type II protein kinase A and enhances steroidogenic acute regulatory protein-mediated steroidogenesis in MA-10 mouse leydig tumor cells
.
Biol Reprod
2008
;
78
:
267
277
.

112.

Dyson
MT
,
Kowalewski
MP
,
Manna
PR
,
Stocco
DM
.
The differential regulation of steroidogenic acute regulatory protein-mediated steroidogenesis by type I and type II PKA in MA-10 cells
.
Mol Cell Endocrinol
2009
;
300
:
94
103
.

113.

Paz
C
,
Cornejo Maciel
F
,
Gorostizaga
A
,
Castillo
AF
,
Mori Sequeiros Garcia
MM
,
Maloberti
PM
,
Orlando
UD
,
Mele
PG
,
Poderoso
C
,
Podesta
EJ
.
Role of protein phosphorylation and tyrosine phosphatases in the adrenal regulation of steroid synthesis and mitochondrial function
.
Front Endocrinol (Lausanne)
2016
;
7
:
60
.

114.

Grozdanov
PN
,
Stocco
DM
.
Short RNA molecules with high binding affinity to the KH motif of A-kinase anchoring protein 1 (AKAP1): implications for the regulation of steroidogenesis
.
Mol Endocrinol
2012
;
26
:
2104
2117
.

115.

Tu
LN
,
Zhao
AH
,
Stocco
DM
,
Selvaraj
V
.
PK11195 effect on steroidogenesis is not mediated through the translocator protein (TSPO)
.
Endocrinology
2015
;
156
:
1033
1039
.

116.

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
.

117.

Tu
LN
,
Morohaku
K
,
Manna
PR
,
Pelton
SH
,
Butler
WR
,
Stocco
DM
,
Selvaraj
V
.
Peripheral benzodiazepine receptor/translocator protein global Knock-out mice are viable with no effects on steroid hormone biosynthesis
.
J Biol Chem
2014
;
289
:
27444
27454
.

118.

Banati
RB
,
Middleton
RJ
,
Chan
R
,
Hatty
CR
,
Wai-Ying Kam
W
,
Quin
C
,
Graeber
MB
,
Parmar
A
,
Zahra
D
,
Callaghan
P
,
Fok
S
,
Howell
NR
et al.
Positron emission tomography and functional characterization of a complete PBR/TSPO knockout
.
Nat Commun
2014
;
5
:
5452
.

119.

Wang
H
,
Zhai
K
,
Xue
Y
,
Yang
J
,
Yang
Q
,
Fu
Y
,
Hu
Y
,
Liu
F
,
Wang
W
,
Cui
L
,
Chen
H
,
Zhang
J
et al.
Global deletion of TSPO does not affect the viability and gene expression profile
.
PLoS ONE
2016
;
11
:
e0167307
.

120.

Amri
H
,
Culty
M
,
Gaillard
JL
,
Teper
G
,
Papadopoulos
V
.
The peripheral-type benzodiazepine receptor and adrenal steroidogenesis
.
Curr Opin Endocrinol Diabetes
1999
;
6
:
179
184
.

121.

Veenman
L
,
Papadopoulos
V
,
Gavish
M
.
Channel-like functions of the 18-kDa translocator protein (TSPO): regulation of apoptosis and steroidogenesis as part of the host-defense response
.
CPD
2007
;
13
:
2385
2405
.

122.

Selvaraj
V
,
Stocco
DM
.
The changing landscape in translocator protein (TSPO) function
.
Trends Endocrinol Metab
2015
;
26
:
341
348
.

123.

Selvaraj
V
,
Stocco
DM
,
Tu
LN
.
Minireview: translocator protein (TSPO) and steroidogenesis: a reappraisal
.
Mol Endocrinol
2015
;
29
:
490
501
.

124.

Selvaraj
V
,
Tu
LN
.
Current status and future perspectives: TSPO in steroid neuroendocrinology
.
J Endocrinol
2016
;
231
:
R1
R30
.

125.

Fan
J
,
Campioli
E
,
Midzak
A
,
Culty
M
,
Papadopoulos
V
.
Conditional steroidogenic cell-targeted deletion of TSPO unveils a crucial role in viability and hormone-dependent steroid formation
.
Proc Natl Acad Sci USA
2015
;
112
:
7261
7266
.

126.

Selvaraj
V
,
Tu
LN
,
Stocco
DM
.
Crucial role reported for TSPO in viability and steroidogenesis is a misconception. Commentary: Conditional steroidogenic Cell-Targeted deletion of TSPO unveils a crucial role in viability and Hormone-Dependent steroid formation
.
Front Endocrinol (Lausanne)
2016
;
7
:
91
.

127.

Owen
DR
,
Fan
J
,
Campioli
E
,
Venugopal
S
,
Midzak
A
,
Daly
E
,
Harlay
A
,
Issop
L
,
Libri
V
,
Kalogiannopoulou
D
,
Oliver
E
,
Gallego-Colon
E
et al.
TSPO mutations in rats and a human polymorphism impair the rate of steroid synthesis
.
Biochem J
2017
;
474
:
3985
3999
.

128.

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
:
3985
3999
.

129.

Kelly-Hershkovitz
E
,
Weizman
R
,
Spanier
I
,
Leschiner
S
,
Lahav
M
,
Weisinger
G
,
Gavish
M
.
Effects of peripheral-type benzodiazepine receptor antisense knockout on MA-10 Leydig cell proliferation and steroidogenesis
.
J Biol Chem
1998
;
273
:
5478
5483
.

130.

Stocco
DM
,
Zhao
AH
,
Tu
LN
,
Morohaku
K
,
Selvaraj
V
.
A brief history of the search for the protein(s) involved in the acute regulation of steroidogenesis
.
Mol Cell Endocrinol
2017
;
441
:
7
16
.

131.

Batoko
H
,
Veljanovski
V
,
Jurkiewicz
P
.
Enigmatic Translocator protein (TSPO) and cellular stress regulation
.
Trends Biochem Sci
2015
;
40
:
497
503
.

132.

Guilarte
TR
,
Loth
MK
,
Guariglia
SR
.
TSPO finds NOX2 in microglia for redox homeostasis
.
Trends Pharmacol Sci
2016
;
37
:
334
343
.

133.

Gut
P
,
Zweckstetter
M
,
Banati
RB
.
Lost in translocation: the functions of the 18-kD translocator protein
.
Trends Endocrinol Metab
2015
;
26
:
349
356
.

134.

Tu
LN
,
Zhao
AH
,
Hussein
M
,
Stocco
DM
,
Selvaraj
V
.
Translocator protein (TSPO) affects mitochondrial fatty acid oxidation in steroidogenic cells
.
Endocrinology
2016
;
157
:
1110
1121
.

135.

Liu
GJ
,
Middleton
RJ
,
Kam
WW
,
Chin
DY
,
Hatty
CR
,
Chan
RH
,
Banati
RB
.
Functional gains in energy and cell metabolism after TSPO gene insertion
.
Cell Cycle
2017
;
16
:
436
447
.

136.

Barron
AM
,
Ji
B
,
Kito
S
,
Suhara
T
,
Higuchi
M
.
Steroidogenic abnormalities in translocator protein knockout mice and significance in the aging male
.
Biochem J
2018
;
475
:
75
85
.

137.

Manna
PR
,
Stocco
DM
.
Crosstalk of CREB and Fos/Jun on a single cis-element: transcriptional repression of the steroidogenic acute regulatory protein gene
.
J Mol Endocrinol
2007
;
39
:
261
277
.

138.

Manna
PR
,
Dyson
MT
,
Stocco
DM
.
Role of basic leucine zipper proteins in transcriptional regulation of the steroidogenic acute regulatory protein gene
.
Mol Cell Endocrinol
2009
;
302
:
1
11
.

139.

Clark
BJ
,
Stocco
DM
.
The Steroidogenic Acute Regulatory Protein (StAR)
. In:
Clark
JB
,
Stocco
MD
(eds.),
Cholesterol Transporters of the START Domain Protein Family in Health and Disease: START Proteins - Structure and Function
.
New York, NY
:
Springer
;
2014
:
15
47
.

140.

Bahrami
S
,
Drablos
F
.
Gene regulation in the immediate-early response process
.
Adv Biol Regul
2016
;
62
:
37
49
.

141.

Clem
BF
,
Hudson
EA
,
Clark
BJ
.
Cyclic adenosine 3΄,5΄-monophosphate (cAMP) enhances cAMP-responsive element binding (CREB) protein phosphorylation and phospho-CREB interaction with the mouse steroidogenic acute regulatory protein gene promoter
.
Endocrinology
2005
;
146
:
1348
1356
.

142.

Manna
PR
,
Dyson
MT
,
Eubank
DW
,
Clark
BJ
,
Lalli
E
,
Sassone-Corsi
P
,
Zeleznik
AJ
,
Stocco
DM
.
Regulation of steroidogenesis and the steroidogenic acute regulatory protein by a member of the cAMP response-element binding protein family
.
Mol Endocrinol
2002
;
16
:
184
199
.

143.

Manna
PR
,
Wang
XJ
,
Stocco
DM
.
Involvement of multiple transcription factors in the regulation of steroidogenic acute regulatory protein gene expression
.
Steroids
2003
;
68
:
1125
1134
.

144.

Wooton-Kee
CR
,
Clark
BJ
.
Steroidogenic factor-1 influences protein-deoxyribonucleic acid interactions within the cyclic adenosine 3΄,5΄-monophosphate-responsive regions of the murine steroidogenic acute regulatory protein gene
.
Endocrinology
2000
;
141
:
1345
1355
.

145.

Karin
M
,
Liu
Z
,
Zandi
E
.
AP-1 function and regulation
.
Curr Opin Cell Biol
1997
;
9
:
240
246
.

146.

Meier
RK
,
Clark
BJ
.
Angiotensin II-dependent transcriptional activation of human steroidogenic acute regulatory protein gene by a 25-kDa cAMP-responsive element modulator protein isoform and Yin Yang 1
.
Endocrinology
2012
;
153
:
1256
1268
.

147.

Nogueira
EF
,
Rainey
WE
.
Regulation of aldosterone synthase by activator transcription factor/cAMP response element-binding protein family members
.
Endocrinology
2010
;
151
:
1060
1070
.

148.

Sugawara
T
,
Sakuragi
N
,
Minakami
H
.
CREM confers cAMP responsiveness in human steroidogenic acute regulatory protein expression in NCI-H295R cells rather than SF-1/Ad4BP
.
J Endocrinol
2006
;
191
:
327
337
.

149.

Nogueira
EF
,
Bollag
WB
,
Rainey
WE
.
Angiotensin II regulation of adrenocortical gene transcription
.
Mol Cell Endocrinol
2009
;
302
:
230
236
.

150.

Hiroi
H
,
Christenson
LK
,
Chang
L
,
Sammel
MD
,
Berger
SL
,
Strauss
JF 3rd
.
Temporal and spatial changes in transcription factor binding and histone modifications at the steroidogenic acute regulatory protein (stAR) locus associated with stAR transcription
.
Mol Endocrinol
2004
;
18
:
791
806
.

151.

Clark
JB
,
Stocco
MD
.
The Steroidogenic Acute Regulatory Protein (StAR)
. In:
Clark
JB
,
Stocco
MD
(eds.),
Cholesterol Transporters of the START Domain Protein Family in Health and Disease: START Proteins - Structure and Function
.
New York, NY
:
Springer
;
2014
:
15
47
.

152.

Abdou
HS
,
Villeneuve
G
,
Tremblay
JJ
.
The calcium signaling pathway regulates leydig cell steroidogenesis through a transcriptional cascade involving the nuclear receptor NR4A1 and the steroidogenic acute regulatory protein
.
Endocrinology
2013
;
154
:
511
520
.

153.

Martin
LJ
,
Boucher
N
,
Brousseau
C
,
Tremblay
JJ
.
The orphan nuclear receptor NUR77 regulates hormone-induced StAR transcription in Leydig cells through cooperation with Ca2+/calmodulin-dependent protein kinase I
.
Mol Endocrinol
2008
;
22
:
2021
2037
.

154.

Martin
LJ
,
Tremblay
JJ
.
The nuclear receptors NUR77 and SF1 play additive roles with c-JUN through distinct elements on the mouse Star promoter
.
J Mol Endocrinol
2009
;
42
:
119
129
.

155.

Clark
BJ
,
Combs
R
.
Angiotensin II and cyclic adenosine 3΄,5΄-monophosphate induce human steroidogenic acute regulatory protein transcription through a common steroidogenic factor-1 element
.
Endocrinology
1999
;
140
:
4390
4398
.

156.

Sugawara
T
,
Kiriakidou
M
,
McAllister
JM
,
Kallen
CB
,
Strauss
JF 3rd
.
Multiple steroidogenic factor 1 binding elements in the human steroidogenic acute regulatory protein gene 5΄-flanking region are required for maximal promoter activity and cyclic AMP responsiveness
.
Biochemistry
1997
;
36
:
7249
7255
.

157.

Reinhart
AJ
,
Williams
SC
,
Clark
BJ
,
Stocco
DM
.
SF-1 (steroidogenic factor-1) and C/EBP beta (CCAAT/enhancer binding protein-beta) cooperate to regulate the murine StAR (steroidogenic acute regulatory) promoter
.
Mol Endocrinol
1999
;
13
:
729
741
.

158.

Tremblay
JJ
,
Hamel
F
,
Viger
RS
.
Protein kinase A-dependent cooperation between GATA and CCAAT/enhancer-binding protein transcription factors regulates steroidogenic acute regulatory protein promoter activity
.
Endocrinology
2002
;
143
:
3935
3945
.

159.

Nalbant
D
,
Williams
SC
,
Stocco
DM
,
Khan
SA
.
Luteinizing Hormone-Dependent gene regulation in leydig cells may be mediated by CCAAT/Enhancer-Binding Protein-β
.
Endocrinology
1998
;
139
:
272
279
.

160.

Silverman
E
,
Yivgi-Ohana
N
,
Sher
N
,
Bell
M
,
Eimerl
S
,
Orly
J
.
Transcriptional activation of the steroidogenic acute regulatory protein (StAR) gene: GATA-4 and CCAAT/enhancer-binding protein beta confer synergistic responsiveness in hormone-treated rat granulosa and HEK293 cell models
.
Mol Cell Endocrinol
2006
;
252
:
92
101
.

161.

Yivgi-Ohana
N
,
Sher
N
,
Melamed-Book
N
,
Eimerl
S
,
Koler
M
,
Manna
PR
,
Stocco
DM
,
Orly
J
.
Transcription of steroidogenic acute regulatory protein in the rodent ovary and placenta: alternative modes of cyclic adenosine 3΄, 5΄-monophosphate dependent and independent regulation
.
Endocrinology
2009
;
150
:
977
989
.

162.

Lefrancois-Martinez
AM
,
Blondet-Trichard
A
,
Binart
N
,
Val
P
,
Chambon
C
,
Sahut-Barnola
I
,
Pointud
JC
,
Martinez
A
.
Transcriptional control of adrenal steroidogenesis
.
J Biol Chem
2011
;
286
:
32976
32985
.

163.

Li
J
,
Feltzer
RE
,
Dawson
KL
,
Hudson
EA
,
Clark
BJ
.
Janus kinase 2 and calcium are required for angiotensin II-dependent activation of steroidogenic acute regulatory protein transcription in H295R human adrenocortical cells
.
J Biol Chem
2003
;
278
:
52355
52362
.

164.

Ozbay
T
,
Merrill
AH Jr
,
Sewer
MB
.
ACTH regulates steroidogenic gene expression and cortisol biosynthesis in the human adrenal cortex via sphingolipid metabolism
.
Endocr Res
2004
;
30
:
787
794
.

165.

Takemori
H
,
Okamoto
M
.
Regulation of CREB-mediated gene expression by salt inducible kinase
.
J Steroid Biochem Mol Biol
2008
;
108
:
287
291
.

166.

Ito
M
,
Yu
R
,
Jameson
JL
.
DAX-1 inhibits SF-1-mediated transactivation via a carboxy-terminal domain that is deleted in adrenal hypoplasia congenita.
Mol Cell Biol
1997
;
17
:
1476
1483
.

167.

Manna
PR
,
Dyson
MT
,
Jo
Y
,
Stocco
DM
.
Role of dosage-sensitive sex reversal, adrenal hypoplasia congenita, critical region on the X chromosome, gene 1 in protein kinase A- and protein kinase C-mediated regulation of the steroidogenic acute regulatory protein expression in mouse Leydig tumor cells: mechanism of action
.
Endocrinology
2009
;
150
:
187
199
.

168.

Zazopoulos
E
,
Lalli
E
,
Stocco
DM
,
Sassone-Corsi
P
.
DNA binding and transcriptional repression by DAX-1 blocks steroidogenesis
.
Nature
1997
;
390
:
311
315
.

169.

Ehrlund
A
,
Anthonisen
EH
,
Gustafsson
N
,
Venteclef
N
,
Robertson Remen
K
,
Damdimopoulos
AE
,
Galeeva
A
,
Pelto-Huikko
M
,
Lalli
E
,
Steffensen
KR
,
Gustafsson
J-Å
,
Treuter
E
.
E3 Ubiquitin ligase RNF31 cooperates with DAX-1 in transcriptional repression of steroidogenesis
.
Mol Cell Biol
2009
;
29
:
2230
2242
.

170.

Tajima
K
,
Dantes
A
,
Yao
Z
,
Sorokina
K
,
Kotsuji
F
,
Seger
R
,
Amsterdam
A
.
Down-regulation of steroidogenic response to gonadotropins in human and rat preovulatory granulosa cells involves mitogen-activated protein kinase activation and modulation of DAX-1 and steroidogenic factor-1
.
J Clin Endocrinol Metab
2003
;
88
:
2288
2299
.

171.

Buholzer
CF
,
Arrighi
JF
,
Abraham
S
,
Piguet
V
,
Capponi
AM
,
Casal
AJ
.
Chicken ovalbumin upstream promoter-transcription factor is a negative regulator of steroidogenesis in bovine adrenal glomerulosa cells
.
Mol Endocrinol
2005
;
19
:
65
75
.

172.

Shibata
H
,
Ikeda
Y
,
Mukai
T
,
Morohashi
K
,
Kurihara
I
,
Ando
T
,
Suzuki
T
,
Kobayashi
S
,
Murai
M
,
Saito
I
,
Saruta
T
.
Expression profiles of COUP-TF, DAX-1, and SF-1 in the human adrenal gland and adrenocortical tumors: possible implications in steroidogenesis
.
Mol Genet Metab
2001
;
74
:
206
216
.

173.

Shibata
H
,
Kurihara
I
,
Kobayashi
S
,
Yokota
K
,
Suda
N
,
Saito
I
,
Saruta
T
.
Regulation of differential COUP-TF-coregulator interactions in adrenal cortical steroidogenesis
.
J Steroid Biochem Mol Biol
2003
;
85
:
449
456
.

174.

Bakke
M
,
Lund
J
.
Mutually exclusive interactions of two nuclear orphan receptors determine activity of a cyclic adenosine 3΄,5΄-monophosphate-responsive sequence in the bovine CYP17 gene
.
Mol Endocrinol
1995
;
9
:
327
339
.

175.

Jo
Y
,
Stocco
DM
.
Regulation of steroidogenesis and steroidogenic acute regulatory protein in R2C cells by DAX-1 (Dosage-Sensitive Sex Reversal, Adrenal Hypoplasia Congenita, Critical Region on the X Chromosome, Gene-1)
.
Endocrinology
2004
;
145
:
5629
5637
.

176.

Osman
H
,
Murigande
C
,
Nadakal
A
,
Capponi
AM
.
Repression of DAX-1 and induction of SF-1 expression. Two mechanisms contributing to the activation of aldosterone biosynthesis in adrenal glomerulosa cells
.
J Biol Chem
2002
;
277
:
41259
41267
.

177.

Sandhoff
TW
,
McLean
MP
.
Repression of the rat steroidogenic acute regulatory (StAR) protein gene by PGF2alpha is modulated by the negative transcription factor DAX-1
.
Endocrine
1999
;
10
:
83
92
.

178.

Z-n
Wang
,
H
Takemori
,
Halder
SK
,
Nonaka
Y
,
Okamoto
M
.
Cloning of a novel kinase (SIK) of the SNF1/AMPK family from high salt diet-treated rat adrenal1
.
FEBS Lett
1999
;
453
:
135
139
.

179.

Katoh
Y
,
Takemori
H
,
Min
L
,
Muraoka
M
,
Doi
J
,
Horike
N
,
Okamoto
M
.
Salt-inducible kinase-1 represses cAMP response element-binding protein activity both in the nucleus and in the cytoplasm
.
Eur J Biochem
2004
;
271
:
4307
4319
.

180.

Duan
H
,
Cherradi
N
,
Feige
JJ
,
Jefcoate
C
.
cAMP-dependent posttranscriptional regulation of steroidogenic acute regulatory (STAR) protein by the zinc finger protein ZFP36L1/TIS11b
.
Mol Endocrinol
2009
;
23
:
497
509
.

181.

Lee
J
,
Yamazaki
T
,
Dong
H
,
Jefcoate
C
.
A single cell level measurement of StAR expression and activity in adrenal cells
.
Mol Cell Endocrinol
2017
;
441
:
22
30
.

182.

Lee
J
,
Tong
T
,
Duan
H
,
Foong
YH
,
Musaitif
I
,
Yamazaki
T
,
Jefcoate
C
.
Regulation of StAR by the N-terminal domain and coinduction of SIK1 and TIS11b/Znf36l1 in single cells
.
Front Endocrinol (Lausanne)
2016
;
7
:
107
.

183.

Li
C
,
Gao
S
,
Chen
S
,
Chen
L
,
Zhao
Y
,
Jiang
Y
,
Zheng
X
,
Zhou
X
.
Differential expression of microRNAs in luteinising hormone-treated mice TM3 Leydig cells
.
Andrologia
2018
;
50
:
e12824
.

184.

Hu
Z
,
Shen
W-J
,
Cortez
Y
,
Tang
X
,
Liu
L-F
,
Kraemer
FB
,
Azhar
S
.
Hormonal regulation of microRNA expression in steroid producing cells of the ovary, testis and adrenal gland
.
PLoS ONE
2013
;
8
:
e78040
.

185.

Men
Y
,
Fan
Y
,
Shen
Y
,
Lu
L
,
Kallen
AN
.
The steroidogenic acute regulatory protein (StAR) is regulated by the H19/let-7 Axis
.
Endocrinology
2017
;
158
:
402
409
.

186.

Kallen
AN
,
Zhou
XB
,
Xu
J
,
Qiao
C
,
Ma
J
,
Yan
L
,
Lu
L
,
Liu
C
,
Yi
JS
,
Zhang
H
,
Min
W
,
Bennett
AM
et al.
The imprinted H19 lncRNA antagonizes let-7 microRNAs
.
Mol Cell
2013
;
52
:
101
112
.

187.

Geng
XJ
,
Zhao
DM
,
Mao
GH
,
Tan
L
.
MicroRNA-150 regulates steroidogenesis of mouse testicular Leydig cells by targeting STAR
.
Reproduction
2017
;
154
:
129
136
.

188.

Pisarska
MD
,
Bae
J
,
Klein
C
,
Hsueh
AJ
.
Forkhead L2 is expressed in the ovary and represses the promoter activity of the steroidogenic acute regulatory gene
.
Endocrinology
2004
;
145
:
3424
3433
.

189.

Dai
A
,
Sun
H
,
Fang
T
,
Zhang
Q
,
Wu
S
,
Jiang
Y
,
Ding
L
,
Yan
G
,
Hu
Y
.
MicroRNA-133b stimulates ovarian estradiol synthesis by targeting Foxl2
.
FEBS Lett
2013
;
587
:
2474
2482
.

190.

Prasad
M
,
Pawlak
KJ
,
Burak
WE
,
Perry
EE
,
Marshall
B
,
Whittal
RM
,
Bose
HS
.
Mitochondrial metabolic regulation by GRP78
.
Sci Adv
2017
;
3
:
e1602038
.

191.

Bose
HS
,
Baldwin
MA
,
Miller
WL
.
Evidence that StAR and MLN64 act on the outer mitochondrial membrane as molten globules
.
Endocr Res
2000
;
26
:
629
637
.

192.

Wilhelm
LP
,
Wendling
C
,
Vedie
B
,
Kobayashi
T
,
Chenard
MP
,
Tomasetto
C
,
Drin
G
,
Alpy
F
.
STARD3 mediates endoplasmic reticulum-to-endosome cholesterol transport at membrane contact sites
.
EMBO J
2017
;
36
:
1412
1433
.

This article is published and distributed under the terms of the Oxford University Press, Standard Journals Publication Model (https://academic.oup.com/journals/pages/about_us/legal/notices)