Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2013 Jan 2.
Published in final edited form as: Mol Cell Endocrinol. 2011 Aug 16;348(1):165–175. doi: 10.1016/j.mce.2011.08.003

Sphingosine-1-Phosphate Rapidly Increases Cortisol Biosynthesis and the Expression of Genes Involved in Cholesterol Uptake and Transport in H295R Adrenocortical Cells

Natasha C Lucki a, Donghui Li b, Marion B Sewer b
PMCID: PMC3508734  NIHMSID: NIHMS318955  PMID: 21864647

Abstract

In the acute phase of adrenocortical steroidogenesis, adrenocorticotrophin (ACTH) activates a cAMP/PKA-signaling pathway that promotes the transport of free cholesterol to the inner mitochondrial membrane. We have previously shown that ACTH rapidly stimulates the metabolism of sphingolipids and the secretion of sphingosine-1-phosphate (S1P) in H295R cells. In this study, we examined the effect of S1P on genes involved in the acute phase of steroidogenesis. We show that S1P increases the expression of steroidogenic acute regulatory protein (StAR), 18-kDa translocator protein (TSPO), low-density lipoprotein receptor (LDLR), and scavenger receptor class B type I (SR-BI). S1P-induced StAR mRNA expression requires Gαi signaling, phospholipase C (PLC), Ca2+/calmodulin-dependent kinase II (CamKII), and ERK1/2 activation. S1P also increases intracellular Ca2+, the phosphorylation of hormone sensitive lipase (HSL) at Ser563, and cortisol secretion. Collectively, these findings identify multiple roles for S1P in the regulation of glucocorticoid biosynthesis.

Keywords: sphingosine-1-phosphate, StAR, H295R, cortisol, acute steroidogenesis

1. Introduction

In the zona fasciculata of the human adrenal cortex, cortisol is synthesized from cholesterol via the concerted action of P450 heme-containing monooxygenases (CYPs) and 3β-hydroxysteroid dehydrogenase (3β-HSD) enzymes (Miller WL, 2008, Payne AH and Hales DB, 2004, Sewer MB et al., 2007). Additionally, StAR, 18-kDa translocator protein (TSPO), cholesterol hydrolase HSL, low-density lipoprotein receptor (LDLR), and scavenger receptor class B type I (SR-BI) are equally important in assuring adequate amounts of substrate for steroid hormone production. Cortisol biosynthesis is primarily regulated by the peptide hormone ACTH, which upon binding to the melanocortin 2 receptor, activates a cAMP/PKA-dependent pathway. Acutely, ACTH promotes cholesterol uptake and mobilization from intracellular stores to the inner mitochondrial membrane, where the first enzymatic step in steroid hormone production occurs. In the chronic phase of steroidogenesis, ACTH coordinately activates the transcription of all genes in the steroid hormone biosynthetic pathway, thus maintaining optimal steroidogenic capacity (Arlt W and Stewart PM, 2005, Dammer EB et al., 2007, Sewer MB and Waterman MR, 2003).

Import of free cholesterol into the inner mitochondrial membrane is facilitated by the formation of a macromolecular protein complex at the outer mitochondrial membrane containing TSPO, StAR, the TSPO-associated protein PAP7, and the regulatory subunit RIαof PKA (Hauet T et al., 2002, Liu J et al., 2006). StAR is essential for cholesterol trafficking (Caron KM et al., 1997b) and is rapidly synthesized in response to activation of the ACTH signaling pathway (Stocco DM and Clark BJ, 1996). The regulation of StAR transcription has been extensively studied [reviewed in (Manna PR et al., 2009a)], and the nuclear receptor steroidogenic factor 1 (SF-1) is required for StAR gene expression (Caron KM et al., 1997a) via its binding to a cAMP-responsive region located within the first 350 base pairs upstream of the transcription initiation site (Clark BJ and Combs R, 1999, Sandhoff TW et al., 1998, Sugawara T et al., 1996). Alternatively, Ca2+ influx through L-type Ca2+ channels abrogates StAR transcription by inducing the expression of the transcription factor DAX1 (dosage-sensitive sex reversal, adrenal hypoplasia critical region, on chromosome X, gene 1), a repressor of the StAR gene (Pandey AK et al., 2010). ACTH/cAMP rapidly increases StAR mRNA and protein levels (Lin D et al., 1995, Sewer MB and Waterman MR, 2001, Stocco D et al., 2005) as well as its enzymatic activity through PKA-mediated phosphorylation events at conserved serine residues (Arakane F et al., 1997). Further, the protein kinase C (PKC) signaling pathway also plays an integral role in StAR regulation by potentiating cAMP-stimulated StAR expression and phosphorylation (Manna PR et al., 2009c). Of note, Manna et al. have recently shown that different PKC isoforms, including PKCα, PKCδ, PKCεand PKD, exhibit diverse effects on PMA-mediated StAR transcription and steroidogenesis (Manna PR et al., 2011). Finally, A-kinase anchoring protein 121 (AKAP121) enhances the post-transcriptional regulation of StAR by recruiting type II PKA regulatory subunit α (PKAR2α) and StAR mRNA transcripts to the outer mitochondria membrane, thus promoting spatially localized protein synthesis and increasing steroidogenic efficiency (Dyson MT et al., 2008). Similarly, HSL, which is responsible for the neutral cholesteryl ester hydrolase activity in steroidogenic tissues (Kraemer FB and Shen WJ, 2002), plays a vital role in steroid hormone production by increasing the availability of free cholesterol. HSL interacts with StAR and this partnering between the two proteins facilitates the transport of cholesterol from lipid droplets to mitochondria (Luo Y et al., 2011).

Sphingolipids are structural components of cell membranes and key signaling mediators of many cellular processes (Cuvillier O, 2002, Goni FM and Alonso A, 2006, Hannun YA, 1996, Huwiler A et al., 2000, Kihara A et al., 2007, Lucki N and Sewer MB, 2008, Merrill AH Jr et al., 1999, Merrill Jr. AH, 2002, Ogretmen B, 2006, Zeidan YH and Hannun YA, 2007). Several studies have reported various roles for sphingolipids in adrenal and gonadal steroidogenesis (Brizuela L et al., 2006, Budnik LT et al., 1999, Degnan BM et al., 1996, Kwun C et al., 1999, Lucki N and Sewer MB, 2008, Lucki N and Sewer MB, 2009, McClellan DR et al., 1997, Meroni SB et al., 2000, Ozbay T et al., 2004, Porn MI et al., 1991, Santana P et al., 1996, Urs AN et al., 2006). As a bioactive lipid mediator, S1P regulates a broad array of physiological functions, including cell proliferation and survival (Olivera A et al., 1999, Olivera A and Spiegel S, 1993, Spiegel S and Milstien S, 2002), chemotaxis (Hla T et al., 1999, Wang F et al., 1999, Yamamura S et al., 1996), and protection against ceramide-mediated apoptosis (Cuvillier O et al., 1996). S1P has also been shown to stimulate cortisol production in zona fasciculata bovine adrenal cells in a PKC- and Ca2+-dependent manner (Rabano M et al., 2003), and promote aldosterone secretion in bovine glomerulosa cells via the phosphoinositide 3-kinase (PI3K)/Akt and mitogen-activated protein kinase (MAPK/ERK) pathways (Brizuela L et al., 2007, Brizuela L et al., 2006). S1P also mediates cAMP-dependent cortisol secretion in H295R human adrenocortical cells by promoting the SREBP-1-dependent transcription of CYP17 (Ozbay T et al., 2006).

S1P not only functions as an intracellular messenger (Hla T et al., 2001, Olivera A and Spiegel S, 2001) but also exerts many of its effects through cell surface G-protein coupled receptors (Alvarez SE et al., 2007). The mechanism of S1P export from cells is not completely understood, however, studies have provided evidence for the involvement of the ATP-binding cassette (ABC) family of transporters in this process (Anada Y et al., 2007, Kobayashi N et al., 2006, Mitra P et al., 2006, Takabe K et al., 2010). Five S1P receptors (S1PR1–5) have been identified (Sanchez T and Hla T, 2004, Spiegel S and Milstien S, 2000), four of which (S1PR1, S1PR2, S1PR3, S1PR5) are expressed in H295R steroidogenic cells (Ozbay T et al., 2006). Differences in signaling through these receptors are primarily due to differential coupling to G-proteins. S1PR1 signals through Gαi (Ancellin N and Hla T, 1999, Windh RT et al., 1999) whereas S1PR2 and S1PR3 couple to Gαi, Gαq, and Gα13 (Ancellin N and Hla T, 1999). S1PR4 associates with Gαi and Gα12/13 (Graler MH et al., 2003, Van Brocklyn JR et al., 2000, Yamazaki Y et al., 2000) and S1PR5 couples to Gαi/o and Gα12 (Malek RL et al., 2001).

We have shown that ACTH rapidly stimulates sphingolipid metabolism in H295R cells (Ozbay T et al., 2004). ACTH and the cAMP analog Bt2cAMP decrease cellular amounts of sphingomyelin, ceramide, and sphingosine, while simultaneously increasing the secretion of S1P (Ozbay T et al., 2004). The S1P produced stimulates cortisol secretion from H295R cells by promoting the maturation and binding of sterol regulatory element binding protein 1 (SREBP1) to the CYP17 promoter, thereby inducing gene transcription (Ozbay T et al., 2006). These findings implicate S1P as a paracrine mediator of ACTH-dependent CYP17 transcription. Therefore, the aim of the present study was to characterize the role of S1P in mediating the acute phase of steroidogenesis in H295R cells. We show that S1P rapidly increases cortisol biosynthesis and the mRNA expression of multiple genes involved in the acute phase of steroid hormone production including StAR, TSPO, SR-BI, and LDLR. In addition, we demonstrate that S1P acutely increases the phosphorylation of HSL at Ser563 and show that S1P-stimulated StAR gene expression is pertussis toxin sensitive and dependent on the activation of PLC, CamKII, and ERK1/2.

2. Materials and Methods

2.1. Reagents

Bt2cAMP was obtained from Sigma (St. Louis, MO). Adrenocorticotropin (ACTH 1–39) was purchased from American Peptide Co. (Sunnyvale, CA). S1P and VPC23019 were obtained from Avanti Polar Lipids Inc (Alabaster, AL) and prepared by dissolving in ethanol and dimethylamine, followed by evaporation and solubilization in 4 mg/ml fatty-acid-free BSA. Pertussis toxin (PTX), U0126, KN-93, U73122, U73343, and Fluo 3/AM were obtained from EMD Biosciences (La Jolla, CA).

2.2. Cell culture and Treatment

H295R adrenocortical cells (Rainey WE et al., 1994, Rainey WE et al., 2004, Staels B et al., 1993) were generously donated by Dr. William E. Rainey (Medical College of Georgia, Augusta, GA) and cultured in Dulbecco's modified Eagle's/F12 medium (DMEM/F12) (Mediatech, Manassas, VA) supplemented with 10% Nu-Serum I (BD Biosciences, Palo Alto, CA), 1% ITS Plus (BD Biosciences, Palo Alto, CA), antibiotics, and antimycotics. H295R cells were sub-cultured in 12-well plates and pretreated for 18 h with 10 µM VPC23019 (S1PR1 and S1PR3 inhibitor), 5 pg/ml PTX, 10 µM U73122 (PLC inhibitor), 10 µM U73343 (inactive PLC analog), 10 µM KN-93 (CamKII inhibitor), or 10 µM U0126 (MEK1 inhibitor) followed by treatment with 1 µM S1P, 50 nM ACTH, or 0.4 mM Bt2cAMP for 2 h.

2.3. Cortisol and DHEA Assay

Cells were cultured in 12-well plates and treated with 0.4 mM Bt2cAMP or 1 µM S1P for 3- to 48 h. Cortisol and DHEA released into the media were determined in triplicate against standards made up in DMEM/F12 medium using a 96-well plate enzyme-linked immunosorbent assays (ELISA; Enzo Life Sciences Inc., Plymouth Meeting, PA). Results are expressed as nanomoles per milligram of cellular protein in each sample.

2.4. RNA isolation and real time RT-PCR

Total RNA was extracted using Isol-RNA Lysis Reagent (5 Prime, Inc., Gaithersburg, MD) and amplified using a One-Step SYBR Green RT-PCR Kit (Thermo Scientific Inc., Waltham, MA) and the following primer sets: β-actin (forward 5’-ACG GCT CCG GCA TGT GCA AG-3’ and reverse 5’-TGA CGA TGC CGT GCT GCA TG-3’), StAR (forward 5’-GCT CTC TAC TCG GTT CTC-3’ and reverse 5’-GCT GAC TCT CCT TCT TCC-3’), TSPO (forward 5’-GCA GAT TCC GTG ATT ACA GTG-3’ and reverse 5’- TCC TCC TCG TCG TCA TCG-3’), HSL (forward 5’-CAC TAC AAA CGC AAC GAG AC-3’ and reverse 5’-CCA GAG ACG ATA GCA CTT CC-3’), SR-BI (forward 5’-CCA TCC TCA CTT CCT CAA-3’ and reverse 5’-CCA CAG GCT CAA TCT TCC-3’, LDLR (forward 5’-ACG GTG GAG ATA GTG ACA ATG-3’ and reverse 5’-AGA CGA GGA GCA CGA TGG-3’). mRNA expression of StAR, TSPO, HSL, SR-BI, and LDLR was normalized to the transcript levels of β-actin and calculated using the delta-delta cycle threshold (ΔΔCT) method.

2.5. Measurement of Inositol 1, 4, 5-Trisphosphate (IP3)

Cells were treated with 1 µM S1P or 1 mM Bt2cAMP for 5 to 60 min. Cells were harvested into PBS and centrifuged 5 min at 4,000 rpm. Supernatant was removed and the cells resuspended in 20% (w/v) perchloric acid on ice for 30 min. Cells were centrifuged and the supernatants neutralized with HEPES-KOH solution. Amounts of inositol 1, 4, 5-trisphosphate in each sample were measured by a radioreceptor assay with the D-myo-inositol 1, 4, 5-triphosphate [3H] assay kit (TRK 1000, Amersham Biosciences, Piscataway, NJ).

2.6. Intracellular Ca2+Quantification

H295R cells were subcultured onto coverslips for 24 h and intracellular calcium was labeled with Fluo 3/AM (Burnier M et al., 1994, Orlicky J et al., 2004, Tsai J-H et al., 2010). Fluo 3/AM (5 µM) was loaded into the cells in serum-free medium for 1 hour at 37 °C. 0.02% pluronic acid (Sigma, St. Louis, MO) was added to the medium to disperse Fluo 3/AM (EMD Biosciences). After the incubation, cells were washed three times with medium. Then cells were placed in a chamber with 0.5 ml medium for fluorescence imaging. Fluorescence images were acquired using UltraVIEW Vox Spinning Disk Confocal Microscope (Perkin Elmer Inc., Waltham, MA). Intensity ratio of Fluo 3/AM was quantified using the Volocity software (Perkin Elmer Inc.).

2.7. Western Blotting (WB)

For analysis of StAR protein expression, cells were treated with 1 µM S1P (0 – 18 h). Phospho-ERK1/2 (pERK1/2) and total ERK2 expression was assessed in lysates purified from cells that were serum-starved for 30 h, pretreated with 10 µM KN-93, or 5 pg/ml PTX for 1 h followed by treatment with 1 µM S1P. Phospho-CamKII (pCamKII) and total CamKII expression was assessed in lysates purified from cells that were pretreated with 10 µM VPC23019, 10 µM U0126, or 5 pg/ml PTX for 1 h followed by treatment with 1 µM S1P for 30 min. To assess the phosphorylation status of HSL, cells were treated for 15, 30, or 60 min with 1 µM S1P. Cells were harvested into RIPA buffer (1X PBS, 1% Nonidet P-40, 0.5% sodium deoxycholate, 0.1% SDS) containing 1X Protease Inhibitor Cocktail Set I (EMD Biosciences) and lysed by sonication (one 2 sec burst) followed by incubation on ice for 30 min. Lysates were centrifuged at 12,000 rpm for 15 min at 4°C and the supernatant collected for analysis by SDS-PAGE. Aliquots of each sample (30 µg of protein) were run on 8% SDS-PAGE gels and transferred to Immobilon FL polyvinylidene difluoride (PVDF) membranes (Millipore). Blots were probed with anti-phospho-ERK1/2 (sc-7383, Santa Cruz), ERK2 (sc-154, Santa Cruz), StAR (sc-25806, Santa Cruz), phospho-Ser563-HSL (4139, Cell Signaling), HSL (sc-25843, Santa Cruz), phospho- Thr286-CamKII (06-881, Millipore), CamKII (04-1079, Millipore), or GAPDH (sc-25778, Santa Cruz). Expression was detected using an ECF western blotting kit (GE Healthcare, Piscataway, NJ) and visualized by scanning the blots on a VersaDoc 4000 imager (Bio-Rad). Protein concentrations were determined by the bicinchoninic acid (BCA) protein assay (Pierce, Rockford, IL).

2.8. Statistical Analysis

One-way ANOVA and Tukey-Kramer multiple comparison test were performed using GraphPad InStat software (GraphPad Software Inc., San Diego, CA). Significant differences from compared value were defined as p<0.05 and denoted by asterisks (*) and carats (^).

3. Results

3.1. S1P stimulates cortisol biosynthesis

As previously discussed, we have demonstrated that ACTH and Bt2cAMP rapidly stimulate sphingolipid metabolism (Ozbay T et al., 2004) and S1P secretion from H295R cells (Ozbay T et al., 2006). Additionally, S1P induces CYP17 transcription (Ozbay T et al., 2006). Therefore, in order to investigate the role of S1P in the acute phase of steroid hormone biosynthesis, we determined the effect of S1P on cortisol biosynthesis in H295R cells. Cells were treated for time points ranging from 3 to 48 h and the concentrations of steroid hormone secreted into the media were quantified as described in Materials and Methods. As shown in Fig. 1A, S1P maximally increased cortisol production by 2.5-fold at the 12-h time point. While S1P resulted in a significant increase in cortisol production at the 48 h time point (Fig. 1A), the bioactive sphingolipid was unable to evoke a sustained increase in DHEA secretion over the same time period (Fig. 1B). Bt2cAMP treatment robustly stimulated the secretion of both cortisol and DHEA at all time points assayed, with a 7-fold increase in cortisol production and a 6-fold increase in DHEA biosynthesis at the 24 h time point.

Fig. 1. S1P increases cortisol secretion.

Fig. 1

H295R cells were treated for time periods ranging from 3 to 48 h with Bt2cAMP (0.4 mM) or S1P (1 µM) as described in Materials and Methods. Cortisol (A) or DHEA (B) released into the media was quantified by ELISA. Values represent the mean ± SEM of 3 experiments, each performed in triplicate and normalized to the total cellular protein content. Asterisks (*) denote p < 0.05 for agonist-stimulated hormone production versus vehicle-treated controls.

3.2. S1P induces the mRNA expression of various acute phase steroidogenic genes

To characterize the mechanism by which S1P acutely increases cortisol secretion, we determined the effect of S1P on the mRNA expression of multiple genes involved in the acute steroidogenic response. StAR is essential for the delivery of free cholesterol into the inner mitochondrial membrane (Caron KM et al., 1997b, Hauet T et al., 2002), the rate-limiting step in steroid hormone biosynthesis. To determine the effect of S1P on the mRNA expression of StAR, H295R cells were treated for 0.5-, 1- or 2 h with 0.4 mM Bt2cAMP or 1 µM S1P and total RNA was isolated for quantification by qRT-PCR. As shown in Fig. 2A, S1P and Bt2cAMP induced StAR mRNA expression by 2.0- and 1.8-fold after 1 h treatment, respectively. S1P further increased StAR expression at the 2-h time point by 5.0-fold, as compared to the 14-fold induction in StAR mRNA expression elicited by Bt2cAMP (Fig. 2A).

Fig. 2. S1P acutely increases steroidogenic gene expression.

Fig. 2

(A) H295R cells were treated with 0.4 mM Bt2cAMP or 1 µM S1P for 30 min to 2 h. StAR mRNA levels were quantified by qRT-PCR and normalized to the mRNA expression of β-actin. (B) H295R cells were treated with 0.4 mM Bt2cAMP or 1 µM S1P for 2 h and mRNA expression of SR-BI, HSL, LDLR, and TSPO was quantified by qRT-PCR and normalized to the mRNA levels of β-actin. Data graphed are expressed as percent of control group mean and represent the mean ± SEM of 3 experiments, each performed in triplicate. Asterisk (*) denotes p < 0.05 for Bt2cAMP- or S1P-stimulated mRNA expression versus vehicle-treated control.

In addition to examining StAR transcription, we quantified the mRNA expression of four additional genes that are involved in the biosynthesis of cortisol: TSPO, SR-BI, LDLR and HSL. TSPO forms a macromolecular complex with StAR to facilitate the import of cholesterol into the inner mitochondrial membrane (Papadopoulos V, 1993). SR-BI mediates the selective uptake of cholesteryl esters (CE) from high-density lipoprotein particles (HDL) in rodents and humans (Briand O et al., 2003, MA, 2009, Pilon A et al., 2003) and HSL has neutral cholesteryl ester hydrolase activity and is critical for hydrolyzing internalized CEs (Kraemer FB et al., 2004). Importantly, cAMP has been shown to regulate both HSL enzymatic activity and SR-BI mRNA expression (Krintel C et al., 2008). LDLR enables receptor-mediated endocytic delivery of CEs from LDL particles (Kraemer FB, 2007). In humans, LDLR provides most of the cholesterol necessary for steroid hormone production (Carr BR and Simpson ER, 1981). We examined the mRNA expression of these genes in H295R cells that were treated with 1 µM S1P or 0.4 mM Bt2cAMP for 2 h. As shown in Fig. 2B, S1P induced the mRNA expression of SR-BI, LDLR, and TSPO by 2.5-, 4.7-, and 4.5-fold, respectively. The mRNA expression of HSL was not affected by S1P; though Bt2cAMP significantly increased the expression of all four genes.

3.3. S1P-mediated induction of StAR gene expression is pertussis toxin sensitive

As discussed earlier, many of the effects elicited by S1P are mediated through G-protein coupled S1P receptors (S1PR1–5) that activate multiple downstream signaling cascades (Alvarez SE et al., 2007, Sanchez T and Hla T, 2004, Taha TA et al., 2004). In order to define the signaling pathway that mediates S1P-induced StAR gene expression, we determined the effect of the Gαi inhibitor PTX on S1P-dependent StAR gene expression. H295R cells were pretreated with 5 pg/ml PTX, followed by stimulation with 1 µM S1P for 2 h. As shown in Fig. 3A, PTX prevented S1P-induced expression of StAR. Because multiple S1PR isoforms have been reported to couple to Gαi pathways (Hla T et al., 2001, Taha TA et al., 2004, Windh RT et al., 1999, Yamazaki Y et al., 2000), we next sought to identify the specific S1PR that mediates S1P-induced StAR transcription by pre-treating H295R cells with the S1PR1/S1PR3 antagonist VPC23019 followed by S1P stimulation. VPC23019 abolished S1P-mediated StAR gene expression (Fig. 3A). Taken together, these data suggest that S1P acutely activates StAR transcription through activation of Gαi-protein coupled receptors S1PR1 and/or S1PR3. Because we have previously found that ACTH/cAMP signaling promotes the secretion of S1P from H295R cells (Ozbay T et al., 2004, Ozbay T et al., 2006), we assessed the effect of PTX and VPC23019 on ACTH/cAMP-stimulated StAR mRNA expression and found that neither inhibitor affected the ability of Bt2cAMP to induce StAR expression (Fig. 4B). However, VPC23019 significantly reduced ACTH-dependent StAR mRNA levels by 34% (Fig 4B). The ability of the S1PR1/3 antagonist to partially repress ACTH-stimulated StAR expression is consistent with a second messenger role for S1P in ACTH-dependent steroidogenesis. To further define the relationship between S1P-dependent StAR transcription and ACTH signaling, we examined the effect of S1P on ACTH- and Bt2cAMP-induced on StAR mRNA expression. Neither Bt2cAMP nor ACTH had an additive effect on the S1P response (Fig. 3C). However, unexpectedly ACTH-stimulated StAR transcription was suppressed in the presence of S1P (Fig. 3C).

Fig. 3. S1P-stimulated StAR gene expression occurs through receptor-mediated G αi signaling.

Fig. 3

(A) H295R cells were treated for 2 h with 1 µM S1P in the presence or absence of 10 µM VPC23019 or 5 pg/ml PTX. Total mRNA was isolated and StAR mRNA expression was quantified by qRT-PCR and normalized to the mRNA levels of β-actin. Data graphed represent the mean ± SEM of 3 experiments, each performed in triplicate. Asterisk (*) denotes p < 0.05 compared to vehicle-treated controls. (B) H295R cells were pre-treated with 10 µM VPC23019 or 5 pg/ml PTX for 18 h and then stimulated for 2 h with 50 nM ACTH or 0.4 mM Bt2cAMP. Total mRNA was isolated and StAR mRNA expression was quantified by qRT-PCR and normalized to the mRNA levels of β-actin. Data graphed represent the mean ± SEM of 3 experiments, each performed in triplicate. Asterisks (*) denote p < 0.05 compared to vehicle-treated control and carat (^) denotes statistically different from ACTH-treated group (p < 0.05). (C) H295R cells were treated for 2 h 1 µM S1P in the presence or absence of 0.4 mM Bt2cAMP or 50 nM ACTH. Total mRNA was isolated and StAR mRNA expression was quantified by qRT-PCR and normalized to the mRNA levels of β-actin. Data graphed represent the mean ± SEM of 3 experiments, each performed in triplicate. Asterisk (*) and carat (^) denote statistically different (p < 0.05) compared to vehicle-treated controls or ACTH-treated, respectively.

Fig. 4. S1P-mediated StAR transcription is dependent on PLC activation.

Fig. 4

(A) H295R cells were pretreated with U73122 or U73343 (10 µM) and then stimulated for 2 h with 1 µM S1P. Total RNA was isolated and StAR mRNA expression assessed by qRT-PCR and normalized to the mRNA levels of β-actin. Data are graphed as percent of control group mean and represent the mean ± SEM of 3 experiments, each performed in triplicate. Asterisks (*) denote statistically different from the control group (p < 0.05). (B) H295R cells were treated with 1 µM S1P or 0.4 mM Bt2cAMP for time periods ranging from 5 to 60 min. IP3 content was determined by radioreceptor assay as described in Materials and Methods. Data graphed are expressed as pmol of IP3 per milligram of protein and represent the mean ± SEM of 2 experiments, each performed in triplicate. Asterisks (*) denote p < 0.05 of S1P-treated cells versus vehicle-treated controls. (C) Representative immunofluorescence live cell images of cytosolic Ca2+ in H295R cells treated for 0 to 15 min with S1P (top panel) or Bt2cAMP (bottom panel). Cytosolic Ca2+ was visualized using Fluo 3/AM (green). (D) Quantification of cytosolic Ca2+ in H295R cells treated for 5 to 15 min with 0.4 mM Bt2cAMP or 1 µM S1P using the Volocity software. Data are graphed as mean intensity over area and represent the mean ± SEM of 4 experiments, each performed in triplicate.

3.4. S1P-dependent StAR transcription requires PLC activation

Signaling through Gαi affects multiple downstream effectors including adenylyl cyclase, the small G protein Rac, Src, PI3K/Akt, MAPK, and PLC. The inhibitory effect of PTX and VPC23019 on S1P-mediated StAR mRNA expression suggested that the stimulatory effect of S1P on gene expression occur through the interaction of S1P with S1PR1/3. Therefore, in order to identify effectors in the signaling pathway downstream of Gαi, we determined the effect of PLC inhibition on S1P-stimulated StAR transcription. H295R cells were pre-treated with 10 µM U73122 followed by treatment with 1 µM S1P for 2 h. PLC inhibition completely abolished S1P-induced StAR mRNA expression, whereas the inactive PLC analog U73343 had no significant effect on the ability of S1P to increase StAR mRNA expression (Fig. 4A).

3.5. S1P stimulates IP3 release and increases cytoplasmic Ca2+

Based on the findings described above demonstrating that S1P-stimulated StAR gene transcription is dependent on PLC activity, we quantified the amount of IP3 release in response to S1P. H295R cells were treated with 1 µM S1P for 5 to 60 min and IP3 intracellular levels were determined as described in Materials and Methods. As shown in Fig. 4B, S1P significantly increased intracellular IP3 by 2.5- and 4.8-fold after 5 and 15 min, respectively, compared to untreated controls. IP3 levels in S1P-treated cells gradually decreased after 15 min of stimulation, although they remained significantly higher than untreated controls. Notably, Bt2cAMP had no significant effect on IP3 release (Fig. 4B), suggesting that the mechanism by which S1P induces StAR expression exhibits distinct elements when compared to Bt2cAMP. The increase in IP3 was concomitant with a 2.9- and 5.8-fold increase in cytosolic Ca2+ after 5 and 15 min of S1P stimulation, respectively (Fig. 4D). Bt2cAMP also stimulated an increase in cytosolic Ca2+, however, the magnitude of the response was lower than the increase elicited by S1P.

3.6. Induction of StAR mRNA expression requires CamKII and ERK1/2 activation

The increases in IP3 and Ca2+ led us to postulate that S1P-stimulated StAR expression requires CamKII. To test this hypothesis, H295R cells were treated for 2 h with 1 µM S1P in the presence or absence of the CamKII inhibitor KN-93 (10 µM) and StAR mRNA transcript levels were quantified by qRT-PCR. As shown in Fig. 5A, inhibiting CamKII activity attenuated S1P-induced StAR mRNA expression. Consistent with this finding, S1P stimulation activated CamKII after 30 min, and pre-treatment with VPC23019 and PTX abrogated S1P-stimulated CamKII phosphorylation (Fig.5B). Because CamKs can crosstalk with many intracellular pathways, including MAPK signaling (Blanc A et al., 2004, Li N et al., 2009, Nguyen A et al., 2004, Schmitt JM et al., 2004), we next investigated the effect of S1P on the activation of ERK1/2. As shown in Fig. 5C, S1P activates ERK1/2 after 30 min of stimulation. To determine if ERK1/2 activation is upstream of CamKII, the phosphorylation status of ERK1/2 was quantified by western blotting from H295R cells pre-treated with 10 µM KN-93 and stimulated with 1 µM S1P for 15 or 30 min. As shown in Fig. 5D, pre-treatment with KN-93 prevented ERK1/2 phosphorylation in S1P-treated cells. Additionally, pre-treatment of H295R cells with the MEK1 inhibitor U0126 completely ablated S1P-induced StAR gene expression (Fig. 5E). Taken together, our results indicate that S1P-dependent StAR transcription requires the sequential activities of CamKII and ERK1/2.

Fig. 5. S1P-stimulated StAR transcription requires CamKII and ERK1/2 activation.

Fig. 5

(A) H295R cells were pretreated with 10 µM KN-93 and then stimulated with with 1 µM S1P for 2 h. Cells were harvested and total RNA isolated for analysis of StAR mRNA expression by qRT-PCR and normalized to the mRNA levels of β-actin. Data graphed are expressed as percent of control group mean and represent the mean ± SEM of 3 experiments, each performed in triplicate. Asterisk (*) denotes p < 0.05 for S1P-stimulated mRNA expression versus vehicle-treated control. (B) H295R cells were pre-treated with 10 µM VPC23019 or 5 ng/mL PTX for 1 h, and then treated with 1 µM S1P for 30 min. Cell lysates were harvested and separated by SDS-PAGE. CamKII activity was assessed in cell lysates by western blotting using an antibody against phospho-Thr286-CamKII (top panel) or CamKII (bottom panel). Data are expressed as fold change in phospho-CamKII over control group mean ± STD DEV, normalized to total CamKII, and represent the densitometric analysis of western blots from three independent experiments, each performed in triplicate. (C) H295R cells were serum-starved for 30 h, treated for 15 or 30 min with 1 µM S1P, and cell lysates were harvested and separated by SDS-PAGE followed by western blotting using anti-phospho-ERK1/2 (top panel) and ERK2 (bottom panel) antibodies. Data are expressed as fold change in phospho-ERK1/2 over control group mean ± STD DEV, normalized to total ERK2 and represent the densitometric analysis of western blots from three independent experiments, each performed in triplicate. (D) H295R cells were serum-starved for 30 h, pre-treated with 10 µM KN-93 for 1h, and then treated for 15 or 30 min with 1 µM S1P. Some cells were treated with 25 ng/ml EGF for 10 min. ERK activity was assessed in cell lysates by western blotting using an antibody against phospho-ERK1/2 (top panel) or ERK2 (bottom panel). Data are expressed as fold change in phospho-ERK1/2 over control group mean ± STD DEV, normalized to total ERK2 and represent the densitometric analysis of western blots from three independent experiments, each performed in triplicate. (E) H295R cells were treated for 2 h with 1 µM S1P in the presence or absence of 10 µM U0126. Total RNA was isolated and StAR mRNA expression was quantified by qRT-PCR and normalized to the mRNA levels of β-actin. Data graphed represent the mean ± SEM of 3 separate experiments, each performed in triplicate. Asterisk (*) denotes statistically significant difference (p < 0.05) compared to vehicle-treated control. (F) H295R cells were serum-starved for 30 h, pre-treated with 5 pg/ml PTX for 1 h, and then treated for 15 or 30 min with 1 µM S1P. Some cells were treated with 25 ng/ml EGF for 10 min (positive control). Cell lysates were harvested and separated by SDS-PAGE followed by western blotting using antibodies against phospho-ERK1/2 (top panel) or ERK2 (bottom panel). Data are expressed as fold change in phospho-ERK1/2 over control group mean ± STD DEV, normalized to total ERK2 and represent the densitometric analysis of western blots from three independent experiments, each performed in triplicate.

3.7. S1P-mediated ERK1/2 activation is dependent on G αi signaling

Our data show that ERK1/2 activation is required for S1P-dependent StAR gene transcription and that CamKII is upstream of ERK1/2. Moreover, Gαi and PLC are also required for the S1P response. Therefore, in order to determine if ERK1/2 phosphorylation occurs through activation of Gαi signaling, we carried out western blotting analysis of H295R cells pre-treated with 5 pg/ml PTX and stimulated with 1 µM S1P for 15 or 30 min. As shown in Fig. 5F, suppression of Gi signaling by PTX prevents S1P-induced ERK1/2 phosphorylation.

3.8. S1P acutely increases StAR protein expression and HSL phosphorylation

To determine if the increase in StAR transcription resulted in an increase in protein expression, we quantified StAR protein levels in S1P-treated H295R cells by western blotting. As shown in Fig. 6A and B, S1P significantly increased StAR protein expression by 1.9- and 2.5-fold after 3 and 6 h of stimulation, respectively. Significantly, although S1P had no effect on HSL mRNA (Fig. 2B) or protein expression (data not shown), densitometric analysis of western blots revealed that S1P stimulated the phosphorylation of HSL at Ser563, a PKA target site that stimulates HSL activity (Carmen GY and Victor SM, 2006, Garton AJ et al., 1988, Stralfors P and Belfrage P, 1983, Stralfors P et al., 1984), by 2.4- and 2.8- fold after 15 and 30 min, respectively (Fig. 6C and 6D).

Fig. 6. S1P increases StAR protein expression and HSL phosphorylation.

Fig. 6

(A). H295R cells were treated with 1 µM S1P for 0 to 18 h and cell lysates were harvested and separated by SDS-PAGE followed by western blotting analysis using anti-StAR and anti-GAPDH antibodies. (B). Densitometric analysis of western blots of StAR protein expression, normalized to GAPDH protein content in cells treated for 0 to 18 h with 1 µM S1P. Data graphed represent the mean ± SEM of 2 separate experiments, each done in triplicate. Asterisks (*) denote statistically significant difference (p < 0.05) versus untreated controls. (C) H295R cells were treated with 1 µM S1P for 15 to 60 min and cell lysates were harvested and separated by SDS-PAGE followed by western blot analysis using antibodies against phospho-Ser563-HSL (top panel) and anti-HSL (bottom panel). (D) Data are graphed as fold change in phospho-HSL over control group mean ± STD DEV, normalized to total HSL and represent the densitometric analysis of western blots from three independent experiments, each performed in triplicate.

4. Discussion

ACTH promotes glucocorticoid production by coordinating a series of spatially and temporally distinct cellular processes. In what is classically known as the acute phase of steroidogenesis, ACTH-stimulated PKA activation leads to an increase in the hydrolysis of cholesterol esters and the subsequent trafficking of this substrate to the inner mitochondrial membrane, the site of the first enzymatic reaction in the steroid hormone biosynthetic pathway. During the acute phase, ACTH signaling rapidly induces the expression of StAR (Clark BJ et al., 1994, Stocco D et al., 2005) and the activation of HSL (Cook KG et al., 1982). Based on our previous studies showing ACTH-stimulated sphingolipid metabolism and S1P secretion (Ozbay T et al., 2006), we sough to define the role of S1P in the acute phase of steroidogenesis. We report herein that S1P induces the transcription of multiple acute steroidogenic genes, stimulates the phosphorylation of HSL at Ser563, and increases cortisol production. Further, we show that S1P-stimulated StAR gene expression occurs through an S1PR1/SIPR3-mediated Gαi signaling pathway that involves PLC, CamKII, and ERK1/2 activation.

Although ACTH is the main regulator of cortisol biosynthesis, multiple additional signaling molecules can modulate steroidogenic gene transcription, including growth factors, prostaglandins, and phospholipids (Doi J et al., 2001, Schinner S et al., 2007, Shah BH et al., 2005, Watanabe M et al., 2006). S1P regulates aldosterone secretion in zona glomerulosa adrenocortical cells (Brizuela L et al., 2007, Brizuela L et al., 2006) and stimulates cortisol secretion in zona fasciculata bovine adrenal cells (Rabano M et al., 2003). We show that S1P is a novel modulator of acute steroid hormone biosynthesis in human adrenocortical cells. Although the effect of S1P and Bt2cAMP were similar at earlier time points, the magnitude of the response elicited by Bt2cAMP was significantly greater at later time points (Fig. 1A). Notably, S1P only increased DHEA secretion transiently at the 6 and 12 h time points (Fig. 1B). Because the biosynthesis of cortisol requires the movement of intermediate metabolites between the ER and mitochondria and we have previously shown that ACTH/cAMP-mediated mitochondrial movement plays a pivotal role in glucocorticoid biosynthesis (Li D and Sewer MB, 2010), it is tempting to speculate that S1P is stimulating cortisol production by facilitating substrate delivery to mitochondria. Indeed, we have found the S1P increases the rate of mitochondrial trafficking (Sewer MB and Li D, 2008).

During the acute phase of steroidogenesis, ACTH-dependent cholesterol mobilization is mirrored by an increase in the transcription of genes essential for its uptake, transport, and de-esterification, including StAR and TSPO (Thomson M, 1997). Our data show that S1P rapidly increases the transcription of TSPO, SR-BI, LDLR, and StAR (Fig. 2). Significantly, although S1P had no effect on HSL mRNA expression (Fig. 2B), phosphorylation of HSL at Ser563 was acutely stimulated S1P (Fig. 6C). As discussed earlier, the activity of HSL is regulated by phosphorylation at multiple residues (Anthonsen MW et al., 1998, Degerman E et al., 1990, Garton AJ and Yeaman SJ, 1990, Greenberg AS et al., 2001). However, the phosphorylation of HSL at Ser563 by PKA (Anthonsen MW et al., 1998) triggers HSL translocation to the surface of lipids droplets where it hydrolyzes cholesteryl esters (Clifford GM et al., 2000, Egan JJ et al., 1992). Moreover, HSL has been found to interact with StAR (Shen WJ et al., 2003). Therefore, our data identify multiple roles for S1P in facilitating increased substrate availability and delivery for cortisol biosynthesis (Fig. 7).

Fig. 7. Proposed model for S1P-mediated cortisol biosynthesis in H295R cells.

Fig. 7

S1P binds to S1PR1 and/or SIPR3 and activates Gαi. Gαi couples to PLC, thereby increasing intracellular IP3 and subsequently cytosolic Ca2+. Intracellular Ca2+ accumulation activates CamKII, which in turn mediates the phosphorylation of ERK1/2. Acute activation of the S1P signaling pathway culminates in an induction of StAR gene expression and an increase in cortisol secretion. S1P also stimulates PKA-mediated phosphorylation of HSL at Ser563, which increases cholesterol ester hydrolysis and substrate availability for cortisol production.

Most of the physiological effects elicited by S1P are the result of binding to tissue-specific S1P receptors (S1PR1–5) (Alvarez SE et al., 2007, An S et al., 1998). In this manner, S1P has been shown to activate PLC, PI3K/Akt, and MAPK signaling as well as Ca2+ mobilization (Brizuela L et al., 2007, Brizuela L et al., 2006, Graler MH et al., 2003, Olivera A and Spiegel S, 1993, Spiegel S and Milstien S, 2003, Suhaiman L et al., 2010, Yamazaki Y et al., 2000). Our results demonstrate that S1P-dependent StAR gene expression occurs through a pertussis toxin-sensitive Gαi-coupled receptor (Fig. 3A). Moreover, because the S1PR1/SIPR3 antagonist VPC23019 prevented S1P-induced gene expression, it is likely that S1P signals through one or both of these receptor isoforms. Notably, both isoforms are expressed in the H295R cells line (Ozbay T et al., 2006) and can couple to Gαi (Ancellin N and Hla T, 1999, Brizuela L et al., 2007, van Brocklyn JR et al., 1998, Windh RT et al., 1999). Further, S1P and Bt2cAMP did not have an additive effect on StAR mRNA expression (Fig. 3C), suggesting that these stimuli activate a common downstream effector. Conversely, S1P abrogated ACTH-stimulated StAR transcription (Fig. 3C). Because our data suggests that S1P signals through Gαi to induce StAR mRNA expression (Fig. 3A) and ACTH activates a Gαs-cAMP-PKA signaling pathway (Mountjoy KG et al., 1992), it is tempting to speculate that S1P-mediated Gαi activation downregulates ACTH-dependent Gαs signaling by inhibiting adenylyl cyclase (Taussig R et al., 1993). Gαi-mediated downregulation of cAMP production by S1P has been previously reported (Jongsma M et al., 2009, Jun DJ et al., 2006, Means CK et al., 2008, Sensken SC et al., 2008).

Many reports support the activation of Ca2+ mobilization by S1P through GPCR-mediated PLC activation (Bornfeldt KE et al., 1995, Noh SJ et al., 1998, Okajima F et al., 1996, Suhaiman L et al., 2010). Consistent with these findings, we show that the transcriptional response evoked by S1P requires PLC activation and results in the accumulation of cytoplasmic IP3 and Ca2+ (Fig. 4). It should be noted that S1P has been shown to increase cytosolic Ca2+ by stimulating release from intracellular stores and by activating Ca2+ influx (Bischoff A et al., 2001, Coussin F et al., 2002, Mathieson FA and Nixon GF, 2006). Although our studies do not distinguish between these two sources, the increase in IP3 levels elicited by S1P (Fig. 4B) suggest that elevated cytosolic Ca2+ in response to S1P stimulation stems, at least in part, from intracellular stores. Notably, even though Bt2cAMP also elevated cytosolic Ca2+ levels (Fig. 4D), it did so independently of cellular IP3 levels (Fig. 4B), which supports the notion that the molecular mechanism of S1P-mediated glucocorticoid production displays unique components when compared to ACTH/cAMP signaling.

In order to further characterize the signaling molecules involved in the S1P transcriptional response, we show that downstream of PLC, CamKII and ERK1/2 activation are required events for S1P-dependent StAR gene transcription (Fig. 5). CamKII is a well-established Ca2+-activated kinase, thus its activation by increased cytoplasmic Ca2+ is plausible. In addition, S1P has been shown to activate this enzyme in vascular smooth muscle cells (Mathieson FA and Nixon GF, 2006). Consistent with previous reports on the crosstalk between CamK and the MAPK/ERK pathway (Illario M et al., 2003, Li N et al., 2009, Nguyen A et al., 2004, Schmitt JM et al., 2004) and the activation of MAPK signaling by S1P (Brizuela L et al., 2007, Rakhit S et al., 1999, Wu J et al., 1995), we show that S1P rapidly activated ERK1/2 (Fig. 5B) in a CamKII-dependent manner (Fig. 5C). Further, ERK1/2 activation was required for S1P-dependent StAR mRNA expression (Fig. 5D). Importantly, MAPK/ERK activation in response to S1PR1 and S1PR3 couples to Gαi (Brizuela L et al., 2007, Lee MJ et al., 1996, van Brocklyn JR et al., 1998, Windh RT et al., 1999, Zondag GC et al., 1998), further supporting our findings.

ERK1/2-mediated StAR gene expression is well established (Gyles SL et al., 2001, Manna PR et al., 2006, Manna PR et al., 2007) and both PKA and PKC signaling can activate ERK1/2 (Manna PR et al., 2007). Importantly, PKA (reviewed in (Arakane F et al., 1997, Stocco DM, 2000)) and PKC (Fiedler EP et al., 1999, Manna PR et al., 2007) signaling regulate StAR transcription. Our data identify S1P as a novel regulator of StAR gene expression via the activation of the MAPK/ERK pathway. Regulation of StAR gene transcription involves multiple transcription factors including the bZIP family of transcription factors (cAMP-regulatory element binding protein (CREB)/CRE modulator (CREM)/activating transcription factor (ATF)), which binds to conserved 5’-CRE half-sites (Manna PR et al., 2009b, Manna PR et al., 2003). In addition, increased Ca2+ and subsequent activation of CamK leads to CREB phosphorylation in a MAPK-dependent manner (Wu GY et al., 2001). Further, S1P was shown to partially activate CREB in this manner (Coussin F et al., 2003). Thus, it is tempting to speculate that S1P-induced StAR transcription is mediated, at least in part, through ERK1/2-dependent CREB phosphorylation. Moreover, because ERK1/2 regulate StAR function by phosphorylation at Ser232 (Poderoso C et al., 2009), it is possible that S1P stimulation triggers StAR phosphorylation in addition to inducing its gene expression. Future studies are needed to test these hypotheses as well as uncover other regulatory proteins involved in S1P-dependent acute steroidogenic gene expression.

We (Ozbay T et al., 2006) and others (Brizuela L et al., 2007, Brizuela L et al., 2006, Rabano M et al., 2003) have reported that S1P mediates steroid hormone biosynthesis by activating various intracellular cascades. Rabano et al. (2003) demonstrated that S1P-mediated cortisol secretion in bovine adrenocortical cells is dependent on PKC and intracellular Ca2+ in a pertussis toxin-sensitive manner. Our results support a similar mechanism of action for S1P in human H295R cells, although we demonstrate that PLC activation is also required. Importantly, we show that S1P induces the phosphorylation of HSL at Ser563, providing evidence that S1P regulates steroidogenesis by acting via multiple. Given that ACTH/cAMP promotes the secretion of S1P from H295R cells (Ozbay T et al., 2006), it is plausible that adrenocortical cells might utilize an ACTH/S1P feed-forward mechanism to facilitate rapid steroidogenic output and fine-tune sustained hormone production. In summary, we demonstrate that S1P acts at multiple levels to promote cortisol production.

Highlights.

  • >

    S1P induces StAR, TSPO, LDLR, and SR-BI expression and cortisol secretion.

  • >

    StAR expression requires a Ca2+-dependent pathway that involves PLC, CamKII, and ERK1/2.

  • >

    S1P rapidly stimulates the phosphorylation of hormone sensitive lipase (HSL) at Ser563.

Acknowledgement

This study was supported by NIH DK084178 (to M.B.S.) and by an American Heart Association Predoctoral Fellowship (to N.C.L.).

Abbreviations

S1P

sphingosine-1-phosphate

ACTH

adrenocorticotropin

PLC

phospholipase C

StAR

steroidogenic acute regulatory protein

TSPO

translocator protein

CamKII

Ca2+/calmodulin-dependent kinase II

IP3

inositol 1,4,5-triphosphate

PKA

protein kinase A

PKC

protein kinase C

HSL

hormone sensitive lipase

LDLR

low-density lipoprotein receptor

SR-BI

scavenger receptor class B type I

SF-1

steroidogenic factor 1

MAPK

mitogen-activated protein kinase

ERK

extracellular signal-regulated kinase

PI3K

phosphoinositide 3-kinase

S1PR

S1P receptor

Bt2cAMP

dibutyryl-cAMP

PTX

pertussis toxin

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  1. Alvarez SE, Milstien S, Spiegel S. Autocrine and paracrine roles of sphingosine-1-phosphate. Trends Endocrinol Metab. 2007;18:300–307. doi: 10.1016/j.tem.2007.07.005. [DOI] [PubMed] [Google Scholar]
  2. An S, Goetzl EJ, Lee H. Signaling mechanisms and molecular characteristics of G protein-coupled receptors for lysophosphatidic acid and sphingosine-1-phosphate. J Cell Biochem Suppl. 1998;30–31:147–157. [PubMed] [Google Scholar]
  3. Anada Y, Igarashi Y, Kihara A. The immunomodulator FTY720 is phosphorylated and released from platelets. Eur J Pharmacol. 2007;568:106–111. doi: 10.1016/j.ejphar.2007.04.053. [DOI] [PubMed] [Google Scholar]
  4. Ancellin N, Hla T. Differential pharmacological properties and signal transduction of the sphingosine-1-phosphate receptors EDG-1, EDG-3, and EDG-5. J Biol Chem. 1999;274:18997–19002. doi: 10.1074/jbc.274.27.18997. [DOI] [PubMed] [Google Scholar]
  5. Anthonsen MW, Ronnstrand L, Wernstedt C, Degerman E, Holm C. Identification of novel phosphorylation sites in hormone-sensitive lipase that are phosphorylated in response to isoproterenol and govern activation properties in vitro. J Biol Chem. 1998;273:215–221. doi: 10.1074/jbc.273.1.215. [DOI] [PubMed] [Google Scholar]
  6. Arakane F, King SR, Du Y, Kallen CB, Walsh LP, Watari H, Stocco DM, Strauss JF., 3rd Phosphorylation of Steroidogenic Acute Regulatory Protein (StAR) Modulates Its Steroidogenic Activity. J Biol Chem. 1997;272:32656–32662. doi: 10.1074/jbc.272.51.32656. [DOI] [PubMed] [Google Scholar]
  7. Arlt W, Stewart PM. Adrenal corticosteroid biosynthesis, metabolism, and action. Endocrinol Metab Clin North Am. 2005;34:293–313. doi: 10.1016/j.ecl.2005.01.002. [DOI] [PubMed] [Google Scholar]
  8. Bischoff A, Finger J, Michel MC. Nifedipine inhibits sphinogosine-1-phosphate-induced renovascular contraction in vitro and in vivo. Naunyn Schmiedebergs Arch Pharmacol. 2001;364:179–182. doi: 10.1007/s002100100446. [DOI] [PubMed] [Google Scholar]
  9. Blanc A, Pandey NR, Srivastava AK. Distinct roles of Ca2+, calmodulin, and protein kinase C in H2O2-induced activation of ERK1/2, p38 MAPK, and protein kinase B signaling in vascular smooth muscle cells. Antioxid Redox Signal. 2004;6:353–366. doi: 10.1089/152308604322899422. [DOI] [PubMed] [Google Scholar]
  10. Bornfeldt KE, Graves LM, Raines EW, Igarashi Y, Wayman G, Yamamura S, Yatomi Y, Sidhu JS, Krebs EG, Hakomori S, Ross R. Sphingosine-1-phosphate inhibits PDGF-induced chemotaxis of human arterial smooth muscle cells: spatial and temporal modulation of PDGF chemotactic signal transduction. J Cell Biol. 1995;130:193–206. doi: 10.1083/jcb.130.1.193. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Briand O, Lestavel S, Pilon A, Torpier G, Fruchart JC, Clavey V. SR-BI does not require raft/caveola localisation for cholesteryl ester selective uptake in the human adrenal cell line NCI-H295R. Biochimica Biophys Acta. 2003;1631:42–50. doi: 10.1016/s1388-1981(02)00354-2. [DOI] [PubMed] [Google Scholar]
  12. Brizuela L, Rabano M, Gangoiti P, Narbona N, Macarulla JM, Trueba M, Gomez-Munoz A. Sphingosine-1-phosphate stimulates aldosterone secretion through a mechanism involving the PI3K/PKB and MEK/ERK1/2 pathways. J Lipid Res. 2007;48:2264–2274. doi: 10.1194/jlr.M700291-JLR200. [DOI] [PubMed] [Google Scholar]
  13. Brizuela L, Rabano M, Pena A, Gangoiti P, Macarulla JM, Trueba M, Gomez-Munoz A. Sphingosine-1-phosphate: a novel stimulator of aldosterone secretion. J Lipid Res. 2006;47:1238–1249. doi: 10.1194/jlr.M500510-JLR200. [DOI] [PubMed] [Google Scholar]
  14. Budnik LT, Jahner D, Mukhopadhyay AK. Inhibitory effects of TNF alpha on mouse tumor Leydig cells: possible role of ceramide in the mechanism of action. Mol Cell Endocrinol. 1999;150:39–46. doi: 10.1016/s0303-7207(99)00029-5. [DOI] [PubMed] [Google Scholar]
  15. Burnier M, Centeno G, Burki E, Brunner HR. Confocal microscopy to analyze cytosolic and nuclear calcium in cultured vascular cells. Cell Physiol. 1994;35:C1118–C1127. doi: 10.1152/ajpcell.1994.266.4.C1118. [DOI] [PubMed] [Google Scholar]
  16. Carmen GY, Victor SM. Signalling mechanisms regulating lipolysis. Cell Signal. 2006;18:401–408. doi: 10.1016/j.cellsig.2005.08.009. [DOI] [PubMed] [Google Scholar]
  17. Caron KM, Ikeda Y, Soo SC, Stocco DM, Parker KL, Clark BJ. Characterization of the promoter region of the mouse gene encoding the steroidogenic acute regulatory protein. Mol Endocrinol. 1997a;11:137–147. doi: 10.1210/mend.11.2.9880. [DOI] [PubMed] [Google Scholar]
  18. 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. 1997b;94:11540–11545. doi: 10.1073/pnas.94.21.11540. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Carr BR, Simpson ER. Lipoprotein utilization and cholesterol synthesis by the human fetal adrenal gland. Endocr Rev. 1981;2:306–326w. doi: 10.1210/edrv-2-3-306. [DOI] [PubMed] [Google Scholar]
  20. 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. doi: 10.1210/endo.140.10.7085. [DOI] [PubMed] [Google Scholar]
  21. Clark BJ, Wells J, King SR, Stocco DM. The purification, cloning, and expression of a novel leutenizing 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. [PubMed] [Google Scholar]
  22. Clifford GM, Londos C, Kraemer FB, Vernon RG, Yeaman SJ. Translocation of hormone-sensitive lipase and perilipin upon lipolytic stimulation of rat adipocytes. J Biol Chem. 2000;275:5011–5015. doi: 10.1074/jbc.275.7.5011. [DOI] [PubMed] [Google Scholar]
  23. Cook KG, Yeaman SJ, Stralfors P, Fredrikson G, Belfrage P. Direct evidence that cholesteryl ester hydrolase from adrenal cortex is the same enzyme as hormone-sensitive lipase from adipose tissue. Eur J Biochem. 1982;125:245–249. doi: 10.1111/j.1432-1033.1982.tb06675.x. [DOI] [PubMed] [Google Scholar]
  24. Coussin F, Scott RH, Nixon GF. Sphingosine-1-phosphate induces CREB activation in rat cerebral atery via a protein kinase C-mediated inhibition of voltage-gated K+ channels. Biochem Pharmacol. 2003;66:1861–1870. doi: 10.1016/s0006-2952(03)00546-x. [DOI] [PubMed] [Google Scholar]
  25. Coussin F, Scott RH, Wise A, Nixon GF. Comparison of sphingosine 1-phosphate-induced intracellular signaling pathways in vascular smooth muscles: differential role in vasoconstriction. Circ Res. 2002;91:151–157. doi: 10.1161/01.res.0000028150.51130.36. [DOI] [PubMed] [Google Scholar]
  26. Cuvillier O. Sphingosine in apoptosis signaling. Biochim Biophys Acta. 2002;1585:153–162. doi: 10.1016/s1388-1981(02)00336-0. [DOI] [PubMed] [Google Scholar]
  27. Cuvillier O, Pirianov G, Kleuser B, Vanek PG, Coso OA, Gutkind S, Spiegel S. Suppression of ceramide-mediated programmed cell death by sphingosine-1-phosphate. Nature. 1996;381:800–803. doi: 10.1038/381800a0. [DOI] [PubMed] [Google Scholar]
  28. Dammer EB, Leon A, Sewer MB. Coregulator Exchange and Sphingosine-Sensitive Cooperativity of Steroidogenic Factor-1, General Control Nonderepressed 5, p54, and p160 Coactivators Regulate Cyclic Adenosine 3',5'-Monophosphate-Dependent Cytochrome P450c17 Transcription Rate. Mol Endocrinol. 2007;21:415–438. doi: 10.1210/me.2006-0361. [DOI] [PubMed] [Google Scholar]
  29. Degerman E, Smith CJ, Tornqvist H, Vasta V, Belfrage P, Manganiello VC. Evidence that insulin and isoprenaline activate the cGMP-inhibited low-Km cAMP phosphodiesterase in rat fat cells by phosphorylation. Proceedings of the National Academy of Sciences of the USA. 1990;87:533–537. doi: 10.1073/pnas.87.2.533. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Degnan BM, Bourdelat-Parks B, Daniel A, Salata K, Francis GL. Spingomyelinase inhibits in vitro Leydig cell function. Ann. Clin. Lab. Sci. 1996;26:242–243. [PubMed] [Google Scholar]
  31. Doi J, Takemori H, Ohta M, Nonaka Y, Okamoto M. Differential regulation of 3β-hydroxysteroid dehydrogenase type II and 17α-hydroxylase/lyase P450 in human adrenocortical carcinoma cells by epidermal growth factor and basic fibroblast growth factor. J. Endocrinol. 2001;168:87–94. doi: 10.1677/joe.0.1680087. [DOI] [PubMed] [Google Scholar]
  32. Dyson MT, Jones JK, Kowaleswski 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. doi: 10.1095/biolreprod.107.064238. [DOI] [PubMed] [Google Scholar]
  33. Egan JJ, Greenberg AS, Chang MK, Wek SA, Moos MC, Jr, Londos C. Mechanism of hormone-stimulated lipolysis in adipocytes: translocation of hormone-sensitive lipase to the lipid storage droplet. Proc Natl Acad Sci USA. 1992;89:8537–8541. doi: 10.1073/pnas.89.18.8537. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Fiedler EP, Plouffe L, Jr, Hales DB, Hales KH, Khan I. Prostaglandin F(2alpha) induces a rapid decline in progesterone production and steroidogenic acute regulatory protein expression in isolated rat corpus luteum without altering messenger ribonucleic acid expression. Biol Reprod. 1999;61:643–650. doi: 10.1095/biolreprod61.3.643. [DOI] [PubMed] [Google Scholar]
  35. Garton AJ, Campbell DG, Cohen P, Yeaman SJ. Primary structure of the site on bovine hormone-sensitive lipase phosphorylated by cyclic AMP-dependent protein kinase. FEBS Lett. 1988;229:68–72. doi: 10.1016/0014-5793(88)80799-3. [DOI] [PubMed] [Google Scholar]
  36. Garton AJ, Yeaman SJ. Identification and role of the basal phosphorylation site on hormone-sensitive lipase. Eur J Biochem. 1990;191:245–250. doi: 10.1111/j.1432-1033.1990.tb19116.x. [DOI] [PubMed] [Google Scholar]
  37. Goni FM, Alonso A. Biophysics of spingolipids I. Membrane properties of sphingosine, ceramides and other simple sphingolipids. Biochimica Biophys Acta. 2006;1758:1902–1921. doi: 10.1016/j.bbamem.2006.09.011. [DOI] [PubMed] [Google Scholar]
  38. Graler MH, Grosse R, Kusch A, Kremmer E, Gudermann T, Lipp M. The sphingosine-1-phosphate receptor S1P4 regulates cell shape and motility via coupling to Gi and G12/13. J Cell Biochem. 2003;89:507–519. doi: 10.1002/jcb.10537. [DOI] [PubMed] [Google Scholar]
  39. Greenberg AS, Shen WJ, Muliro K, Patel S, Souza SC, Roth RA, Kraemer FB. Stimulation of lipolysis and hormone-sensitive lipase via the extracellular signal-regulated kinase pathway. J Biol Chem. 2001;276:45456–45461. doi: 10.1074/jbc.M104436200. [DOI] [PubMed] [Google Scholar]
  40. Gyles SL, Burns CJ, Whitehouse BJ, Sugden D, Marsh PJ, Persuad SJ, Jones PM. ERKs regulate cyclic AMP-induced steroid synthesis through transcription of the steroidogenic acute regulatory (StAR) gene. J Biol. Chem. 2001;276:34888–34895. doi: 10.1074/jbc.M102063200. [DOI] [PubMed] [Google Scholar]
  41. Hannun YA. Functions of ceramide in coordinating cellular responses to stress. Science. 1996;274:1855–1859. doi: 10.1126/science.274.5294.1855. [DOI] [PubMed] [Google Scholar]
  42. Hauet T, Liu J, Li H, Gazouli M, Culty M, Papadopoulos V. PBR, StAR, and PKA: partners in cholesterol transport in steroidogenic cells. Endocr Res. 2002;28:395–401. doi: 10.1081/erc-120016814. [DOI] [PubMed] [Google Scholar]
  43. Hla T, Lee MJ, Ancellin N, Liu CH, Thangada S, Thompson BD, Kluk MJ. Sphingosine-1-phosphate: extracellular mediator or intracellular second messenger? Biochem Pharmacol. 1999;58:201–207. doi: 10.1016/s0006-2952(99)00086-6. [DOI] [PubMed] [Google Scholar]
  44. Hla T, Lee MJ, Ancellin N, Paik JH, Kluk MJ. Lysophospholipids--receptor revelations. Science. 2001;294:1875–1878. doi: 10.1126/science.1065323. [DOI] [PubMed] [Google Scholar]
  45. Huwiler A, Kolter T, Pfeilschifter J, Sandhoff K. Pysiology and pathophysiology of sphingolipid metabolism and signaling. Biochim Biophys Acta. 2000;1485:63–99. doi: 10.1016/s1388-1981(00)00042-1. [DOI] [PubMed] [Google Scholar]
  46. Illario M, Cavallo AL, Bayer KU, Di Matola T, Fenzi G, Rossi G, Vitale M. Calcium/calmodulin-dependent protein kinase II binds to Raf-1 and modulates integrin-stimulated ERK activation. J Biol Chem. 2003;278:45101–45108. doi: 10.1074/jbc.M305355200. [DOI] [PubMed] [Google Scholar]
  47. Jongsma M, van Unen J, van Loenen PB, Michel MC, Peters SL, Alewijnse AE. Different response patterns of several ligands at the sphingosine-1-phosphate receptor subtype 3 (S1P(3)) Br J Pharmacol. 2009;156:1305–1311. doi: 10.1111/j.1476-5381.2009.00134.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Jun DJ, Lee JH, Choi BH, Koh TK, Ha DC, Jeong MW, Kim KT. Sphingosine-1-phosphate modulates both lipolysis and leptin production in differentiated rat white adipocytes. Endocrinology. 2006;147:5835–5844. doi: 10.1210/en.2006-0579. [DOI] [PubMed] [Google Scholar]
  49. Kihara A, Mitsutake S, Mizutani Y, Igarashi Y. Metabolism and biological functions of two phosphorylated sphingolipids, sphingosine 1-phosphate and ceramide 1-phosphate. Prog Lipid Res. 2007;46:126–144. doi: 10.1016/j.plipres.2007.03.001. [DOI] [PubMed] [Google Scholar]
  50. Kobayashi N, Nishi T, Hirata T, Kihara A, Sano T, Igarashi Y, Yamaguchi A. Sphingosine 1-phosphate is released from the cytosol of rat platelets in a carrier-mediated manner. J Lipid Res. 2006;47:614–621. doi: 10.1194/jlr.M500468-JLR200. [DOI] [PubMed] [Google Scholar]
  51. Kraemer FB. Adrenal cholesterol utilization. Mol Cell Endocrinol. 2007;265–266:42–45. doi: 10.1016/j.mce.2006.12.001. [DOI] [PubMed] [Google Scholar]
  52. Kraemer FB, Shen WJ. Hormone-sensitive lipase: control of intracellular tri-(di-)acylglycerol and cholesteryl ester hydrolysis. J Lipid Res. 2002;43:1585–1594. doi: 10.1194/jlr.r200009-jlr200. [DOI] [PubMed] [Google Scholar]
  53. Kraemer FB, Shen WJ, Harada K, Patel S, Osuga J, Ishibashi S, Azhar S. Hormone-sensitive lipase is required for high-density lipoprotein cholesteryl ester-supported adrenal steroidogenesis. Mol Endocrinol. 2004;18:549–557. doi: 10.1210/me.2003-0179. [DOI] [PubMed] [Google Scholar]
  54. Krintel C, Osmark P, Larsen MR, Resjö S, Logan DT, Holm C. Ser649 and Ser650 are the major determinants of protein kinase A-mediated activation of human hormone-sensitive lipase against lipid substrates. PLoS One. 2008;3:e3756. doi: 10.1371/journal.pone.0003756. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Kwun C, Patel A, Pletcher S, Lyons B, Abdelrahim M, Nicholson D, Morris E, Salata K, Francis GL. Ceramide increases steroid hormone production in MA-10 Leydig cells. Steroids. 1999;64:499–509. doi: 10.1016/s0039-128x(99)00013-6. [DOI] [PubMed] [Google Scholar]
  56. Lee MJ, Evans M, Hla T. The inducible G protein-coupled receptor edg-1 signals via the G(i)/mitogen-activated protein kinase pathway. J Biol Chem. 1996;271:11272–11279. doi: 10.1074/jbc.271.19.11272. [DOI] [PubMed] [Google Scholar]
  57. Li D, Sewer MB. RhoA and DIAPH1 Mediate Adrenocorticotropin-Stimulated Cortisol Biosynthesis by Regulating Mitochondrial Trafficking. Endocrinology. 2010;151:4313–4323. doi: 10.1210/en.2010-0044. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Li N, Wang C, Wu Y, Liu X, Cao X. Ca(2+)/calmodulin-dependent protein kinase II promotes cell cycle progression by directly activating MEK1 and subsequently modulating p27 phosphorylation. J Biol Chem. 2009;284:3021–3027. doi: 10.1074/jbc.M805483200. [DOI] [PubMed] [Google Scholar]
  59. Lin D, Sugawara T, Strauss JR, III, 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. doi: 10.1126/science.7892608. [DOI] [PubMed] [Google Scholar]
  60. Liu J, Rone MB, Papadopoulos V. Protein-protein interactions mediate mitochondrial cholesterol transport and steroid biosynthesis. J Biol Chem. 2006;281:38879–38893. doi: 10.1074/jbc.M608820200. [DOI] [PubMed] [Google Scholar]
  61. Lucki N, Sewer MB. Multiple roles for sphingolipids in steroid hormone biosynthesis. Subcell Biochem. 2008;49:387–412. doi: 10.1007/978-1-4020-8831-5_15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Lucki N, Sewer MB. The cAMP-responsive element binding protein (CREB) regulates the expression of acid ceramidase (ASAH1) in H295R human adrenocortical cells. Biochim Biophys Acta. 2009;1791:706–713. doi: 10.1016/j.bbalip.2009.03.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Luo Y, Zhou H, Liu L, Shen T, Chen W, Xu B, Han X, Zhang F, Scott RS, Alexander JS, Alam A, Huang S. The fungicide ciclopirox inhibits lymphatic endothelial cell tube formation by suppressing VEGFR-3-mediated ERK signaling pathway. Oncogene. 2011;30:2098–2107. doi: 10.1038/onc.2010.590. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. MA Connelly. SR-BI-mediated HDL cholesteryl ester delivery in the adrenal gland. Mol Cell Endocrinol. 2009;300:83–88. doi: 10.1016/j.mce.2008.09.011. [DOI] [PubMed] [Google Scholar]
  65. Malek RL, Toman RE, Edsall LC, Wong S, Chiu J, Letterle CA, Van Brocklyn JR, Milstien S, Spiegel S, Lee NH. Erg-1 belongs to the endothelial differentiation gene family of G protein-coupled sphingosine-1-phosphate receptors. J Biol Chem. 2001;276:5692–5699. doi: 10.1074/jbc.M003964200. [DOI] [PubMed] [Google Scholar]
  66. Manna PR, Chandrala SP, King SR, Jo Y, Counis R, Huhtaniemi IT, Stocco DM. Molecular mechanisms of insulin-like growth factor-I mediated regulation of the steroidogenic acute regulatory protein in mouse leydig cells. Mol Endocrinol. 2006;20:362–378. doi: 10.1210/me.2004-0526. [DOI] [PubMed] [Google Scholar]
  67. Manna PR, Dyson MT, Stocco DM. Regulation of the steroidogenic acute regulatory protein gene expression: present and future perspectives. Mol Hum Reprod. 2009a;15:321–333. doi: 10.1093/molehr/gap025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. 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. 2009b;302:1–11. doi: 10.1016/j.mce.2008.12.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Manna PR, Huhtaniemi IT, Stocco DM. Mechanisms of protein kinase C signaling in the modulation of 3',5'-cyclic adenosine monophosphate-mediated steroidogenesis in mouse gonadal cells. Endocrinology. 2009c;150:3308–3317. doi: 10.1210/en.2008-1668. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Manna PR, Jo Y, Stocco DM. Regulation of Leydig cell steroidogenesis by extracellular signal-regulated kinase 1/2: role of protein kinase A and protein kinase C signaling. J Endocrinol. 2007;193:53–63. doi: 10.1677/JOE-06-0201. [DOI] [PubMed] [Google Scholar]
  71. Manna PR, Soh JW, Stocco DM. The involvement of specific PKC isoenzymes in phorbol ester-mediated regulation of steroidogenic acute regulatory protein expression and steroid synthesis in mouse leydig cells. Endocrinology. 2011;152:313–325. doi: 10.1210/en.2010-0874. [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. 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. doi: 10.1016/j.steroids.2003.07.009. [DOI] [PubMed] [Google Scholar]
  73. Mathieson FA, Nixon GF. Sphingolipids differentially regualte mitogen-activated protein kinases and intracellular Ca2+ in vascular smooth muscle: effects on CREB activation. Br J Pharmacol. 2006;147:351–359. doi: 10.1038/sj.bjp.0706600. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. McClellan DR, Bourdelat-Parks B, Salata K, Francis GL. Sphingomyelinase affects hormone production in Jeg-3 choriocarcinoma cells. Endocrinol Metab. 1997;3:19–24. [Google Scholar]
  75. Means CK, Miyamoto S, Chun J, Brown JH. S1PR1 receptor localization confers selectivity for Gi-mediated cAMP contractile responses. J Biol Chem. 2008;283:11954–11963. doi: 10.1074/jbc.M707422200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Meroni SB, Pellizzari EH, Canepa DF, Cigorraga SB. Possible involvement of ceramide in the regulation of rat Leydig cell function. J of Steroid Biochemistry and Molecular Biology. 2000;75:307–313. doi: 10.1016/s0960-0760(00)00188-6. [DOI] [PubMed] [Google Scholar]
  77. Merrill AH, Jr, Nikolova-Karakashian M, Schmelz EM. Regulation of cytochrome P450 expression by sphingolipids. Chemistry and Physics of Lipids. 1999;102:131–130. doi: 10.1016/s0009-3084(99)00081-x. [DOI] [PubMed] [Google Scholar]
  78. Merrill AH., Jr De novo sphingolipid biosynthesis: a necessary, but dangerous, pathway. J Biol Chem. 2002;277:25843–25846. doi: 10.1074/jbc.R200009200. [DOI] [PubMed] [Google Scholar]
  79. Miller WL. Steroidogenic enzymes. Endocr Dev. 2008;13:1–18. doi: 10.1159/000134751. [DOI] [PubMed] [Google Scholar]
  80. Mitra P, Oskeritzian CA, Payne SG, Beaven MA, Milstien S, Spiegel S. Role of ABCC1 in export of sphingosine-1-phosphate from mast cells. Proc Natl Acad Sci USA. 2006;103:16394–16399. doi: 10.1073/pnas.0603734103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Mountjoy KG, Robbins LS, Mortrud MT, Cone RD. The cloning of a family of genes that encode the melanocortin receptors. Science. 1992;257:1248–1251. doi: 10.1126/science.1325670. [DOI] [PubMed] [Google Scholar]
  82. Nguyen A, Chen P, Cai H. Role of CaMKII in hydrogen peroxide activation of ERK1/2, p38 MAPK, HSP27 and actin reorganization in endothelial cells. FEBS Lett. 2004;572:307–313. doi: 10.1016/j.febslet.2004.06.061. [DOI] [PubMed] [Google Scholar]
  83. Noh SJ, Kim MJ, Shim S, Han JK. Different signaling pathway between sphingosine-1-phosphate and lysophosphatidic acid in Xenopus oocytes: functional coupling of the sphingosine-1-phosphate receptor to PLC-xbeta in Xenopus oocytes. J Cell Physiol. 1998;176:412–423. doi: 10.1002/(SICI)1097-4652(199808)176:2<412::AID-JCP20>3.0.CO;2-3. [DOI] [PubMed] [Google Scholar]
  84. Ogretmen B. Sphingolipids in cancer: Regulation of pathogenesis and therapy. FEBS Lett. 2006;580:5467–5476. doi: 10.1016/j.febslet.2006.08.052. [DOI] [PubMed] [Google Scholar]
  85. Okajima F, Tomura H, Sho K, Nochi H, Tamoto K, Kondo Y. Involvement of pertussis toxin-sensitive GTP-binding proteins in sphingosine 1-phosphate-induced activation of phospholipase C-Ca2+ system in HL60 leukemia cells. FEBS Lett. 1996;379:260–264. doi: 10.1016/0014-5793(95)01526-4. [DOI] [PubMed] [Google Scholar]
  86. Olivera A, Kohama T, Edsall L, Nava V, Cuvillier O, Poulton S, Spiegel S. Sphingosine kinase expression increases intracellular sphingosine-1-phosphate and promotes cell growth and survival. J Cell Biol. 1999;147:545–558. doi: 10.1083/jcb.147.3.545. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Olivera A, Spiegel S. Sphingosine-1-phosphate as a second messenger in cell proliferation induced by PDGF and FCS mitogens. Nature. 1993;365:557–560. doi: 10.1038/365557a0. [DOI] [PubMed] [Google Scholar]
  88. Olivera A, Spiegel S. Sphingosine kinase: a mediator of vital cellular functions. Prostaglandins Other Lipid Mediators. 2001;64:123–134. doi: 10.1016/s0090-6980(01)00108-3. [DOI] [PubMed] [Google Scholar]
  89. Orlicky J, Sulova Z, Dovinova I, Fiala R, Zahradnikova A, Jr, Breier A. Functional Fluo-3/AM assay on p-glycoprotein transport activity in L1210/VCR cells by confocal microscopy. Gen Physiol Biophys. 2004;23:357–366. [PubMed] [Google Scholar]
  90. 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. doi: 10.1081/erc-200044040. [DOI] [PubMed] [Google Scholar]
  91. Ozbay T, Rowan A, Leon A, Patel P, Sewer MB. Cyclic adenosine 5'-monophosphate-dependent sphingosine-1-phosphate biosynthesis induces human CYP17 gene transcription by activating cleavage of sterol regulatory element binding protein 1. Endocrinology. 2006;147:1427–1437. doi: 10.1210/en.2005-1091. [DOI] [PubMed] [Google Scholar]
  92. Pandey AK, Li W, Yin X, Stocco DM, Grammas P, Wang X. Blocking L-type calcium channels reduced the threshold of cAMP-induced steroidogenic acute regulatory gene expression in MA-10 Leydig cells. J Endocrinol. 2010;204:67–74. doi: 10.1677/JOE-09-0206. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Papadopoulos V. Peripheral-type benzodiazepine/diazepam binding inhibitor receptor: biological role in steroidgenic cell function. Endocr Rev. 1993;14:222–240. doi: 10.1210/edrv-14-2-222. [DOI] [PubMed] [Google Scholar]
  94. Payne AH, Hales DB. Overview of steroidogenic enzymes in the pathway from cholesterol to active steroid hormones. Endocr Rev. 2004;25:947–970. doi: 10.1210/er.2003-0030. [DOI] [PubMed] [Google Scholar]
  95. Pilon A, Martin G, Bultel-Brienne S, Junquero D, Delhon A, Fruchart JC, Staels B, Clavey V. Regulation of the scavenger receptor BI and the LDL receptor by activators of aldosterone production, angiotensin II and PMA, in the human NCI-H295R adrenocortical cell line. Biochimica Biophys Acta. 2003;1631:218–228. doi: 10.1016/s1388-1981(03)00020-9. [DOI] [PubMed] [Google Scholar]
  96. Poderoso C, Maloberti P, Duarte A, Neuman I, Paz C, Maciel FC, Podesta EJ. Hormonal activation of a kinase cascade localized at the mitochondria is required for StAR protein activity. Mol Cell Endocrinol. 2009;300:37–42. doi: 10.1016/j.mce.2008.10.009. [DOI] [PubMed] [Google Scholar]
  97. Porn MI, Tenhunen J, Slotte JP. Increased steroid hormone secretion in mouse Leydig tumor cells after induction of cholesterol translocation by sphingomyelin degradation. Biochim Biophys Acta. 1991;1093:7–12. doi: 10.1016/0167-4889(91)90131-g. [DOI] [PubMed] [Google Scholar]
  98. Rabano M, Pena A, Brizuela L, Marino A, Macarulla JM, Trueba M, Gomez-Munoz A. Sphingosine-1-phosphate stimulates cortisol secretion. FEBS Lett. 2003;535:101–105. doi: 10.1016/s0014-5793(02)03882-6. [DOI] [PubMed] [Google Scholar]
  99. Rainey WE, Bird IM, Mason JI. The NCI-H295 cell line: a pluripotent model for human adrenocortical studies. Mol Cell Endocrinol. 1994;99:R17–R20. doi: 10.1016/0303-7207(94)90277-1. [DOI] [PubMed] [Google Scholar]
  100. Rainey WE, Saner K, Schimmer BP. Adrenocortical cell lines. Mol Cell Endocrinol. 2004;228:23–38. doi: 10.1016/j.mce.2003.12.020. [DOI] [PubMed] [Google Scholar]
  101. Rakhit S, Conway AM, Tate R, Bower T, Pyne NJ, Pyne S. Sphingosine 1-phosphate stimulation of the p42/p44 mitogen-activated protein kinase pathway in airway smooth muscle. Role of endothelial differentiation gene 1, c-Src tyrosine kinase and phosphoinositide 3-kinase. Biochem J. 1999;338:643–649. [PMC free article] [PubMed] [Google Scholar]
  102. Sanchez T, Hla T. Structural and functional characteristics of S1P receptors. J Cell Biochem. 2004;92:913–922. doi: 10.1002/jcb.20127. [DOI] [PubMed] [Google Scholar]
  103. Sandhoff TW, Hales DB, Hales KH, McLean MP. Transcriptional regulation of steroidogenic acute regulatory protein gene by steroidogenic factor 1. Endocrinology. 1998;139:4820–4831. doi: 10.1210/endo.139.12.6345. [DOI] [PubMed] [Google Scholar]
  104. Santana P, LLanes L, Hernandez I, Gonzalez-Robayna I, Tabraue C, Gonzalez-Reyes J, Quintana J, Estevez F, Ruiz de Galarretta CM, Fanjul LF. Interleukin-1beta stimulates sphingomyelin hydrolysis in cultured granulosa cells: evidence for a regulatory role of ceramide on progesterone and prostaglandin biosynthesis. Endocrinology. 1996;137:2480–2489. doi: 10.1210/endo.137.6.8641202. [DOI] [PubMed] [Google Scholar]
  105. Schinner S, Willenberg HS, Krause D, Schott M, Lamounier-Zepter V, Krug AW, Ehrhart-Bornstein M, Bornstein SR, Scherbaum WA. Adipocyte-derived products induce the transcription of the StAR promoter and stimulate aldosterone and cortisol secretion from adrenocortical cells through the Wnt-signaling pathway. Int J Obes (Lond) 2007;31:864–870. doi: 10.1038/sj.ijo.0803508. [DOI] [PubMed] [Google Scholar]
  106. Schmitt JM, Wayman GA, Nozaki N, Soderling TR. Calcium activation of ERK mediated by calmodulin kinase I. J Biol Chem. 2004;279:24064–24072. doi: 10.1074/jbc.M401501200. [DOI] [PubMed] [Google Scholar]
  107. Sensken SC, Stäubert C, Keul P, Levkau B, Schöneberg T, Gräler MH. Selective activation of G alpha i mediated signaling of S1P3 by FTY720-phosphate. Cell Signal. 2008;20:1125–1133. doi: 10.1016/j.cellsig.2008.01.019. [DOI] [PubMed] [Google Scholar]
  108. Sewer MB, Dammer EB, Jagarlapudi S. Transcriptional regulation of adrenocortical steroidogenic gene expression. Drug Metab Rev. 2007;39:371–388. doi: 10.1080/03602530701498828. [DOI] [PubMed] [Google Scholar]
  109. Sewer MB, Li D. Regulation of steroid hormone biosynthesis by the cytoskeleton. Lipids. 2008;43:1109–1115. doi: 10.1007/s11745-008-3221-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Sewer MB, Waterman MR. Insights into the transcriptional regulation of steroidogenic enzymes and StAR. Rev. in Endocrine and Metabolic Disorders. 2001;2:269–274. doi: 10.1023/a:1011516532335. [DOI] [PubMed] [Google Scholar]
  111. Sewer MB, Waterman MR. ACTH modulation of transcription factors responsible for steroid hydroxylase gene expression in the adrenal cortex. Microsc Res Tech. 2003;61:300–307. doi: 10.1002/jemt.10339. [DOI] [PubMed] [Google Scholar]
  112. Shah BH, Baukal AJ, Shah FB, Catt KJ. Mechanisms of extracellularly regulated kinases 1/2 activation in adrenal glomerulosa cells by lysophosphatidic acid and epidermal growth factor. Mol Endocrinol. 2005;19:2535–2548. doi: 10.1210/me.2005-0082. [DOI] [PubMed] [Google Scholar]
  113. Shen WJ, Patel K, natu V, Hong R, Wang J, Azhar S, Kraemer FB. Interaction of hormone-sensitive lipase with steroidogenic acute regulatory protein: facilitation of cholesterol transfer in adrenal. J Biol Chem. 2003;278:43870–43876. doi: 10.1074/jbc.M303934200. [DOI] [PubMed] [Google Scholar]
  114. Spiegel S, Milstien S. Functions of a new family of sphingosine-1-phosphate receptors. Biochim Biophys Acta. 2000;1484:107–116. doi: 10.1016/s1388-1981(00)00010-x. [DOI] [PubMed] [Google Scholar]
  115. Spiegel S, Milstien S. Sphingosine 1-phosphate, a key cell signaling molecule. J Biol Chem. 2002;277:25851–25854. doi: 10.1074/jbc.R200007200. [DOI] [PubMed] [Google Scholar]
  116. Spiegel S, Milstien S. Sphingosine-1-phosphate: an enigmatic signalling lipid. Nat Rev Mol Cell Biol. 2003;4:397–407. doi: 10.1038/nrm1103. [DOI] [PubMed] [Google Scholar]
  117. Staels B, Hum DW, Miller WL. Regulation of steroidogenesis in NCI-H295 cells: a cellular model of the human fetal adrenal. Mol Endocrinol. 1993;7:423–433. doi: 10.1210/mend.7.3.8387159. [DOI] [PubMed] [Google Scholar]
  118. Stocco D, Wang X, Jo Y, Manna PR. Multiple signaling pathways regulating steroidogenesis and steroidogenic acute regulatory protein expression: more complicated than we thought. Mol Endocrinol. 2005;19:2647–2659. doi: 10.1210/me.2004-0532. [DOI] [PubMed] [Google Scholar]
  119. Stocco DM. The role of the StAR proein in steroidogenesis: challenges for the future. J Endocrinol. 2000;164:247–253. doi: 10.1677/joe.0.1640247. [DOI] [PubMed] [Google Scholar]
  120. Stocco DM, Clark BJ. Regulation of the acute production of steroids in steroidogenic cells. Endocr Rev. 1996;17:221–244. doi: 10.1210/edrv-17-3-221. [DOI] [PubMed] [Google Scholar]
  121. Stralfors P, Belfrage P. Phosphorylation of hormone-sensitive lipase by cyclic AMP-dependent protein kinase. J Biol Chem. 1983;258:15146–15152. [PubMed] [Google Scholar]
  122. Stralfors P, Bjorgell P, Belfrage P. Hormonal regulation of hormone-sensitive lipase in intact adipocytes: identification of phosphorylated sites and effects on the phosphorylation by lipolytic hormones and insulin. Proc Natl Acad Sci USA. 1984;81:3317–3321. doi: 10.1073/pnas.81.11.3317. [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. Sugawara T, Holt JA, Kiriakidou M, Strauss JR., III Steroidogenic factor 1-dependent promoter activity of the human steroidogenic acute regulatory protein (StAR) gene. Biochemistry. 1996;35:9052–9059. doi: 10.1021/bi960057r. [DOI] [PubMed] [Google Scholar]
  124. Suhaiman L, De Blas GA, Obeid LM, Darszon A, Mayorga LS, Belmonte SA. Sphingosine 1-phosphate and sphingosine kinase are involved in a novel signaling pathway leading to acrosomal exocytosis. J Biol Chem. 2010;285:16302–16314. doi: 10.1074/jbc.M109.072439. [DOI] [PMC free article] [PubMed] [Google Scholar]
  125. Taha TA, Argraves KM, Obeid LM. Sphingosine-1-phosphate receptors: receptor specificity versus functional redundancy. Biochim Biophys Acta. 2004;1682:48–55. doi: 10.1016/j.bbalip.2004.01.006. [DOI] [PubMed] [Google Scholar]
  126. Takabe K, Kim RH, Allegood JC, Mitra P, Ramachandran S, Nagahashi M, Harikumar KB, Hait NC, Milstien S, Spiegel S. Estradiol induces export of sphingosine 1-phosphate from breast cancer cells via ABCC1 and ABCG2. J Biol Chem. 2010;285:10477–10486. doi: 10.1074/jbc.M109.064162. [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Taussig R, Iniguez-Lluhi JA, Gilman AG. Inhibition of adenylyl cyclase by Gi alpha. Science. 1993;261:218–221. doi: 10.1126/science.8327893. [DOI] [PubMed] [Google Scholar]
  128. Thomson M. Molecular and cellular mechanisms used in the acute phase of stimulated steroidogenesis. Hormone Metabolism Research. 1997;30:16–28. doi: 10.1055/s-2007-978825. [DOI] [PubMed] [Google Scholar]
  129. Tsai J-H, Chen H-W, Chen Y-W, Liu J-Y, Lii C-K. The protection of hepatocyte cells from the effects of oxidative stress by treatment with vitamin E in conjunction with DTT. J Biomed Biotechnol. 2010;2010 doi: 10.1155/2010/486267. 486267. [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Urs AN, Dammer E, Sewer MB. Sphingosine Regulates the Transcription of CYP17 by Binding to Steroidogenic Factor-1. Endocrinology. 2006;147:5249–5258. doi: 10.1210/en.2006-0355. [DOI] [PubMed] [Google Scholar]
  131. Van Brocklyn JR, Graler MH, Bernhardt G, Hobson JP, Lipp M, Spiegel S. Sphingosine-1-phosphate is a ligand for the G protein-coupled receptor EDG-6. Blood. 2000;95:2624–2629. [PubMed] [Google Scholar]
  132. van Brocklyn JR, Lee MJ, Menzeleev R, Olivera A, Edsall L, Cuvillier O, Thomas DM, Coopman PJ, Thangada S, Liu CH, Hla T, Spiegel S. Dual actions of sphingosine-1-phosphate: extracellular through the Gi-coupled receptor Edg-1 and intracellular to regulate proliferation and survival. J Cell Biol. 1998;142:229–240. doi: 10.1083/jcb.142.1.229. [DOI] [PMC free article] [PubMed] [Google Scholar]
  133. Wang F, Van Brocklyn JR, Edsall L, Nava VE, Spiegel S. Sphingosine-1-phosphate inhibits motility of human breast cancer cells independently of cell surface receptors. Cancer Research. 1999;59:6185–6191. [PubMed] [Google Scholar]
  134. Watanabe M, Noda M, Nakajin S. Effect of epidermal growth factor and prostaglandin on the expression of aromatase (CYP19) in human adrenocortical carcinoma cell line NCI-H295R cells. J Endocrinol. 2006;188:59–68. doi: 10.1677/joe.1.06214. [DOI] [PubMed] [Google Scholar]
  135. Windh RT, Lee MJ, Hla T, An S, Barr MJ, Manning DR. Differential coupling of the sphingosine-1-phosphate receptors EDG-1, EDG-3 and H218/EDG-5 to the G(i), G(q), and G(12) families of heterotrimeric G proteins. J Biol Chem. 1999;274:27351–27358. doi: 10.1074/jbc.274.39.27351. [DOI] [PubMed] [Google Scholar]
  136. Wu GY, Dieisseroth K, Tsien RW. Activity-dependent CREB phosphorylation: convergence of a fast, sensistive calmodulin kinase pathway and a slow, sensitive mitogen-activated protein kianse pathway. Proc Natl Acad Sci USA. 2001;98:2808–2813. doi: 10.1073/pnas.051634198. [DOI] [PMC free article] [PubMed] [Google Scholar]
  137. Wu J, Spiegel S, Sturgill TW. Sphingosine-1-phosphate rapidly activates the mitogen-activated protein kinase pathway by a G protein-dependent mechanism. J Biol Chem. 1995;270:11484–11488. doi: 10.1074/jbc.270.19.11484. [DOI] [PubMed] [Google Scholar]
  138. Yamamura S, Sadahira Y, Ruan F, Hakomori S, Igarashi Y. Sphingosine-1-phosphate inhibits actin nucleation and pseudopodium formation to control cell motility of mouse melanoma cells. FEBS Lett. 1996;382:193–197. doi: 10.1016/0014-5793(96)00175-5. [DOI] [PubMed] [Google Scholar]
  139. Yamazaki Y, Kon J, Sato K, Tomura H, Sato M, Yoneya T, Okazaki H, Okajima F, Ohta H. Edg-6 as a putative sphingosine-1-phosphate receptor coupling to Ca(2+) signaling pathway. Biochem Biophys Res Commun. 2000;268:583–589. doi: 10.1006/bbrc.2000.2162. [DOI] [PubMed] [Google Scholar]
  140. Zeidan YH, Hannun YA. Translational aspects of sphingolipid metabolism. Trends Mol Med. 2007;13:327–336. doi: 10.1016/j.molmed.2007.06.002. [DOI] [PubMed] [Google Scholar]
  141. Zondag GC, Postma FR, Etten IV, Verlaan I, Moolenaar WH. Sphingosine 1-phosphate signalling through the G-protein-coupled receptor Edg-1. Biochem J. 1998;330:605–609. doi: 10.1042/bj3300605. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES