Expression and functional characterization of SCaMPER: a sphingolipid-modulated calcium channel of cardiomyocytes

Amy L. Cavalli1, Nicole W. O'Brien1, Steven B. Barlow1, Romeo Betto3, Christopher C. Glembotski1, Philip T. Palade2, and Roger A. Sabbadini1

1 SDSU Heart Institute and Department of Biology, San Diego State University, San Diego, California 92182-4614; 2 Department of Physiology and Biophysics, University of Texas Medical Branch, Galveston, Texas 77555-0641; and 3 Consiglio Nazionale delle Ricerca Institute of Neuroscience, Section of Muscle Biology and Physiopathology, Department of Biomedical Sciences, University of Padua, 35121 Padua, Italy


    ABSTRACT
TOP
ABSTRACT
INTRODUCTION
MATERIALS AND METHODS
RESULTS
DISCUSSION
REFERENCES

Calcium channels are important in a variety of cellular events including muscle contraction, signaling, proliferation, and apoptosis. Sphingolipids have been recognized as mediators of intracellular calcium release through their actions on a calcium channel, sphingolipid calcium release-mediating protein of the endoplasmic reticulum (SCaMPER). The current study investigates the expression and function of SCaMPER in cardiomyocytes. Northern analyses and RT-PCR cloning and sequencing revealed SCaMPER expression in both human and rat cardiac tissue. Immunofluorescence and Western blot analyses demonstrated that SCaMPER is abundant in cardiac tissue and is localized to the sarcotubular junction. This was confirmed by the colocalization of SCaMPER with dihydropyridine and ryanodine receptors by confocal microscopy. Purified T tubules were shown to contain SCaMPER and immunoelectron micrographs suggested that SCaMPER is located to the junctional T tubules, but a junctional SR localization cannot be ruled out. The sphingolipid ligand for SCaMPER, sphingosylphosphorylcholine (SPC), initiated calcium release from the cardiomyocyte SR. Importantly, antisense knockdown of SCaMPER mRNA produced a substantial reduction of sphingolipid-induced calcium release, suggesting that SCaMPER is a potentially important calcium channel of cardiomyocytes.

sphingolipid calcium release-mediating protein of the endoplasmic reticulum; calcium channel; sphingosylphosphorylcholine; sphingolipids and cardiomyocytes


    INTRODUCTION
TOP
ABSTRACT
INTRODUCTION
MATERIALS AND METHODS
RESULTS
DISCUSSION
REFERENCES

RECENTLY, sphingolipids have emerged as important signaling molecules with diverse effects on the cardiovascular system (8). These effects include promotion of angiogenesis (7, 11), atherogenesis (27, 28), coronary vasoconstriction, and arrhythmias (9, 13, 24). Additionally, sphingolipids have been shown to mediate a variety of cellular events such as proliferation, apoptosis, and calcium signaling (10). The diverse effects of sphingolipid mediators are in part explained by the finding that these signaling molecules have direct intracellular actions as well as acting on surface membrane receptors (4).

Sphingolipid mediators such as sphingosine 1-phosphate (S1P) and sphingosylphosphorylcholine (SPC) act as extracellular signaling molecules (3). For example, many of the actions of extracellular S1P have been attributed to activation of a family of G protein-coupled receptors (GPCRs) belonging to the Edg family of receptors (12). We recently showed that Edg-1 is expressed on the surface of cardiomyocytes and is responsible for S1P's ability to deregulate calcium levels (17) and promote apoptosis in cardiomyocytes (31). Many important physiological actions of S1P and SPC are independent of Edg receptors (26). However, the search for Edg-independent sites of sphingolipid action has remained problematic. The present work was initiated with the intent of identifying other potential protein targets responsible for sphingolipid effects in heart cells.

Recently, S1P and SPC have been recognized as mediators of intracellular calcium release through their actions on a calcium channel, sphingolipid calcium release-mediating protein of the endoplasmic reticulum (SCaMPER) (5). Proof that SCaMPER is a calcium channel came from the finding that a SCaMPER cDNA obtained from a Madin-Darby canine kidney (MDCK) expression library was expressed in oocytes and was subsequently found to mediate sphingolipid-induced intracellular calcium release (14). We previously showed (1) SCaMPER expression in heart, skeletal muscle, and brain tissue and suggested that it may play a calcium regulatory role in the heart; however, we had not determined which cell type in the heart expressed SCaMPER and had not elucidated its functional role in heart cells.

Consequently, this study was initiated to determine the expression, localization, and function of SCaMPER in sphingolipid signaling and calcium regulation of heart cells. We report here that SCaMPER is expressed in cardiomyocytes. Furthermore, immunological techniques localize SCaMPER to the transverse tubules, an important site of calcium regulation. Using antisense expression constructs, we demonstrate SCaMPER's role in calcium signaling in cardiomyocytes.


    MATERIALS AND METHODS
TOP
ABSTRACT
INTRODUCTION
MATERIALS AND METHODS
RESULTS
DISCUSSION
REFERENCES

RT-PCR, cloning, and DNA sequencing. Kidney, brain, skeletal, and heart tissue isolated from adult rats and adult human heart tissue were used. The tissue was either used fresh or flash frozen, and RNA was isolated with the Qiagen Rneasy Mini Kit (Valencia, CA). Amplification of the sequence was with a Qiagen One-Step RT-PCR kit using the standard protocol. Primer sets included sense primer 5'-CCAGGATTCATCATATGTTAAAAG-3' and antisense primer of 5'-ATCAGTGGGTGCATCAGTAGC-3' for open reading frame (ORF) (designed from GenBank accession no. U33628). Amplification utilized the kit's recommended cycling with an optimization of 40 PCR cycles, which featured 45-s denaturation at 95°C, 45 s of annealing at 50°C, and 1 min of polymerization at 72°C. The SCaMPER PCR products were subcloned into pCR3.1 TOPO (Invitrogen, Carlsbad, CA) with standard techniques. Sequencing with a T7 primer was performed at the San Diego State University DNA Core Facility (San Diego, CA) and was used to diagnose plasmids as sense or antisense orientations. Plasmids were amplified and purified for subsequent transfection studies with QIAfilter plasmid MAXI and MEGA kits (Qiagen).

SCaMPER antibody development. A putative antigenic site for SCaMPER was determined by PC GENE and MCVECTOR software on the basis of its hydrophilicity, surface probability, backbone flexibility, and predicted tertiary structure of the protein (GenBank accession no. U33628). The identified sequence was APDLKIRDPKIEKLYC, which is located near the NH2 terminus of the putative cytosolic domain. A search of the GenBank protein data bank via BLAST revealed that this sequence is unique for the SCaMPER protein. The peptide corresponding to the sequence was synthesized by Sigma-Genosys (Woodlands, TX). A rabbit polyclonal anti-SCaMPER antibody was generated with standard techniques and titered via ELISA. The serum from the final bleed was purified by protein A column chromatography (Bio-Rad, Hercules, CA).

Immunological methods. Microsomal fractions of adult and neonatal rat ventricles were isolated as previously described (2). Sarcoplasmic reticulum (SR) fractions were isolated from total microsomal fractions by iterative loading with calcium oxalate and multistep sucrose density gradient centrifugation as previously described (20). Designated amounts of sample for Western blot analysis were incubated with Laemmli buffer (Bio-Rad), run on SDS-PAGE (18%), and transferred to nitrocellulose. Primary antibody dilutions and incubation times of 1:10,000 SCaMPER IgG for 45 min, 1:1,000 sarco(endo)plasmic reticulum Ca2+-ATPase (SERCA)2 ATPase and phospholamban IgG2a (ABR, Golden, CO) and 1:1,000 Na+/Ca2+ exchanger (cardiac) IgG1 (a kind gift from Kenneth Philipson, UCLA, Los Angeles, CA) for 1 h were used. Secondary antibody dilutions of 1:5,000 anti-rabbit or anti-mouse horseradish peroxidase (HRP)-conjugated (Invitrogen) antibodies for 1 h were used. The blots were developed with the ECL technique according to the manufacturer's protocol (Amersham Pharmacia Biotech, Piscataway, NJ). For the peptide blocking experiments, blotted protein was preincubated for 1 h with 10 µg/ml of peptide (used for SCaMPER antibody immunizations) before addition of the anti-SCaMPER antibody. For the deglycosylation experiments, an enzymatic deglycosylation kit (Glyko, Novato, CA) was used with the kit protocol.

Immunofluorescence microscopy was performed on 6-µm-thick frozen cryostat sections of adult rat ventricle tissue as previously described (31) and on isolated adult and neonatal cardiomyocytes (17). Antibody dilutions and incubation times were 1:150 anti-SCaMPER IgG for 1 h and 1:500 anti-rabbit-FITC or 1:500 anti-mouse-rhodamine red (Jackson, West Grove, PA) for 1 h. Colocalization experiments used an anti-dihydropyridine receptor (DHPR) antibody at 1:500 dilution (ABR), an anti-ryanodine receptor (RyR) antibody at 1:500 dilution (ABR), and an anti-SERCA2 antibody at 1:500 dilution (ABR) along with anti-mouse-rhodamine red-X antibody (Jackson) at a dilution of 1:500. An antinuclear pore O-linked glycoprotein antibody (ABR) was used at a dilution of 1:100. Confocal images were taken on a Leica TCS SP2 confocal inverted light microscope. In other experiments, immunofluorescence was performed with a Nikon Labophot-2 microscope.

Immunoelectron microscopy of cardiac tissue was performed essentially as previously described by Ralston and Ploug (19) with slight modification. Briefly, a fresh rat heart was excised, cannulated, and perfused with 60 ml of ice-cold fixative (4% paraformaldehyde in 0.1 M phosphate buffer). The left ventricle was then cut into thin strips and allowed to incubate further in the fixative for 5 h on ice. Small bundles of cardiac fibers were teased away in 50 mM glycine and incubated in blocking solution (50 mM glycine, 0.25% BSA, 0.1% saponin in 1× PBS) for 30 min. The fibers were then incubated with anti-SCaMPER antibody or preimmune serum at a 1:100 dilution in blocking solution overnight at room temperature. Fibers were washed three times for 30 min in blocking solution and incubated for 4 h with a goat anti-rabbit 10-nm gold-conjugated secondary antibody (BB International). The fibers were then washed twice in blocking solution and once in 1× PBS. Glutaraldehyde fixation, silver enhancement, and further processing for electron microscopy were performed as previously described (19) with a Philips 410A electron microscope. Imaging was performed with the PGT IMIX Image Analysis System.

Northern analysis. Adult rat heart ventricular RNA was isolated with a Qiagen Rneasy MAXI kit using the kit's standard protocol for isolation from animal tissue. mRNA was isolated with the PolyATtract mRNA isolation system (Promega). Ten micrograms of mRNA were fractionated on a 1% agarose-18% formaldehyde gel and blotted onto Zeta-Probe GT genomic membranes (Bio-Rad). A 350-bp probe consisting of the SCaMPER ORF was radioactively labeled with Random Primer Labeling Reaction (Invitrogen). Hybridization was performed in 20 ng/ml 32P-labeled probe (specific activity of 1-109 cpm/µg) in hybridization solution (5× Denhardt's, 6× SSC, 0.5% SDS) with 100 µg/ml herring sperm (Sigma, St. Louis, MO) at 55°C overnight. The blot was washed in 1× SSC-0.5% SDS twice for 20 min and with 0.1× SSC-0.5% SDS twice for 20 min at 55°C. Autoradiography was performed for 24 h on Kodak Biomax film at -70°C.

Expression methods. The SCaMPER ORF was cloned into pcDNA3.1-HIS/V5 TOPO TA cloning vector (Invitrogen) and used for expression in an in vitro rabbit reticulocyte lysate system for molecular weight determination. The TNT Quick Coupled Transcription/Translation System (Promega) was used for expression with the Transcend nonradioactive translation detection system (Promega) utilizing biotinylated lysine incorporation. Protein expression was performed and subjected to Western blotting and detection with the kit protocol. Analysis of cotranslational and initial posttranslational processing events including core glycosylation was with addition of canine pancreatic microsomal membrane (Promega) to the transcription/translation system.

Bacterial expression was from a pRHA-driven expression vector induced with 10 mM rhabanose for 2 h in exponentially growing Escherichia coli MG1655 (a kind gift from Mpex BioScience, San Diego, CA). Bacteria were collected at an OD600 of 1.0, lysed in Laemmli buffer, separated by SDS-PAGE (18%), transferred to nitrocellulose, and probed with the SCaMPER antibody as described above.

For the antisense stress response assay, neonatal cardiomyocytes were transfected (as described below) with previously published plasmids (25). Briefly, either GRP78-ERSE-Luc encoding the active ERSE from the human GRP78 cDNA driving SV40 luciferase or a pGL2 control vector (Promega) was cotransfected either with the antisense-SCaMPER construct (used in calcium analysis) or with empty vector pCDNA3.1. The cardiomyocytes were also cotransfected with a SV40-beta -gal-pCH110-C1 (Amersham Pharmacia Biotech), which encodes for beta -galactosidase, whose expression corrects for transfection efficiency. Cells were cultured under conditions identical to those described below for calcium analyses, and at indicated times cardiomyocytes were lysed and beta -gal and luciferase expression were determined as previously described (25). As a positive control for a stress response, cardiomyocytes were treated with 2.5 mM tunicamycin for 16 h.

Cell culture methods. Cardiomyocytes were isolated from neonatal rat hearts or from adult Sprague-Dawley rat hearts with previously published techniques (6, 23) with modifications including the use of pyrogen-free collagenase (Liberase blendzyme; Roche, Indianapolis, IN) for the digestion of the adult rat hearts. The cardiomyocytes were plated on 5 µg/ml or 25 µg/ml fibronectin-coated cell culture plates or glass coverslips, respectively. The cells were cultured in Dulbecco's modified Eagle's medium (DMEM)-F-12 with 10% fetal bovine serum for 24-48 h for neonatal cells and 1 h for adult cells. All procedures involving experimental animals were performed in accordance with institutional and NIH guidelines for the ethical treatment of animals.

Cell transfections. For neonatal cardiomyocytes, pCR3.1 containing a reverse SCaMPER insert (1,200 bases of the ORF and partial 3' end) was used for antisense expression and a nonsense (not recognized against any specific databased sequence) insert of 1,200 bases was used for control (nonsense) expression. Cardiomyocyte transfections were performed by electroporation, with 3 million neonatal cardiomyocytes/cuvette, 5 µg of pGreen Lantern (Invitrogen), and 20 µg of pcDNA3.1 empty vector, nonsense, or antisense SCaMPER. Electroporations were at 0.55 mV as previously described (31).

Calcium assays. Calcium measurements were measured as described previously (31) with indo 1-AM (Sigma). Results are presented as the ratio of indo 1 fluorescence emission of 405 vs. 485 nm rather than as absolute calcium concentrations. Cardiomyocytes were washed once with Tyrode solution (in mM: 140 NaCl, 5.4 KCl, 0.5 MgCl2, 1 CaCl2, 10 HEPES, 0.25 NaH2PO4; pH 7.3), loaded with 3 mM indo 1-AM for 25 min, and washed again. The coverslips were mounted into a chamber with 1 ml of Tyrode solution and placed in position for calcium analysis with the PTI system. Positively transfected cardiomyocytes were identified with green fluorescent protein (GFP; Photon Technologies) fluorescence at 480 nm and selected for calcium analysis. Selected cardiomyocytes were electrically paced for 20 s at 0.3 Hz with 10-ms stimuli with no delay at a monophasic output. Pacing was turned off, and at the designated time point, 50 µM SPC (final concentration) was carefully added. SPC in CHCl3-MeOH (2:1) was dried down with argon and resuspended in 100 µl of BSA (1 mg/ml) and Tyrode solution to a 100 µM SPC working solution. At the designated time points, 10 mM caffeine (final concentration) was added to elicit the release of calcium from the SR membranes. In selected experiments designed to test for the effects of antisense treatment on phasic calcium release, we compared the calcium transients of electrically paced cells transfected either with empty vector or with the antisense expression construct. The calcium transients were then overlaid to determine whether there was an effect on normal excitation-contraction (EC) coupling.

Statistical analyses were performed with Student's t-test. Means ± SE are shown in all results.


    RESULTS
TOP
ABSTRACT
INTRODUCTION
MATERIALS AND METHODS
RESULTS
DISCUSSION
REFERENCES

SCaMPER expression in cardiac tissue. To investigate SCaMPER expression, total RNA was isolated from rat heart and subjected to RT-PCR. The RT-PCR procedure was carried out with primers against the MDCK clone of SCaMPER (see MATERIALS AND METHODS). For both human and rat cardiac tissue, RT-PCR yielded a product of 1.2 kb (accession nos. AY163814 for human and AY163813 for rat) . Additional RT-PCR analyses demonstrated the expression of SCaMPER mRNA in various other tissues including rat kidney, brain, and skeletal muscle (Fig. 1A).


View larger version (56K):
[in this window]
[in a new window]
 
Fig. 1.   Sphingolipid calcium release-mediating protein of the endoplasmic reticulum (SCaMPER) mRNA is expressed by heart and other tissues. A: RT-PCR analysis of total RNA isolated from rat kidney, brain, skeletal muscle, and heart tissue as well as human heart tissue demonstrates the expression of SCaMPER mRNA. B: Northern analysis of 10 µg of total adult heart rat RNA and 10 µg of poly-A mRNA. C: CLUSTAL W multiple sequence alignment of the open reading frame of the originally published canine, the revised canine, the rat, and the human SCaMPER nucleic acid sequences; * denotes single, fully conserved residues and - indicates no consensus.

Northern blot analyses of isolated cardiac mRNA resulted in a single SCaMPER transcript of ~2.1 kb (Fig. 1B). A 2.1-kb transcript is consistent with the size of our 1.2-kb partial cDNA [0.33 kb ORF + 0.9 kb 3' untranslated region (UTR)]. We conclude that the SCaMPER transcript contains the 333-bp ORF for the full-length SCaMPER protein plus 1.8 kb for both UTRs.

Sequence analysis of the SCaMPER DNA revealed a single ORF encoding a predicted protein of 110 amino acids. Primary nucleotide comparisons of the ORFs and partial 3' UTR of the rat and human to the original canine (MDCK cells) clone revealed 96% and 92% identities, respectively (Fig. 1C) (nucleotide comparison of 3' UTR not shown). A previously published cDNA (14) was kindly supplied by L. A. Kindman (Duke Univ. Medical Center, Durham, NC). We sequenced this cDNA and found a shorter ORF than the published sequence but homologous to the rat sequence reported here. As Fig. 2A shows, an extra base pair appeared in the published sequence, leading us to believe that a sequencing error may have been responsible for the incorrect published sequence.


View larger version (41K):
[in this window]
[in a new window]
 
Fig. 2.   SCaMPER protein is expressed in rat cardiomyocytes and cardiac tissue. A: CLUSTAL W multiple sequence alignment of the open reading frame of canine, rat, and human SCaMPER protein sequence. * Single, fully conserved residues. B: Western analysis of purified adult and neonatal rat cardiac microsomes (10 µg) probed with the anti-SCaMPER antibody (i); in vitro expression of SCaMPER with and without posttranslation modification (CPM, canine pancreatic microsomes; ii); bacterial expression of SCaMPER with and without inducer (iii); and peptide blocking assay of 10 µg of adult rat cardiac microsomes preincubated with 10 µg/ml of blocking peptide (iv). C: indirect immunofluorescence with anti-SCaMPER antibody of cultured primary neonatal cardiomyocyte (i), adult cardiomyocytes (ii), and adult rat ventricular tissue (iii).

Interesting secondary structural features of the SCaMPER transcript were predicted by computer analyses of the partial cDNA sequence. The Mfold (Rensselaer Polytechnic Institute) prediction program (32) demonstrated the putative presence of many hairpins, stem loops, and other complex structures that likely account for the difficulties we encountered in cloning the SCaMPER cDNA by RT-PCR.

Western analysis with a polyclonal antibody designed against the predicted cytoplasmic domain of SCaMPER demonstrated SCaMPER protein expression in both adult and neonatal rat cardiac membranes (Fig. 2B,i). The SCaMPER antibody identified a single band corresponding to a protein with an approximate Mr of 24. This Mr is in contradiction to the 12-kDa protein predicted from a transcript size of 333 nucleotides (Fig. 1) and suggests that SCaMPER is either heavily glycosylated or may have been dimerized during membrane preparation. To distinguish between these two possibilities, we expressed SCaMPER in a rabbit reticulocyte lysate in vitro assay system in the presence and absence of canine microsomal membranes designed to glycosylate the translated protein (see MATERIALS AND METHODS). The results demonstrate an expressed protein with a Mr of 12 (Fig. 2B,ii) that did not increase by the addition of posttranslational machinery, suggesting that the apparent Mr of 24 seen in SDS-PAGE is a dimerized form of SCaMPER. Furthermore, expression of SCaMPER in a bacterial system lacking eukaryotic posttranslational modification capability revealed both 12-kDa and 24-kDa immunoreactive protein bands (Fig. 2B,iii). Attempts to deglycosylate the SCaMPER protein by treatment with O- and N-glycanase were unsuccessful in reducing the size of the 24-kDa immunoreactive band. Together, the in vitro translation and deglycosylation data suggest that SCaMPER is not posttranslationally modified.

Considering that SCaMPER dimers persist in denaturing/reducing conditions of the SDS-PAGE, it is likely that SCaMPER proteins are in close association in the native membrane and that SCaMPER may, in fact, be a functional dimer. Peptide blocking assays confirmed the specificity of the antibody against SCaMPER. Figure 2B,iv shows that preincubating blotted cardiac membranes with the free peptide used to generate the original anti-SCaMPER antibody completely prevented the antibody from staining proteins in Western blots. These data not only confirm the antibody's specificity but also indicate that the 24-kDa band seen in blots represents a specific product, which is likely a dimerized form of SCaMPER.

Indirect immunolocalization studies with the SCaMPER anti-peptide antibody detected abundant expression of the protein in both neonatal and adult cultured rat cardiomyocytes, as well as in adult rat cardiac tissue sections (Fig. 2C). Importantly, immunostaining was absent from control cells incubated with preimmune serum (data not shown). In adult cells and cardiac tissue, SCaMPER staining was punctate and striated, forming a pattern consistent with its localization to the sarcotubular system. However, the limited resolution at the light microscopic level was not sufficient to distinguish a SR from a transverse tubular (T tubule) localization.

Importantly, cardiomyocytes transfected with antisense SCaMPER demonstrated a knockdown of SCaMPER signal in immunofluorescence micrographs (Fig. 2D). On the other hand, control cells (Fig. 2D, left) transfected with empty vector exhibited normal levels of SCaMPER protein expression. The lack of immunofluorescence staining in antisense-transfected cells confirms the Western immunoblots in demonstrating that the anti-SCaMPER antibody is highly specific for SCaMPER.

Subcellular localization of SCaMPER. The prominent punctate nature of SCaMPER staining seen in Fig. 2C suggested that SCaMPER was not distributed uniformly along the sarcotubular system and that it could be at the junctional regions where the SR and T tubules are located. Accordingly, confocal immunomicroscopy was used to colocalize SCaMPER with two well-recognized junctional proteins, the DHPR of the junctional T tubules and the RyR of the junctional SR. Other antibodies of interest were used, including an antibody to SERCA, a prominent protein of the longitudinal SR. The data in Fig. 3A show a strong colocalization of the SCaMPER antibody with the DHPR and a weaker colocalization with the RyR. Because colocalization of SCaMPER with SERCA was not observed, SCaMPER is not likely associated with the longitudinal SR. Together, these data suggest that SCaMPER is localized to the sarcotubular junction, either the junctional SR or T tubules.


View larger version (92K):
[in this window]
[in a new window]
 
Fig. 3.   SCaMPER is localized to the sarcotubular junction. A: confocal microscopy of isolated primary cardiomyocytes coimmunostained with anti-SCaMPER and anti-sarco(endo)plasmic reticulum Ca2+-ATPase (SERCA), anti-ryanodine receptor (RyR), anti- dihydropyridine receptor (DHPR), or anti-nucleoporin. B: Western blot of highly purified sarcoplasmic reticulum (SR) membranes and membranes enriched in transverse tubules (TT). C: immunoelectron microscopy of rat cardiac tissue with silver enhancement of 1.4-nm immunogold. i: low magnification of SCaMPER-labeled tissue; arrows indicate regions of periodic localization of label in T tubule-rich regions. ii: several high magnifications demonstrating label of T tubule including 1 T tubule invagination point. iii: low magnification of control (preimmune serum)-labeled tissue section.

Interestingly, the anti-SCaMPER antibody also demonstrated a strong nuclear localization (Figs. 2C and 3A). Costaining of the nuclear pore complex with an antibody against nuclear pore O-linked glycoprotein (anti-nuclear pore complex) revealed a lack of SCaMPER localization to the nuclear envelope. On the other hand, prominent immunostaining was observed in the intranuclear regions.

To further elucidate the subcellular origin of SCaMPER, we performed Western blot analyses on both T tubule-rich surface membranes and highly purified SR membranes prepared by the calcium oxalate loading procedure (see MATERIALS AND METHODS). Na+/Ca2+ exchanger and SERCA antibodies were used as markers for the enriched T tubules and SR membrane fractions, respectively. Figure 3B shows that only the T tubule-enriched surface membranes containing the Na+/Ca2+ exchanger displayed the immunoreactive SCaMPER protein. On the other hand, the SR fraction possessing SERCA (Fig. 3B) and phospholamban (not shown) as SR-specific markers did not display SCaMPER protein staining. These data suggest that SCaMPER may be localized to the T tubules and are consistent with confocal colocalization of SCaMPER to the DHPR (Fig. 3A). Because the T tubular fraction could have possible contamination with junctional SR and because the confocal data in Fig. 3A demonstrate some colocalization of SCaMPER with the RyR, the localization of SCaMPER to the junctional SR cannot be ruled out.

Immunoelectron microscopic techniques were subsequently used to determine whether cardiac SCaMPER is associated with the junctional T tubules. Tissue sections treated with the anti-SCaMPER antibody (Fig. 3C, i and ii) revealed the presence of immunogold-enhanced anti-SCaMPER staining to the T tubule system. Low-magnification micrographs of Fig. 3C,i demonstrate the periodic labeling of antibody corresponding to the typical transverse directionality of T tubules. The periodic labeling is consistent with the striated appearance of confocal micrographs in Fig. 3A. Figure 3C,ii is a set of fields taken at higher magnifications clearly demonstrating the presence of label on the T tubules and what appears to be in the junctional regions where the T tubule meets the SR. Moreover, Fig. 3C,ii demonstrates a strong label at a T tubule invagination point. Figure 3C,iii is a representative control micrograph demonstrating that the preimmune IgG did not appreciably label the rat cardiac sections. These data suggest that SCaMPER is a T tubule protein associated with the junctional regions and are consistent with the confocal images (Fig. 3A) as well as Western blots of isolated membranes (Fig. 3B), which also suggest a T tubule origin of the protein. Furthermore, the lack of staining in the mitochondria clearly demonstrates that SCaMPER is not localized to the mitochondria.

SCaMPER regulates cardiac cell calcium. The expression and localization of SCaMPER to the sarcotubular junction where EC coupling is initiated suggests that SCaMPER may play an important role in the regulation of calcium in cardiomyocytes. Consequently, we examined the ability of SCaMPER to alter calcium signaling in cultured neonatal cardiomyocytes. Overexpression of SCaMPER in neonatal cardiomyocytes proved to be toxic to these cultured cells (data not shown) and thus precluded us from using SCaMPER expression constructs in our calcium studies. Such cellular toxicity was also reported for MDCK cells overexpressing SCaMPER (21). These workers observed severe alterations in morphology, followed by cell death when SCaMPER was overexpressed. Consequently, we turned to antisense techniques to knock down endogenous SCaMPER expression to deduce the role of SCaMPER in calcium signaling. This was achieved by transfection of isolated neonatal ventricular cardiomyocytes with expression plasmids containing the SCaMPER ORF in antisense orientation.

The ability of SCaMPER to mediate calcium signaling was studied by real-time calcium transient analysis. Neonatal cardiomyocytes were incubated with the ratiometric calcium dye indo 1-AM and then monitored for real-time alterations of calcium signaling as described previously (17). The positively transfected cells were electrically paced for 20 s to ensure that they were able to release calcium from the SR. After electrical pacing was discontinued, the calcium transients ceased and at 30 s SPC was added to elicit a calcium release, as previously demonstrated in adult cardiomyocytes (1). Routinely, 10 mM caffeine (final concentration) was added at 120 s to demonstrate whether or not the SR calcium store had been mobilized by the sphingolipid.

Shown in Fig. 4, A and B, are typical responses to SPC in control (empty vector) and antisense-transfected cardiomyocytes. As reported previously in adult cardiomyocytes (1), the neonatal cardiac cells released nearly all of their SR calcium in response to SPC. Resting (diastolic) calcium levels increased dramatically, producing asynchronous and spontaneous calcium transients in the absence of electrical pacing. Eventually, the cell reached calcium overload. By 90 s after SPC addition, the SR was essentially devoid of calcium, as evidenced by the lack of a subsequent caffeine response.


View larger version (27K):
[in this window]
[in a new window]
 
Fig. 4.   SCaMPER mediates the sphingosylphosphorylcholine (SPC)-induced calcium release from cardiomyocytes. Isolated neonatal ventricular cardiomyocytes were transfected with empty vector for controls or with antisense SCaMPER expression constructs. SPC-induced calcium deregulation was determined by epifluorescence after incubating cells with indo 1-AM. The cells were electrically paced for 20 s and then stopped before addition of 50 µM SPC at 30 s; 10 mM caffeine was added at 120 s to determine the SR calcium load. Shown here are typical responses to SPC in control (A) and antisense-treated (B) cells. C: cumulative data from several experiments showing resting calcium levels, peak calcium levels in response to SPC, and the difference between resting and peak calcium release in response to SPC. D: comparison of phasic calcium release of electrically paced control and antisense-treated cardiomyocytes. The resting values are an average of oscillatory and diastolic calcium levels the period between 20 and 30 s and the peak values are for the period between 110 and 120 s for n = 8 controls and n = 7 antisense-treated neonatal cardiomyocytes. * Significant (P > 0.001) differences (Student's t-test) between control and antisense-treated cardiomyocytes.

Before SPC addition, the antisense-transfected cardiomyocytes displayed normal pacing-induced calcium transients compared with control (nontransfected) cardiomyocytes, suggesting that the antisense construct had no adverse effect on normal EC coupling, as evidenced by the lack of effect on phasic calcium release or SR calcium content (Fig. 4D). Importantly, cardiomyocytes transfected with antisense SCaMPER expression constructs failed to elicit a robust calcium release in response to 50 µM SPC (compare Fig. 4B with Fig. 4A). These cardiomyocytes responded to SPC with a much delayed and decreased diastolic calcium release signal. Rarely did the SR become depleted of its calcium, because the subsequent addition of caffeine invariably elicited a strong calcium response (Fig. 4B). Importantly, four of seven antisense-transfected cells completely failed to release any calcium in response to SPC. Figure 4C shows the cumulative calcium release responses of both control cardiomyocytes and those treated with antisense. These results indicate that the antisense construct significantly inhibited the effects of SPC, thus implicating SCaMPER in SPC-induced calcium release. The resting levels are representative of the period of time from the end of pacing to just before addition of SPC (20-30 s), and peak values are an average of calcium levels of SPC influence just before addition of caffeine (110-120 s). Interestingly, resting calcium values were significantly decreased (Student's t-test, P < 0.002) for the antisense cardiomyocytes, suggesting a role for SCaMPER in controlling resting calcium levels. More importantly, the differential between peak amplitudes and resting (diastolic) values were significantly (Student's t-test, P < 0.01) decreased by 73% in antisense-transfected cells.

Importantly, the calcium transients of cells treated with a nonsense insert responded exactly like the control (empty vector or nontransfected) cells. Additionally, we used an ER-stress response reporter, GRP78-ERSE-Luc, that we previously used in cardiomyocytes (25) to determine whether the introduction of antisense-SCaMPER itself induced a stress response. The antisense-treated cardiomyocytes failed to elicit a stress response (0.97 ± 0.2% of control cells) as determined by an enhanced luciferase expression. This lack of response is in contrast to the 550 ± 40% induction when cells were treated with tunicamycin as a positive control. Together, these experiments demonstrate that the decrease in SPC-induced calcium release is due to the knockdown of SCaMPER expression rather than the manipulations of transfection or an induction of stress due to experimental procedures.


    DISCUSSION
TOP
ABSTRACT
INTRODUCTION
MATERIALS AND METHODS
RESULTS
DISCUSSION
REFERENCES

This report is the first demonstration of a possible role of the sphingolipid-modulated calcium channel SCaMPER in cardiomyocytes. With RT-PCR, Northern blotting, and a variety of immunological methods, we demonstrate that SCaMPER is expressed in both adult and neonatal cardiomyocytes as well as in adult cardiac tissue.

Several findings presented in this study suggest that SCaMPER is localized subcellularly to the junctional regions of the sarcotubular membrane system where connections between the T tubules and the SR function to regulate cardiac cell calcium levels. Evidence for the sarcotubular localization comes first from the finding that the anti-SCaMPER antibody colocalizes with the DHPR, a known component of the junctional T tubules, and with a junctional SR marker, the RyR. Confocal imaging was confirmed by the demonstration that T tubule-enriched membrane subfractions display the immunoreactive SCaMPER protein in contrast to longitudinal SR membranes, which do not. Finally, immunoelectron microscopy suggests a T tubular localization of SCaMPER. Together these data suggest that SCaMPER may be localized to the T tubules. On the other hand, it is also possible that SCaMPER could be a junctional SR protein, as evidenced by the apparent colocalization with the RyR. The resolution of indirect fluorescence is not sufficiently robust to enable us to distinguish junctional T tubules from junctional SR.

Because SCaMPER has been shown to be a calcium channel expressed by human basophilic leukemia cells and canine MDCK cells (4, 5, 14), localization of this channel to the junctional regions of cardiac cells suggested that it might have an important role in cardiac calcium control. Accordingly, we demonstrated that SPC, the well-recognized sphingolipid ligand for SCaMPER, elicits profound calcium release from the cardiac SR. Moreover, the finding that antisense expression constructs for SCaMPER significantly reduce the calcium response suggests that SCaMPER is the target of SPC action in controlling SR calcium. We had previously demonstrated (1) that the SPC response could be eliminated by pretreating cardiomyocytes with thapsigargin, an agent that depletes SR calcium. However, until this study, it had not been definitively proven that SCaMPER mediates the sphingolipid-triggered calcium response in heart cells. Surprisingly, endogenous knockdown of SCaMPER expression did not alter phasic calcium release, indicating that SCaMPER does not influence normal EC coupling or alter SR calcium content.

Using a SCaMPER expression construct obtained from canine basophilic leukemia cells, Kindman and colleagues (5) demonstrated that patch-clamped oocytes overexpressing SCaMPER displayed a functional SPC-modulated calcium channel. These workers subsequently described SCaMPER as having a 546-bp ORF (14). However, Zacchetti's group (21) recently argued that the MDCK cDNA is substantially smaller. We have confirmed Zacchetti's group's findings in that the rat and human cardiac SCaMPER cDNA is 333 bp in length. We obtained the clone from Kindman and discovered that the sequence of the clone used in his functional expression studies was in fact 333 bp, suggesting that a sequencing error led Kindman et al. to believe that the ORF was larger than it was. It is important to emphasize that the expression construct used in the current work to prove that SCaMPER is a calcium channel came from the clone that was used in the oocyte expression work previously published (14). Thus we confirm in the present study that SCaMPER is likely a calcium channel or a calcium channel modulator. However, it is likely that the shorter ORF reported here is a single-pass membrane protein rather than the double-pass membrane protein suggested previously (14).

Calcium regulation is an essential feature of cardiac function. Because sphingolipids are gaining increasing attention as important signaling molecules in the cardiovascular system (8), an understanding of the role of SCaMPER in cardiac function will be of great interest in explaining the growing list of physiological effects on the heart attributed to sphingolipid mediators such as SPC and S1P. For example, it was recently shown that SCaMPER's putative ligand, SPC, is a normal constituent of serum (9) and that it may be involved in cardiac hypertrophy (22) and in the regulation of cardiac chronotropy and inotropy (9). Our current findings and our previous work (1) demonstrating that skeletal and cardiac calcium is modulated by SPC are consistent with the well-established role of SPC in regulating cellular calcium in other cell types (15, 29).

One might infer from the finding that SCaMPER is expressed on the surface membranes (i.e., T tubules) that some of the extracellular actions attributed to SPC as an extracellular signaling molecule (16) may be explained by the mediator's effects on SCaMPER. However, additional work is needed to elucidate this and to distinguish extracellular from intracellular actions of sphingolipids in the heart. In addition, it has been demonstrated that SPC is a possible ligand for Edg and other GPCRs (18, 30), some of which are expressed by rat cardiac cells (17). Thus it becomes important to distinguish SCaMPER from the GPCRs as putative targets of sphingolipid action in the heart.

The antisense knockdown approach used in this study provides clear demonstration that the SCaMPER protein has a potential physiological role in mediating SPC effects in heart cells. We have shown not only that SCaMPER is expressed in cardiomyocytes, but more importantly, that it is a modulator of calcium regulation in these cells. Experiments in progress will determine the oligomeric nature of SCaMPER and how it functions as a calcium channel. The expression and localization of SCaMPER to the sarcotubular junction set the stage for this protein to play an important role in the regulation of cardiac calcium.


    ACKNOWLEDGEMENTS

We thank Doriana Sandoná for help with the design of the peptides for antibody development.


    FOOTNOTES

This work was supported by grants from Medlyte, Inc. (R. A. Sabbadini), Consiglio Nazionale delle Ricerca, Italy (R. Betto), and Telethon Italy (no. 1268 to R. Betto).

Address for reprint requests and other correspondence: R. A. Sabbadini, Dept. of Biology, San Diego State Univ., San Diego, CA 92182-4614 (E-mail: rsabba{at}sunstroke.sdsu.edu).

The costs of publication of this article were defrayed in part by the payment of page charges. The article must therefore be hereby marked "advertisement" in accordance with 18 U.S.C. Section 1734 solely to indicate this fact.

First published November 6, 2002;10.1152/ajpcell.00382.2002

Received 22 August 2002; accepted in final form 5 November 2002.


    REFERENCES
TOP
ABSTRACT
INTRODUCTION
MATERIALS AND METHODS
RESULTS
DISCUSSION
REFERENCES

1.   Betto, R, Teresi A, Turcato F, Salviati G, Sabbadini RA, Krown KC, Glembotski CC, Dettbarn C, Yasui K, and Palade PT. Sphingosylphosphorylcholine modulates the ryanodine receptor Ca2+ release channel of cardiac sarcoplasmic reticulum membranes. Biochem J 322: 327-333, 1997[ISI][Medline].

2.   Biral, D, Senter L, and Salviati G. Increased expression of dystrophin, beta -dystroglycan and adhalin in denervated muscles. J Muscle Res Cell Motil 17: 523-632, 1996[ISI][Medline].

3.   Goetzl, EJ, and An S. Diversity of cellular receptors and functions for the lysophospholipid growth factors lysophosphatidic acid and sphingosine-1-phosphate. FASEB J 12: 1589-1598, 1998[Abstract/Free Full Text].

4.   Kim, S, Lakhani V, Costa DJ, Sharara AI, Fitz JG, Huang L, Peters KG, and Kindman LA. Sphingolipid-gated Ca2+ release from intracellular stores of endothelial cells is mediated by a novel Ca2+-permeable channel. J Biol Chem 270: 5266-5269, 1995[Abstract/Free Full Text].

5.   Kindman, LA, Kim S, McDonald TV, and Gardner P. Characterization of a novel intracellular sphingolipid-gated Ca2+-permeable channel from rat basophilic leukemia cells. J Biol Chem 269: 13088-13091, 1994[Abstract/Free Full Text].

6.   Krown, KA, Page MT, Nguyen C, Zechner D, Gutierrez V, Comstock KL, Glembotski CC, Quintana PJE, and Sabbadini RA. TNFalpha -induced apoptosis in cardiac myocytes: involvement of the sphingolipid signaling cascade in cardiac cell death. J Clin Invest 98: 2854-2865, 1996[Abstract/Free Full Text].

7.   Lee, H, Goetzl EJ, and An S. Lysophosphatidic acid and sphingosine 1-phosphate stimulate endothelial cell wound healing. Am J Physiol Cell Physiol 278: C612-C618, 2000[Abstract/Free Full Text].

8.   Levade, T, Auge N, Jan Veldman R, Cuvillieer O, Negre-Salvayre A, and Salvayre R. Sphingolipid mediators in cardiovascular cell biology and pathology. Circ Res 89: 957-968, 2001[Abstract/Free Full Text].

9.   Liliom, K, Sun G, Bunemanns M, Virag T, Nusser N, Baker DL, Wang D, Fabian M, Brandts B, Bender K, Eickel A, Malik K, Miller DD, Desiderio D, Tigyi G, and Pott L. Sphingosylphosphocholine is a naturally occurring lipid mediator in blood plasma: a possible role in regulating cardiac function via sphingolipid receptors. Biochem J 355: 189-197, 2001[ISI][Medline].

10.   Liu, G, Kleine L, and Hebert RL. Advances in the signal transduction of ceramide and related sphingolipids. Crit Rev Clin Lab Sci 36: 511-573, 1999[ISI][Medline].

11.   Liu, Y, Wada R, Yamashita T, Mi Y, Deng C, Hobson JP, Rosenfeldt HM, Nava VE, Chae S, Lee M, Liu CH, Hla T, Spiegel S, and Proia RL. Edg-1, the G protein-coupled receptor for sphingosine-1-phosphate, is essential for vascular maturation. J Clin Invest 106: 951-961, 2000[Abstract/Free Full Text].

12.   Lynch, KR, and Im I. Life on the edg. Trends Pharmacol Sci 20: 473-5, 1999[ISI][Medline].

13.   MacDonell, K, Severson D, and Giles W. Depression of excitability by sphingosine 1-phosphate in rat ventricular myocytes. Am J Physiol Heart Circ Physiol 275: H2291-H2299, 1998[Abstract/Free Full Text].

14.   Mao, C, Kim SH, Almenoff JS, Rudner XL, Kearney DM, and Kindman LA. Molecular cloning and characterization of SCaMPER, a sphingolipid Ca2+ release-mediating protein from endoplasmic reticulum. Proc Natl Acad Sci USA 93: 1993-1996, 1996[Abstract/Free Full Text].

15.   Meyer zu Herindorf, D, Himmel H, and Jakobs K. Sphingosylphosphorylcholine-biological functions and mechanisms of action. Biochim Biophys Acta 1582: 178-179, 2002[ISI][Medline].

16.   Meyer zu Heringdorf, D, Niederdraing N, Neumann E, Frode R, Lass H, Van Koppen CJ, and Jakobs KH. Discrimination between plasma membrane and intracellular target sites of sphingosylphosphorylcholine. Eur J Pharmacol 354: 113-122, 1998[ISI][Medline].

17.   Nakajima, N, Cavalli AL, Biral D, Glembotski CC, McDonough PM, Ho PD, Betto R, Sandora D, Palade PT, Dettbarn CA, Klepper RE, and Sabbadini RA. Expression and characterization of Edg-1 receptors in rat cardiomyocytes: calcium deregulation in response to sphingosine-1-phosphate. Eur J Biochem 267: 5679-5686, 2000[Abstract/Free Full Text].

18.   Okamoto, H, Takuwa N, Yatomi Y, Gonda K, Shigematsu H, and Takuwa Y. EDG3 is a functional receptor specific for sphingosine 1-phosphate and sphingosylphosphorylcholine with signaling characteristics distinct from EDG1 and AGR16. Biochem Biophys Res Commun 260: 203-208, 1999[ISI][Medline].

19.   Ralston, E, and Ploug T. Caveolin-3 is associated with the t-tubules of mature skeletal muscle fibers. Exp Cell Res 246: 510-515, 1999[ISI][Medline].

20.   Sabbadini, RA, and Okamoto VR. The distribution of ATPase activities in purified transverse tubular membranes. Arch Biochem Biophys 223: 107-119, 1983[ISI][Medline].

21.   Schnurbus, R, De Pietri Tonelli D, Grohavaz F, and Zacchetti D. Re-evaluation of primary structure, topology, and localization of scamper, a putative intracellular Ca2+ channel activated by sphingosylphosphocholine. Biochem J 362: 183-189, 2002[ISI][Medline].

22.   Sekiguchi, K, Yokoyama T, Kurabayashi M, Okajima F, and Nagai R. Sphingosylphosphorylcholine induces a hypertrophic growth response through the mitogen-activated protein kinase signaling cascade in rat neonatal cardiac myocytes. Circ Res 85: 1000-1008, 1999[Abstract/Free Full Text].

23.   Shields, P, Dixon JE, and Glembotski CC. The secretion of atrial natriuretic factor-(99-126) by cultured cardiac myocytes is regulated by glucocorticoids. J Biol Chem 263: 12619-12628, 1988[Abstract/Free Full Text].

24.   Sugiyama, A, Yatomi Y, Ozaki Y, and Hashimoto K. Sphingosine 1-phosphate induces sinus tachycardia and coronary vasoconstriction in the canine heart. Cardiovasc Res 46: 119-125, 2000[ISI][Medline].

25.   Thuerauf, D, Hoover H, Meller J, Hernandez J, Su L, Andrews C, Dillmann W, McDonough P, and Glembotski C. Sarco/endoplasmic reticulum calcium ATPase-2 expression is regulated by ATF6 during the endoplasmic reticulum stress response. J Biol Chem 276: 48309-48317, 2001[Abstract/Free Full Text].

26.   Van Brocklyn, JR, Lee M, Menzeleev R, Olivera A, Edsall L, Cuvillier O, Thomas DM, Coopman PJP, Thangada S, Liu CH, Hla T, and Speigel 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 142: 229-240, 1998[Abstract/Free Full Text].

27.   Xia, P, Gamble J, Rye K, Wang L, Hii C, Cockerill P, and Kew-Goodall Y. Tumor necrosis factor-alpha induces adhesion molecule expression through the sphingosine kinase pathway. Proc Natl Acad Sci USA 95: 14196-14201, 1988[Abstract/Free Full Text].

28.   Xia, P, Vadas MA, Rye K, Barter PJ, and Gamble JR. High density lipoproteins (HDL) interrupt the sphingosine kinase signaling pathway. J Biol Chem 274: 33143-33147, 1999[Abstract/Free Full Text].

29.   Xu, Y. Sphingosylphosphorylcholine and lysophosphatidylcholine G protein-coupled receptors and receptor-mediated signal transduction. Biochim Biophys Acta 1582: 81-88, 2002[ISI][Medline].

30.   Xu, Y, Zhu K, Hong G, Wu W, Baudhuin LM, Xiao Y, and Damron DS. Sphingosylphosphorylcholine is a ligand for ovarian cancer G-protein-coupled receptor 1. Nat Cell Biol 2: 261-267, 2000[ISI][Medline].

31.   Zechner, D, Craig R, Hanford D, McDonough PM, Sabbadini RA, and Glembotski CC. MKK6 inhibits myocardial cell apoptosis via a p38 MAP kinase-dependent pathway. J Biol Chem 273: 8232-8239, 1998[Abstract/Free Full Text].

32.   Zuker, M, Mathews D, and Turner D. Algorithms and Thermodynamics for RNA Secondary Structure Prediction: A Practical Guide in RNA Biochemistry and Biotechnology, edited by Barciszewski J, and Clark BFC. Kluwer Academic, 1999, p. 11-43. (NATO ASI Series)


Am J Physiol Cell Physiol 284(3):C780-C790
0363-6143/03 $5.00 Copyright © 2003 the American Physiological Society