ORP3 Splice Variants and Their Expression in Human ... - DRO - Deakin

2 downloads 109 Views 469KB Size Report
Mary Ann Liebert, Inc. Pp. 1–9 ... sterol, 25-hydroxycholesterol (25-OHC) (Dawson et al., 1989). ... action of lipid factors and oxysterols (Lagace et al., 1997;.
Deakin Research Online This is the published version: Collier, F. M., Gregorio-King, C, Apostolopoulos, J., Walder, Ken and Kirkland, M. A. 2003, ORP3 splice variants and their expression in human tissues and hematopoietic cells, DNA and cell biology, vol. 22, no. 1, pp. 1-9. Available from Deakin Research Online: http://hdl.handle.net/10536/DRO/DU:30009437 Reproduced with the kind permissions of the copyright owner. Copyright : 2003, Mary Ann Liebert Publishers

DNA AND CELL BIOLOGY Volume 22, Number 1, 2003 © Mary Ann Liebert, Inc. Pp. 1–9

ORP3 Splice Variants and Their Expression in Human Tissues and Hematopoietic Cells F.M. COLLIER,1 C.C. GREGORIO-KING, 1 J. APOSTOLOPOULOS, 2 K. WALDER, 3 and M.A. KIRKLAND 1

ABSTRACT ORP3 is a member of the newly described family of oxysterol-binding protein (OSBP)-related proteins (ORPs). We previously demonstrated that this gene is highly expressed in CD341 hematopoietic progenitor cells, and deduced that the “full-length” ORP3 gene comprises 23 exons and encodes a predicted protein of 887 amino acids with a C-terminal OSBP domain and an N-terminal pleckstrin homology domain. To further characterize the gene, we cloned ORP3 cDNA from PCR products and identified multiple splice variants. A total of eight isoforms were demonstrated with alternative splicing of exons 9, 12, and 15. Isoforms with an extension to exon 15 truncate the OSBP domain of the predicted protein sequence. In human tissues there was specific isoform distribution, with most tissues expressing varied levels of isoforms with the complete OSBP domain; while only whole brain, kidney, spleen, thymus, and thyroid expressed high levels of the isoforms associated with the truncated OSBP domain. Interestingly, the expression in cerebellum, heart, and liver of most isoforms was negligible. These data suggest that differential mRNA splicing may have resulted in functionally distinct forms of the ORP3 gene.

INTRODUCTION

et al., 2002). It is speculated that the OSBP translocation and binding plays a role in the transport, metabolism, or regulatory action of lipid factors and oxysterols (Lagace et al., 1997; Levine and Munro, 1998). Oxysterols are oxygenated derivatives of cholesterol, which enter the cell from the circulation via the low density lipoprotein (LDL) receptor or can be synthesised de novo. Their binding to OSBP has been well characterized, and their biologic actions include inhibition of cellular proliferation and induction of apoptosis (Brown and Jessup, 1999). Recently, we have shown that oxysterols are potent inhibitors of proliferation of hematopoietic cells (Gregorio-King et al., 2002). It has been postulated that all members of the ORP family have a role in the binding of oxysterols or other lipid moieties, and therefore, may play a role in the transduction of the cellular effects (Lagace et al., 1997). Of the related proteins, ORP1 and ORP2 were found to bind phospholipids but not 25-OHC (Xu et al., 2001). Further characterization implicated ORP2 as a regulator of cellular sterol homeostasis and intracellular membrane trafficking (Laitinen et al., 2002). Full-length ORP4 (ORP4-L) and a related splice vari-

O

RPs (OXYSTEROL -BINDING PROTEIN (OSBP)-related proteins) are a family of closely related gene sequences with two major structural features: a highly conserved oxysterolbinding domain present in the C-terminal region, and, in most cases, an N-terminal pleckstrin homology (PH) domain. Twelve human ORPs or OSBPLs (oxysterol binding protein-like), all containing the eight amino acid oxysterol-binding signature, have now been described (Jaworski et al., 2001; Lehto et al., 2001), although the function of most of these proteins has not been investigated. Early studies (Kandutsch and Thompson, 1980) identified the first ORP (designated OSBP) and showed that it is a cytosolic protein with a high affinity for the oxysterol, 25-hydroxycholesterol (25-OHC) (Dawson et al., 1989). Further studies determined that with ligand (25-OHC) binding, OSBP translocates and binds extrinsically to the Golgi membranes (Ridgway et al., 1992). This action is mediated by the PH domain (Lagace et al., 1997; Levine and Munro, 1998), and involves interaction with VAMP-associated protein-A (Wyles

1 Stem

Cell Laboratory, Douglas Hocking Research Institute, Barwon Health, The Geelong Hospital, Geelong, VIC, Australia. Unit, Australian Red Cross Blood Service, Victoria Research Unit, Southbank, Australia. 3 Metabolic Research Unit, School of Health Sciences, Deakin University, Geelong, VIC, Australia. 2 Research

1

2 ant (ORP4-S) were overexpressed in CHO cells (Wang et al., 2002) and experiments showed that ORP4-L bound 25-OHC with high affinity. The splice variant, in which the PH domain and part of the OSBP domain were deleted, did not bind oxysterols but interacted with the cytoskeletal network of vimentin intermediate filaments. In addition, cells overexpressing ORP4S had a 40% reduction in the esterification of LDL-derived cholesterol, implying that interaction with the vimentin network modifies cholesterol transport. Although normal tissues appear to express all ORPs, there are tissue-specific differences (Jaworski et al., 2001; Lehto et al., 2001). In general, the brain and kidney express the highest levels of the various ORPs, and this has been postulated to be due to the active membrane transport processes in these tissues (Laitinen et al., 1999). The ORPs have been classified into six separate subfamilies according to sequence similarity and shared protein features. It is interesting to note that ORPs within a subfamily, although similar in genomic structure, have varied tissue patterns (Lehto et al., 2001). ORP4 (designated the HeLa metastatic gene) was found to be differentially expressed in metastatic tumour cells (Fournier et al., 1999). In addition, ORP4 mRNA levels have been indicated as a potential marker for hematologic malignancies (Silva et al., 2001), while other studies demonstrate high expression of ORP4 in retina, testis, and fetal liver (Moreira et al., 2001). A recent study (Lehto et al., 2001) investigated whether the expression of human ORPs is influenced by changes in the cellular cholesterol status, and showed that ORP6 was the only gene upregulated in acidified LDL-loaded macrophages. In a previous study we identified the ORP3 gene, and found that it is more highly expressed in CD34 positive (CD341 ) compared with CD34 negative hematopoietic cells (CD342 ) (Gregorio-King et al., 2001). We deduced the complete ORP3 gene from RACE, EST sequences, and homology to genomic clones, to give the “full-length” ORP3, comprising 23 exons and encoding a protein containing 887 amino acids (Accession No: AY008372). ORP3 displays the characteristic PH and OSBP domains, and like some other ORPs, a putative leucine zipper and dimerisation region. The size and distribution of the ORP3 exons is very similar to ORP6 and ORP7, and therefore, they have been grouped into one of the ORP subfamilies (subfamily III) (Lehto et al., 2001). In addition to hematopoietic cells, ORP family studies have demonstrated ORP3 expression in spleen and leukocytes (Laitinen et al., 1999), kidney and thymus (Lehto et al., 2001). Studies focussing on the human ORP gene family, combined with evidence from the NCBI databases, indicate that there are a number of examples of alternative splicing (Jaworski et al., 2001; Lehto et al., 2001); however, there are few specific examples. A recent study (Wang et al., 2002), identified a shortened form of ORP4 that encoded an alternative exon between exons 2 and 3, resulting in deletion of the PH domain and part of the oxysterol-binding domain. Western blot analysis of ORP2 in mouse tissues also revealed two major immunoreactive forms; possibly representing translated variants (Laitinen et al., 2002). Further investigation of the ORP3 gene by our laboratory revealed a large number of previously undescribed splice variants. In this study we identify these isoforms and investigate their expression in various tissues.

COLLIER ET AL.

MATERIALS AND METHODS Preparation of hematopoietic mononuclear cells and cord blood CD341 samples Umbilical cord blood samples (UCB) were obtained after uncomplicated vaginal or cesarean delivery with clamping and cutting of the cord and drainage into sterile collection tubes. All samples were donated by volunteers according to approved institutional guidelines. Mononuclear cells (MNC) were prepared by density gradient separation through Ficoll-Hypaque (Amersham Pharmacia Biotech, Uppsala, Sweden). CD341 cells were isolated immunomagnetically by positive selection using antibody labelling and a MiniMacs™ bead separation kit following the manufacturers instructions (Miltenyi Biotec, Becton Dickinson, Fullerton, CA). Viability was assessed using Trypan Blue staining and purity of CD341 populations was determined by flow cytometric analysis (FacsCalibur, Becton Dickinson). CD341 samples of greater than 70% purity were used for RNA extraction.

Total RNA isolation and reverse transcription Total RNA from MNC and CD341 and CD342 samples were extracted using Trizol (Invitrogen, San Diego, CA) according to the manufacturer’s instructions. A human RNA tissue panel was purchased from Clontech (Palo Alto, CA). Prior to cDNA synthesis all samples were DNase (Promega, Madison, WI) treated to remove any contaminating genomic DNA. One microgram of DNA-free RNA was reverse transcribed using random hexamers in the Reverse Transcription System (Promega, Madison, WI) according to the manufacturer’s guidelines. All reactions were performed on a Perkin-Elmer (Norwalk, CT) 9600 thermocycler. cDNA from eight CD341 and CD342 samples were pooled respectively and used in subsequent PCR reactions.

PCR and cloning Primers were designed (using Primer Express software) to amplify the full coding region of ORP3 (Table 1), and PCR carried out using MNC samples and a liver cDNA library (generously supplied by Professor GR Collier, Deakin University) with Platinum Taq polymerase (Invitrogen) and touchdown cycling (61°C reducing to 58°C) for 44 cycles. Amplified products were visualized by standard agarose gel electrophoresis and ethidium bromide staining. Bands of the correct size were excised from the gel, purified (Clontech Nucleo-Trap), and cloned using PCRscript (Clontech) or TOPO-TA (Invitrogen) cloning kits. Colonies were picked, restriction enzyme digests performed to confirm inserts, plasmid DNA prepared (Eppendorf Perfectprep), and the inserts sequenced using an ABI Prism™ 373 DNA Sequencer (P.E. Applied Biosystems, Foster City, CA) and vector primers.

Semiquantitative PCR Investigation of splice variants was performed on a selection of human tissue cDNA samples (bone marrow, cerebellum, total brain, heart, kidney, liver, lung, skeletal muscle, spleen, thymus, thyroid, and colon) and pooled UCB CD341 and CD342 cDNA. Specific primers (Primer Express® , PE Applied Biosys-

3

ORP3 ISOFORMS AND THEIR TISSUE DISTRIBUTION TABLE 1. Name

SEQUENCES

OF

PRIMER PAIRS

Forward (59-39) primer

Full coding region ORP3 GAPDH ORP3 primer set 1 ORP3 primer set 2 ORP3 primer set 3 ORP3 primer set 4

TGGATAAGAGCTAAACCACCTG CAGTCAGCCGCATCTTCTTTTG GCATCGGACATACTCGGCAC GCATCGGACATACTCGGCAC CAACGCCATCCAGGTCCCGAAAC CCAGAGCTATTGGCAAAGATG

tems) were designed to amplify GAPDH mRNA, and two regions of the ORP3 mRNA. Primer set 1 (Table 1) amplified a region from within exon 8 to the exon 15/16 border, while primer set 2 (Table 1) spanned a region from within exon 8 to an extended region of exon 15 in the ORP3 gene. PCR reactions were performed using Platinum Taq polymerase (Invitrogen), standard conditions and 39 cycles (ORP3, primer set 1 and 2) or 28 cycles (GAPDH). For both of the ORP3 primer sets, four products were amplified representing the various isoforms, although only three bands were easily distinguishable. The products were designated A, B, C, and D for each primer set, and each of the products was confirmed by either restriction enzyme digest, or a subsequent nested PCR (see below), and sequencing (as described above). As PCR products B and C were seen as one band that differed in size by only 15 base pairs, a post-PCR restriction enzyme digest (Nde 1) was performed in an effort to further separate the bands (Primer set 1 only). This resulted in truncation of all the amplicons and bands that were more easily identified on the gel. To further investigate PCR products B and C, nested PCR reactions were performed to separately amplify each isoform. The initial PCR reaction (primer set 1 and 2) was reperformed for 20 cycles on the panel of cDNA, the product diluted 1:2 and used in a secondary reaction with primers for the specific isoforms (primer set 3 and 4; Table 1). PCR reactions were performed using Platinum Taq polymerase (Invitrogen), standard conditions and 21 cycles. All PCR products were visualized by 2% agarose gel electrophoresis and ethidium bromide staining, and the relative band intensities for all PCR reactions evaluated using the Kodak EDAS system and semiquantified relative to GAPDH. Each PCR reaction was repeated in triplicate.

Reverse (59-39) primer AGGCAAGCACAGGAGAAATACAC TGGTTCACACCCATGACGAAC GCTGCCACATATACCATCCTTTCC CCAGACAGCTCTGAGTAATG CATGGATTCTGCGTAAGCGTTC CTGGAGTTTTCGGATGACAGG

while the product derived from the liver library was approximately 400 nucleotides larger (Fig. 1). Both products were ligated into “housekeeping” vectors and the colonies picked revealed a number of different ORP3 splice variants. BLAST searches on the NCBI database indicated that the PCR product from MNC resulted in four isoforms with variations in the inclusion or exclusion of exons 9 and/or 12 (Fig 2i). The products from the liver cDNA library included variants with a larger exon 15, as well as inclusion/exclusion of exon 9. In these cases, the new extended exon 15 was 399 base pairs longer than previously shown, with additional sequence matching to the genomic clone (Accession number, AC004016) and previously regarded as postexon 15 intronic. Further investigation of other tissues by PCR revealed that all exon 9/12 variations were also expressed by the extended isoform 15 (Fig. 2ii). The other exons in the clones were identical to the “full-length” ORP3. To clarify the nomenclature, the eight isoforms were designated ORP3 (1a–1d) 5 all isoforms with “standard exon 15” (Accession No: AF491781–AF491784), or ORP3 (2a–2d) 5 all isoforms with “extended exon 15” (Accession No: AF491785–86 and AF515639–40) (Fig. 2i and ii). Therefore “full-length” ORP3 (1a) is equivalent to the standard exon 15 variant which includes both exon 9 and 12 (Fig. 2i).

Conceptual translations of ORP3 isoforms Translation of “full-length” ORP3 isoform (1a) is initiated by a start codon in exon 2, extends for 2664 nucleotides, and terminates in exon 23. ORP3 isoforms (1b–1d) skip exons 9 and/or 12, which have lengths of 93 nucleotides and 108 nucleotides, respectively. These deletions shorten the ORP3 protein by 31 amino acids and 36 amino acids, respectively, but stay in frame, do not modify the PH or OSBP domains, or cause a premature stop (Fig. 3).

Bioinformatics Nucleic acid and protein sequences were analyzed using software available from the National Center for Biotechnology Information (NCBI) and Swiss Institute of Bioinformatics (SIB) databases. Alignment of splice variant sequences was performed using Bioedit.

RESULTS Identification of ORP3 isoforms PCR was utilised to amplify the full coding region of ORP3 (Accession number, AY008372). The PCR product from the MNC resulted in a band at the expected size (2815 nucleotides),

FIG. 1. PCR amplification of the full coding region of ORP3 for cloning. Lanes (from left) are molecular size ladder, human liver library cDNA (Clontech), and reverse transcribed MNC cDNA.

4

COLLIER ET AL.

In contrast, conceptual translation of the extended exon 15 isoforms (ORP3 (2a–2d)), alters the reading frame 96 amino acids into the OSBP domain, and results in an additional 59 amino acids of novel sequence before the translation terminates. The majority of the oxysterol-binding domain and the OSBP signature are omitted and the novel

sequence does not contain any conserved motifs or protein patterns. Similar deletions of exon 9 and/or 12 occur in the extended exon 15 isoforms (Fig. 3). None of the exon variations (exons 9, 12, or 15) identified here alter the putative dimerization box or leucine zipper region (Gregorio-King et al., 2001).

i

ii

FIG. 2. Splice variants of ORP3 and position of specific primer pairs. Schematic diagram of exon 8 to exon 16 of the ORP3 coding region displaying areas of alternative splicing and the various isoforms for the standard exon 15, ORP3 (1a–1d), (i) and extended exon 15, ORP3 (2a–2d), (ii). The shaded boxes denote the exons, with the alternatively spliced exons represented by open boxes. The primer pairs for ORP3 isoform detection are marked and shown as a thick line. If they span an exon–exon boundary they are joined by a curved line.

ORP3 ISOFORMS AND THEIR TISSUE DISTRIBUTION

5

FIG. 3. Diagrammatic representation of conceptual proteins for ORP3 (1a–1d) and (2a–2d) isoforms. The number of amino acids indicates the predicted size. Group (i): predicted proteins for standard exon 15 isoform, ORP3 (1a–1d). Group (ii): predicted proteins for extended exon 15 isoforms, ORP3 (2a–2d). The pleckstrin homology (PH) and OSBP domains are labeled and shaded. The highly conserved sequence motif “EQVSHHPP” is marked in the OSBP region. In ORP3 (2a–2d) the OSBP domain is truncated, the motif EQVSHHPP deleted, and there is an area of novel sequence, marked by crosshatching.

Semiquantitative PCR To evaluate the respective amounts of the various isoforms in a selection of tissue samples, PCR reactions were performed using specific primer sets (Fig. 2i and ii) for the ORP3 isoforms and semiquantitated relative to GAPDH mRNA expression. Due to the common characteristics of the isoforms, RT-PCR was determined to be the only way to investigate the distribution of all eight isoforms. The standard exon 15 isoforms, ORP3 (1a–1d), were highly expressed in brain, kidney, spleen, thymus, thyroid, colon, and cord blood CD341 and CD342 hematopoietic cells. Interestingly, in contrast to the whole brain, the level of mRNA expression in cerebellum was very low (Fig. 4, Graph i, lane 2). ORP3 (1a) mRNA was the most common isoform with expression in kidney, whole brain, thyroid, and skeletal muscle, but with no demonstrable expression in heart and liver. With the exception of CD342 hematopoietic cells, expression of

ORP3 (1d) was low, with no expression observed in brain and cerebellum (Fig. 4, graph i). To further investigate ORP3 (1b) and (1c) (exon 9-less and exon 12-less) isoforms, nested PCR reactions were performed with primer pairs specific for the variants (primer set 3 and 4, Fig. 2). As with ORP3 (1a), expression of ORP3 (1b) and (1c) was demonstrated in all tissues except heart and liver. Expression of ORP3 (1c) was highest in total brain, kidney, lung, skeletal muscle, thyroid, and cord blood CD341 and CD342 cells. Colon was the only tissue with higher expression of ORP3 (1b) than (1c) (Fig. 4, graph ii). The extended exon 15 isoforms (ORP3 2a–2d), which result in a truncation of the OSBP region, were expressed in most tissues except cerebellum, heart, liver, and lung; with the highest expression in whole brain, kidney, spleen, thymus, thyroid, and CD34 1 cells. Expression of ORP3 (2a) was generally greater than ORP3 (2b), (2c), or (2d). The ORP3 (2d) isoform was expressed in spleen, thymus, colon, and

6

COLLIER ET AL.

FIG. 4. Semiquantitative RT-PCR analysis of (ORP3 (1a–1d)) in bone marrow (1), cerebellum (2), whole brain (3), heart (4), kidney (5), liver (6), lung (7), skeletal muscle (8), spleen (9), thymus (10), thyroid (11), colon (12), cord blood CD341 (13), cord blood CD342 (14), and no template control (15). The far left lane contains the DNA ladder. The numbers below graph (ii) represent the same tissues. (A) PCR analysis of ORP3 (1a–1d) using Primer set 1, with post-PCR Not1 enzyme digest. (B) Nested PCR analysis of ORP3 (1b) using Primer set 3. (C) Nested PCR analysis of ORP3 (1c) using Primer set 4. (D) PCR analysis of GAPDH. Gene expression of the various isoforms was measured by the Kodak EDAS system and quantitated relative to GAPDH. Graph (i): Relative intensities of ORP3 isoforms in gel image (A). Graph (ii): Relative intensities of ORP isoforms in gel images (B) and (C).

CD341 and CD34 2 cells, but at fairly low levels in comparison to the other isoforms. The nested PCR reactions for the separate identification of the ORP3 (2b) and (2c) isoforms revealed expression in all the tissues as predicted from the original PCR. There was far greater expression of ORP3 (2c) than (2b) in the total brain and skeletal muscle; with the reverse seen in thymus bone marrow and CD341 cells. All four extended exon 15 isoforms (ORP3 2a–2d) had very low mRNA expression in CD34 2 cells (Fig. 5).

DISCUSSION It has been estimated that, based on analyses of EST, the transcripts from 35% of human genes are alternatively spliced (Graveley, 2001) and recent studies indicate there is ample evidence for multiple splicing in the ORP family (Jaworski et al., 2001; Lehto et al., 2001). We have demonstrated eight splice variants of the ORP3 gene with specific tissue distributions. The alternative splicing of the isoforms includes exon “skip-

ORP3 ISOFORMS AND THEIR TISSUE DISTRIBUTION

7

FIG. 5. Semiquantitative RT-PCR analysis of ORP3 (2a–2d) in bone marrow (1), cerebellum (2), whole brain (3), heart (4), kidney (5), liver (6), lung (7), skeletal muscle (8), spleen (9), thymus (10), thyroid (11), colon (12), cord blood CD341 (13), cord blood CD342 (14), and no template control (15). The far left lane contains the DNA ladder. The numbers below graph (ii) represent the same tissues. (A) PCR analysis of ORP3 (2a–2d) using Primer set 2. (B) Nested PCR analysis of ORP3 (2b) using Primer set 3. (C) Nested PCR analysis of ORP3 (2c) using Primer set 4. (D) PCR analysis of GAPDH. Gene expression of the various isoforms was measured by the Kodak EDAS system and quantitated relative to GAPDH. Graph (i): Relative intensities of ORP3 isoforms in gel image (A). Graph (ii): Relative intensities of ORP isoforms in gel images (B) and (C). ping” and exon “elongation” at the 59 splice site. The extension to exon 15 is confirmed by a number of sequences from the EST database (Accession Nos: BF771076, BF770483, BI771299, BF771055, BF770475, AW798464); while the skipping of exon 9 and/or 12 do not appear to be currently represented. However, a mouse clone (Accession No: AK004768), with significant homology to ORP3, exhibits the deletion of the equivalent of human ORP3 exon 9. As ORP3, ORP 6, and ORP7 make up ORP subfamily III (Lehto et al., 2001), and their exons 9, 12, and 15 are almost identical in size, it is possible that the same splice variations demonstrated in ORP3 may occur in ORP6 and ORP7. In the current study, aside from the eight iso-

forms described here, we have identified further splice variations at the start of exon 3 (deletion of five base pairs) and exon 11 (deletion of 16 base pairs), respectively (data not shown). The nucleotides at these alternate sites exhibit a conserved 39 splice site sequences (Shapiro and Senapathy, 1987). In the standard exon 15 isoforms (ORP3 (1a–1d)) the alternative splicing of exon 9 and/or 12 shortens the translated protein but keeps both the OSBP and PH domains intact. As these domains are implicated in oxysterol and phospholipid binding and translocation to the Golgi apparatus, respectively, these isoforms would presumably retain the functional characteristics of OSBP. The deletion of exon 9 (31 amino acids) would result

8 in the loss of two PKC phosphorylation sites while exon 12 (36 amino acids) has no commonly recognised patterns. The extension of exon 15 [ORP3 (2a–2d)] results in a shift in the reading frame, the loss of the majority of the oxysterol-binding domain and additional 59 amino acids of novel sequence. Expasy software indicates that this additional sequence displays no currently described domains or motifs. It is, therefore, difficult to predict the function, if any, of these splice variants. However, the truncated ORP3 isoforms are comparable to mutant versions of OSBP, which lack all or part of the oxysterol-binding domain (Ridgway et al., 1992). The mutants were expressed in CHO or COS cells, and although oxysterols binding was not detected, they localized spontaneously to the Golgi apparatus. This was in contrast to the “wild-type” OSBP that translocated to the Golgi only when the oxysterol ligand binds. This study demonstrated that the oxysterol-binding site is in the COOHterminal half of the protein, and it was suggested that this area of the protein exerts a tonic inhibitory effect upon Golgi localisation (Ridgway et al., 1992). A more recent study utilised a yeast two-hydrid screen to identify a binding partner for OSBP; VAMP-associated protein-A (VAPA), a syntaxin-like protein implicated in ER/Golgi vesicle transport. The VAP-A binding region in OSBP was localized to the N-terminus of the OSBP domain and further mutant experiments indicated that intact oxysterol binding was not necessary for VAP-A binding. Both the full-length and truncated ORP3 splice variants contain this binding area. In vitro overexpression studies of the two ORP3 isoforms, (1a) and (2a), will determine if the truncated ORP3 behaves as a possible regulator of the “full-length” ORP3 by blocking Golgi translocation, and whether the splice variants behave in a similar way to the OSBP mutants. Other ORP splice variants have been described for ORP4 (ORP4-S) (Wang et al., 2002) and ORP1 (OSBPL1a) (Jaworski et al., 2001), but in these cases the PH domain was deleted and the major part of the OSBP region intact. As this is the opposite of the ORP3 isoform truncation, the actions of ORP4 cannot be compared with our predicted proteins. Our previous study confirmed the size of the “full-length” ORP3 by northern blot analysis (approx 7.0 kb) and demonstrated two other transcripts (approx 3.6 and 4.4 kb) (GregorioKing et al., 2001). The probe used in this study was specific for “standard exon 15” isoforms, and we are unable to ascertain whether these smaller transcripts relate to any of the isoforms ORP3 (1a–1d); which were amplified from the coding region only and therefore do not include the long 39 sequence. It is more likely, however, that the variations observed in the northern blot are a consequence of large variations in the noncoding region, as seen in an ORP3 clone from a uterine leiomyosarcoma (Genbank Accession No. BC017731), in which the ORP3 (1a) sequence is conserved but the 39 untranslated region is shortened by approximately 3.4 kb. The mRNA expression of each isoform varied considerably between tissues and even within specific compartments of the tissues. ORP3 isoforms (1a–1c) and (2a–2c) were highly expressed in whole brain tissue but negligible in the cerebellum. Interestingly, in a study characterizing ORP2, the reverse expression pattern was found, with the highest expression in the cerebellum and lower expression in the whole brain (Laitinen et al., 2002). Of the exon 9/12 variations, the exon 12 deletion was most common in the brain and an earlier cloning experi-

COLLIER ET AL. ment from hypothalamus confirmed this finding (data not shown). Interestingly, the expression of the extended exon 15 isoforms that results in a putative shortened protein was quite similar to the isoforms of the standard exon 15. The higher mRNA expression previously reported in CD341 hematopoietic cells compared with CD342 , was seen in the extended exon 15 isoforms but not in the standard exon 15 group. This may be due to the reduced sensitivity of semiquantitative RT-PCR compared with real-time PCR technology (which was utilized in the earlier study). In addition, the primers used in the realtime experiments were designed within exon 23 at the end of the coding region and would have amplified all the isoforms. Their combined expression gave an overall greater expression in CD341 hematopoietic cells, which is confirmed by this present study. It is difficult to compare previous ORP3 expression profiles as the results depend on the position of the primers or probes. In one ORP family study, the primers were designed out of the coding region, and therefore, would have amplified all the reported isoforms. The expression of OSBPL3 (5ORP3) was described as quite low, but appeared to be present in pineal and lung and ARPE-19 cells (retinal epithelium) (Jaworski et al., 2001). More comprehensive data from a multiple expression filter array (Lehto et al., 2001) found ORP3 expression highest in the kidney, small intestine, thymus, spleen, leukocytes, and lung. Unfortunately, the size and position of the ORP3 probe used in the hybridization was not made clear, but aside from the lung, the tissue distribution was similar to our results for ORP3 (1a–d). In conclusion, we have identified at least eight isoforms of ORP3 with specific gene and tissue distribution. The theoretical translation of these isoforms result in a number of protein sequences. Further studies are needed to investigate which of these isoforms are expressed as proteins and what role they play in the binding and regulation of the inhibitory effects of oxysterols, particularly in hematopoietic cells. Note added in proof: Since the submission of this paper, annotation of the database by the NCBI resulted in these ORP3 isoforms also being designated OSBPL3, Variants 1–8.

ACKNOWLEDGMENTS This work was supported in part by the Australian Red Cross Blood Service. We are also grateful to the medical and nursing staff of The Geelong Hospital and St. John of God, Geelong, Victoria, Australia, for collection of UCB samples.

REFERENCES BROWN, A.J. and JESSUP, W. (1999). Oxysterols and atherosclerosis. Atherosclerosis 142(1), 1–28. DAWSON, P.A., VAN DER WESTHUYZEN, D.R., GOLDSTEIN, J.L., and BROWN, M.S. (1989). Purification of oxysterol binding protein from hamster liver cytosol. J. Biol. Chem. 264(15), 9046–9052. FOURNIER, M.V., GUIMARAES, F.C., PASCHOAL, M.E., RONCO, L.V., CARVALHO, M.G., and PARDEE, A.B. (1999). Identification of a gene encoding a human oxysterol-binding protein-homo-

ORP3 ISOFORMS AND THEIR TISSUE DISTRIBUTION logue: A potential general molecular marker for blood dissemination of solid tumors. Cancer Res. 59(15), 3748–3753. GRAVELEY, B.R. (2001). Alternative splicing: Increasing diversity in the proteomic world. Trends Genet. 17(2), 100–107. GREGORIO-KING, C.C., COLLIER, G.R., MCMILLAN, J.S., WAUGH, C.M., MCLEOD, J.L., COLLIER, F.M., and KIRKLAND, M.A. (2001). ORP-3, a human oxysterol-binding protein gene differentially expressed in hematopoietic cells. Blood 98(7), 2279–2281. GREGORIO-KING, C.C., COLLIER, F.M., BOLTON, K.A., FERGUSON, M., HOSKING, J.B., COLLIER, G.R., and KIRKLAND, M.A. (2002). Effect of oxysterols on hematopoietic progenitor cells. Exp Hematol. 30(7), 670–678. JAWORSKI, C.J., MOREIRA, E., LI, A., LEE, R., and RODRIGUEZ, I.R. (2001). A family of 12 human genes containing oxysterol-binding domains. Genomics 78(3), 185–196. KANDUTSCH, A.A., and THOMPSON, E.B. (1980). Cytosolic proteins that bind oxygenated sterols. J. Biol. Chem. 255(22), 10813–10826. LAGACE, T.A., BYERS, D.M., COOK, H.W., and RIDGWAY, N.D. (1997). Altered regulation of cholesterol and cholesterol ester synthesis in Chinese hamster ovary cells overexpressing the oxysterol binding protein is dependant on the pleckstring homolgy domain. Biochem. J. 326(Pt 1), 205–213. LAITINEN, S., LEHTO, M., LEHTONEN, S., HYVARINEN, K., HEINO, S., LEHTONEN, E., EHNHOLM, C., IKONEN, E., and OLKKONEN, V.M. (2002). ORP2, a homolog of oxysterol binding protein, regulates cellular cholesterol metabolism. J. Lipid Res. 43(2), 245–255. LAITINEN, S., OLKKONEN, V.M., EHNHOLM, C., and IKONEN, E. (1999). Family of human oxysterol binding protein (OSBP) homologues: A novel member implicated in brain sterol metabolism. J. Lipid Res. 40, 2204–2211. LEHTO, M., LAITINEN, S., CHINETTI, G., JOHANSSON, M., EHNHOLM, C., STAELS, B., IKONEN, E., and OLKKONEN, V.M. (2001). The OSBP-related protein family in humans. J. Lipid Res. 42(8), 1203–1213. LEVINE, T.P., and MUNRO, S. (1998). The pleckstrin homology domain of oxysterol-binding protein recognises a determinant specific to Golgi membranes. Curr. Biol. 8(13), 729–739. MOREIRA, E.F., JAWORSKI, C., LI, A., and RODRIGUEZ, I.R. (2001). Molecular and biochemical characterization of a novel oxys-

9 terol-binding protein (OSBP2) highly expressed in retina. J. Biol Chem. 276(21), 18570–18578. RIDGWAY, N.D., DAWSON, P.A., HO, Y.K., BROWN, M.S., and GOLDSTEIN, J.L. (1992). Translocation of oxysterol binding protein to Golgi apparatus triggered by ligand binding. J. Cell Biol. 116(2), 307–319. SHAPIRO, M.B., and SENAPATHY, P. (1987). RNA splice junctions of different classes of eukaryotes: Sequence statistics and functional implications in gene expression. Nucleic Acids Res. 15(17), 7155–7174. SILVA, N.H., PIMENTA, G., PULCHERI, W.A., FOURNIER, M.V., SPECTOR, N., and DA COSTA CARVALHO, M.G. (2001). Detection of messenger RNA in leukocytes or plasma of patients with chronic myeloid leukemia. Oncol. Rep. 8(3), 693–696. WANG, C., JEBAILEY, L., and RIDGWAY, N.D. (2002). Oxysterolbinding-protein (OSBP)-related protein 4 binds 25-hydroxycholesterol and interacts with vimentin intermediate filaments. Biochem. J. 361(Pt 3), 461–472. WYLES, J.P., MCMASTER, C.R., and RIDGWAY, N.D. (2002). VAMP-associated protein-A (VAP-A) interacts with the oxysterol binding protein (OSBP) to modify export from the endoplasmic reticulum. J. Biol Chem. 277, 29908–29918. XU, Y., LIU, Y., RIDGWAY, N.D., and MCMASTER, C.R. (2001). Novel members of the human oxysterol-binding protein family bind phospholipids and regulate vesicle transport. J. Biol. Chem. 276(21), 18407–18414.

Address reprint requests to: Fiona Collier Stem Cell Laboratory Douglas Hocking Research Institute Barwon Health The Geelong Hospital Geelong, Victoria, 3220 E-mail: [email protected] Received for publication August 15, 2002; accepted September 18, 2002.

This article has been cited by: 1. Terhi Vihervaara, Maurice Jansen, Riikka-Liisa Uronen, Yuki Ohsaki, Elina Ikonen, Vesa M. Olkkonen. 2011. Cytoplasmic oxysterol-binding proteins: sterol sensors or transporters?. Chemistry and Physics of Lipids . [CrossRef] 2. Mike H. Ngo, Terry R. Colbourne, Neale D. Ridgway. 2010. Functional implications of sterol transport by the oxysterol-binding protein gene family. Biochemical Journal 429:1, 13-24. [CrossRef] 3. Daoguang Yan, Vesa M. OlkkonenCharacteristics of Oxysterol Binding Proteins 265, 253-285. [CrossRef] 4. Daoguang Yan, Vesa M Olkkonen. 2007. OSBP-related proteins: lipid sensors or transporters?. Future Lipidology 2:1, 85-94. [CrossRef] 5. Markku Lehto, Riikka Hynynen, Katja Karjalainen, Esa Kuismanen, Kati Hyvärinen, Vesa M. Olkkonen. 2005. Targeting of OSBP-related protein 3 (ORP3) to endoplasmic reticulum and plasma membrane is controlled by multiple determinants. Experimental Cell Research 310:2, 445-462. [CrossRef] 6. Jessica P Wyles, Neale D Ridgway. 2004. VAMP-associated protein-A regulates partitioning of oxysterolbinding protein-related protein-9 between the endoplasmic reticulum and Golgi apparatus. Experimental Cell Research 297:2, 533-547. [CrossRef] 7. Vesa M Olkkonen. 2004. Oxysterol binding protein and its homologues: new regulatory factors involved in lipid metabolism. Current Opinion in Lipidology 15:3, 321-327. [CrossRef] 8. Vesa M Olkkonen, Timothy P Levine. 2004. Oxysterol binding proteins: in more than one place at one time?. Biochemistry and Cell Biology 82:1, 87-98. [CrossRef]