Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2007 Mar 7.
Published in final edited form as: J Lipid Res. 2002 Apr;43(4):565–578.

Molecular cloning, genomic organization, genetic variations, and characterization of murine sterolin genes Abcg5 and Abcg8

Kangmo Lu *, Mi-Hye Lee *, Hongwei Yu *, Yuehua Zhou *, Shelley A Sandell *, Gerald Salen , Shailendra B Patel *,1
PMCID: PMC1815568  NIHMSID: NIHMS4356  PMID: 11907139

Abstract

Mammalian physiological processes can distinguish between dietary cholesterol and non-cholesterol, retaining very little of the non-cholesterol in their bodies. We have recently identified two genes, ABCG5 and ABCG8, encoding sterolin-1 and -2 respectively, mutations of which cause the human disease sitosterolemia. We report here the mouse cDNAs and genomic organization of Abcg5 and Abcg8. Both genes are arranged in an unusual head-to-head configuration, and only 140 bases separate their two respective start-transcription sites. A single TATA motif was identified, with no canonical CCAT box present between the two genes. The genes are located on mouse chromosome 17 and this complex spans no more than 40 kb. Expression of both genes is confined to the liver and intestine. For both genes, two different sizes of transcripts were identified which differ in the lengths of their 3′ UTRs. Additionally, alternatively spliced forms for Abcg8 were identified, resulting from a CAG repeat at the intron 1 splice-acceptor site, causing a deletion of a glutamine. We screened 20 different mouse strains for polymorphic variants. Although a large number of polymorphic variants were identified, strains reported to show significant differences in cholesterol absorption rates did not show significant genomic variations in Abcg5 or Abcg8.—Lu, K., M-H. Lee, H. Yu, Y. Zhou, S. A. Sandell, G. Salen, and S. B. Patel. Molecular cloning, genomic organization, genetic variations, and characterization of murine sterolin genes Abcg5 and Abcg8.

Keywords: dietary cholesterol, sitosterolemia, genetics, sterol transporter, ATP-binding cassette, inbred mouse strains

Abbreviations: EST, expressed sequence tag; IMAGE, integrated molecular analysis of genomes and their expression; LXR, liver X receptor; QTL, quantitative trait loci; SNP, single nucleotide change


Sitosterolemia (also known as phytosterolemia, MIM 210250) is a rare autosomal recessively inherited metabolic disorder, that was first discovered in 1974 in two sisters (1). Sitosterolemia patients develop tendon and tuberous xanthomas, hemolytic episodes, arthralgias, and arthritis, and premature coronary and aortic atherosclerosis leading to cardiac fatalities (24). Affected individuals have very high levels of plasma plant sterols (sitosterol, campesterol, stigmasterol, avenosterol, and others) and 5∝-saturated stanols, particularly sitostanol, but their blood cholesterol levels may be normal or only moderately increased. Affected individuals show not only very high levels of dietary cholesterol absorption, they absorb all types of sterols and retain them in the body. Thus in contrast to normals, affected individuals absorb and retain plant sterols (2,3), as well as shellfish sterols (5). Clinical studies further show that affected individuals have an inability to excrete sterols, both plant sterols as well as cholesterol, into bile (6,7). Increased intestinal absorption, decreased hepatic excretion of sitosterol (the major plant sterol), and abnormally low cholesterol biosynthesis are reported to be features of sitosterolemia (59).

Previous studies on sitosterolemia have suggested that the defect in sitosterolemia may involve a putative sterol transporter expressed in the intestine and/or the liver (10,11). Using linkage analyses, the STSL locus has been mapped to human chromosome 2p21 and no evidence of genetic heterogeneity was detected in a large cohort of sitosterolemia pedigrees (11,12). By using positional cloning procedures (1113), we isolated two novel genes from STSL locus, chromosome 2p21, between D2S2294 and D2S2298. Both genes encode for proteins that contain an ATP binding signature sequence, characteristic of the ABC family of proteins, at the N-terminal and a transmembrane domain located at the carboxyl terminus consisting of 6 membrane-spanning helices. These findings have also been reported independently by Berge et al. (14,15). These genes, using the Human Genome Organization nomenclature, have been named ABCG5, encoding sterolin-1, and ABCG8, encoding sterolin-2, respectively. Mutations in ABCG5 and ABCG8 genes are responsible for causing sitosterolemia (1416). Based upon the clinical defects in sitosterolemia, these proteins are predicted to play a pivotal role in the selective absorption of cholesterol from the intestinal lumen and in the selective excretion of non-cholesterol sterols into bile (4,16,17). We report here (1) the murine genomic organization, cDNA, and tissue expression patterns of these genes, (2) an unusual, but naturally occurring alternative splicing of a mouse ABCG8 cDNA resulting in an in-frame deletion of a single amino acid, and (3) polymorphic variations in Abcg5 and Abcg8 in 20 inbred mouse strains.

MATERIALS AND METHODS

Database searches

Human ABCG5 and ABCG8 cDNA sequence information was used to search the databases for any murine homologous sequences using the Basic Local Alignment Search Tool (18, 19). Three murine expressed sequence tag (EST) integrated molecular analysis of genomes and their expression (IMAGE) clones (20) containing sequences homologous to human ABCG5 cDNA and three clones homologous to human ABCG8 cDNA were identified. These clones were obtained from Research Genetics (Huntsville, AL) and fully sequenced. IMAGE clones 1885393 and 1885458 contained Abcg8 coding sequences and were used for further analyses.

Northern-blot analysis

A mouse multiple-tissue poly(A)+ RNA Northern blot (Origene, Bethesda, MD) was hybridized with a radio-labeled full-length mouse Abcg5 or Abcg8 cDNA as previously described (14). The hybridized filter was washed stringently with 0.1 × SSC-0.1% SDS at 68°C, exposed to a phosphorimager cassette, striped, and re-probed with a mouse β-actin probed for comparison of RNA loading.

cDNA library screening

A mouse cDNA lambda phage library (Stratagene, La Jolla CA) was screened with a partial length Abcg8 cDNA probe to identify full-length cDNA clones (21). Screening an estimated 3 × 106 for Abcg8, we identified more than 80 positive clones, 69 of which were plaque purified and characterized further. A library screen for Abcg5 cDNAs was not performed (see Results). All purified positive clones were excised in vivo into insert-containing pBSK phagemid vector by infection with ExAssist helper phage followed by transduction of filamentous phage particles into E. coli XLOLP according to the manufacturer’s protocols. The phagemid DNA was prepared by plasmid miniprep Kit (Roche, Indianapolis, IN) and cycle-sequenced using BigDye Terminator Cycle Sequencing Ready Reaction Kit and an ABI 373 DNA genetic analyzer.

Reverse transcriptase and rapid amplification of cDNA ends-PCR

Total RNA was isolated from freshly isolated mouse liver using TRIzol reagent (Life Technologies, Gaithersburg, MD). Reverse transcription of 10 μg of total RNA was performed using oligonucleotide d(T) and the single-strand cDNA used as template to amplify mouse Abcg5 cDNA using primers mg5-F4 and mg5-R3 (Table 1). Rapid amplification of cDNA ends (RACE) was performed using a commercially available SMART RACE cDNA Amplification Kit (Clontech, Palo Alto, CA) according to the manufacturer’s protocol, using total liver or intestinal RNA. Five prime and three prime RACE cDNA was synthesized using primers supplied with the kit. PCR products were directly sequenced or sub-cloned using TA cloning and isolated clones sequenced as described above.

TABLE 1.

Oligonucleotide primers used for amplification of Abcg5 and Abcg8

Abcg5
Sense Primers
Antisense Primers
Primer Name Position in cDNA Sequences 5′ to 3′ Primer Name Position in cDNA Sequences 5′ to 3′
mg5-F1 142-162 GGTGAGCTGCCCTTTCTGAGT mg5-2R 363-343 CAAGGAGACATCTTTGAGGAT
mg5-F4 343-363 ATCCTCAAAGATGTCTCCTTG mg5-3R 537-518 GACGTAGGAGAAGCAGTCTT
mg5-3F 469-489 GAAGGGGAGGTGTTTGTGAA mg5-4R 608-589 AGGGCCAGCATCGCTGTGTA
mg5-4F 553-573 TTTCTGAGCAGCCTCACTGTG mg5-5R 775-757 TGGGGTCCTGAAGGAGTTG
mg5-5F 643-662 GTAGAGGCAGTCATGACAGA mg5-6R 915-896 TTGGAAGAGCTCAGAGCGAG
mg5-6F 776-796 AGGTCATGATGCTAGATGAGC mg5-7R 1036-1016 CAAAGGGATTGGAATGTTCAG
mg5-7F 919-939 TTCGACAAAATTGCCATCCTG mg5-8R 1250-1231 ACACCAAGCTTGCCGAACAT
mg5-8F 1050-1069 CTTGACATCAGTGGACACCC mg5-9R 1462-1442 GATTCACAGCATTGAGCATGC
mg5-9F 1260-1282 GCGAGTAACAAGAAACTTAATGA mg5-10R 1597-1577 ACACACTGCTGAAAATGACCG
mg5-10F 1466-1485 TTCCCATGCTGAGAGCCGTC mg5-11R 1781-1762 CCAGATCCAATAAGCAGCCC
mg5-11F 1605-1624 GACTCTGGGCTTGTATCCTG mg5-12R 1776-1756 CATTGACCACGAGAATCTCAC
mg5-12F 1756-1776 GTGAGATTCTCGTGGTCAATG mg5-13R 2160-2140 TACTTCTCTGTGCTCCACAGT
mg5-13F 1968-1988 GAGAAAACCTGCCCAGGTGC mg5-R3 1806-1486 TCCTGACTCTCCTGGTCGCT
Abcg8
Sense Primers
Antisense Primers
Primer Name Position in cDNA Sequences 5′ to 3′ Primer Name Position in cDNA Sequences 5′ to 3′
mg8-F1 42-62 AAAGACAGAGAGAGCCCAACA mg8-2R 264-244 AGGTGAGATCTCTGACCTCC
mg8-F3 188-208 CTCCTCGGAAAGTGACAACAG mg8-R11 417-397 CCCTATGATGGCCAGCATCT
mg8-3F 285-305 CAGGTGCCTTGGTTTGAGCAG mg8-R10 544-525 CACTTCCTCACCAGCTGAGG
mg8-F2 432-451 GGGAGAGCCTCACTACTCGA mg8-4R 742-723 ACCCCACGTACATACGTGTT
mg8-4F 723-742 AACACGTATGTACGTGGGGT mg8-R7 939-920 GCCTGAAGATGTCAGAGCGA
mg8-F7 920-939 TCGCTCTGACATCTTCAGGC mg8-7R 1223-1203 GTGGGTGCTTGTGTTGAGTTC
mg8-F4 1075-1094 TGACCAGCATCGACAGACGC mg8-8R 1274-1254 CTCAACAGCAGTCCCACAGTC
mg8-8F 1254-1274 GACTGTGGGACTGCTGTTGAG mg8-9R 1505-1485 GACGACATCCAGGATGACATT
mg8-9F 1320-1340 ATTTCCAATGACTTCCGGGAC mg8-10R 1565-1545 AGTGTACAGCCCGTCTTCCAG
mg8-10F 1545-1565 CTGGAAGACGGGCTGTACACT mg8-R12 1807-1788 GAGTTGTAGAGGGCATTGCA
mg8-F12 1691-1709 ACACTTCCTGCTCGTGTGG mg8-R13 1973-1953 GATGGAGAAGGTGAAGTTGCC
mg8-F13 1953-1973 GGCAACTTCACCTTCTCCATC mg8-R2 2230-2211 GGTGGCTGCTGCCTGAGAGA

BAC library screening

The sequences of Abcg5 and Abcg8 cDNA were used to design primer sets (Table 1) and used to screen a Mouse BAC library CitbCJ7 (Research Genetics) by PCR of DNA superpools and plates, using a modified protocol as previously described (13). The primers used for screening for Abcg5 were mg5-9F, 9R, mg5-13F, and mg5-13R. Primers mg8-1F, mg8-1R, mg8-13F, and mg8-13R were used for Abcg8 screening. Five μl of pooled DNA from the BAC library was used in a 10 μl PCR reaction containing 1.5 mM MgCl2, 16 mM (NH4) 2SO4, 0.1 mM dNTPs, 67 mM Tris (pH 8.8), 1 μM primer, and 1 unit of Taq polymerase with the following protocol: 94°C for 30 s, 60°C for 30 s, 72°C for 1 min for 35 cycles. Positive mouse BAC clones were obtained from Research Genetics, Inc., plated on agar plates containing 12.5 μg/ml chloramphenicol, and confirmed by colony PCR using the above primer sets.

Direct BAC Sequencing and exon/intron boundary

BAC plasmid DNA was prepared and direct BAC sequencing was performed as previously described (13).

In addition to direct sequencing of the BAC, amplification of genomic fragments, using primers selected from the cDNA sequences (Table 1), was used to determine exon/intron boundary sequences.

Analysis of alternative splice forms

To quantitate the levels of the two alternatively spliced forms of Abcg8 RNAs, we utilized the loss of EcoO109 I restriction enzyme recognition site, as a result of the loss of the CAG triplet codon from the cDNA. Fragments containing the splice difference were amplified by RT-PCR from cDNA using mouse liver, jejunum, ileum, or colon total RNA using primers mg8-1F and mg8-R11, to yield an expected PCR product of 322–325 bp (Table 1). PCR reactions were spiked with 32P end-labeled primer mg8-R11 at the ratio of 1:10 unlabeled primer for detection of PCR products. PCR products were digested with EcoO109 I and separated by 5% acrylamide gel electrophoresis under non-denaturing conditions. The gels were dried and exposed to a phosphorimager cassette (Molecular Dynamics, CA) and analyzed using ImageQuant v1.11 software. Although a 322–325 bp sized PCR products were expected, two additional bands migrating slightly slower were consistently identified following agarose or acrylamide gel electrophoresis. These DNA bands were purified and directly sequenced using the amplification primers and were found to represent heteroduplexes between the spliced two forms (see Results).

Chromosomal localization

To localize Abcg5 and Abcg8 on the mouse chromosomes, we performed electronic PCR. We initially identified the corresponding mouse chromosome syntenic to human chromosome 2p21 (http://www.ncbi.nlm.nih.gov/). We selected markers spanning the STSL locus whose murine counterparts were also known and had been mapped (http://www.informatics.jax.org/). To confirm the mapping, we identified mapped mouse EST sequences (http://www.genome.wi.mit.edu/mouse_rh/index.html) that shared sequence identity with Abcg5 and Abcg8 cDNA (see Results).

Animals

Various mouse inbred strain genomic DNAs were obtained from the DNA resource, Jackson Laboratory, Bar Harbor, ME (http://www.jax.org/resources/documents/dnares/index.html) and polymorphic variants detected by direct PCR amplification and sequencing. Mouse strains were chosen based up the reported plasma cholesterol levels (22) (http://aretha.jax.org/pub-cgi/phenome/mpdcgi?rtn=docs/home), as well from studies reported to show variations in dietary cholesterol absorption (2326). For plasma sitosterol levels, two females of each selected strain were housed for 1 week, fed standard rodent chow, and fasted 2 h before sacrifice for blood and tissue sampling. Plasma sitosterol and campesterol levels were determined as previously described (27).

RESULTS

Identification of mouse Abcg5 and Abcg8 cDNA

By searching the mouse EST database with the human ABCG5 cDNA sequence, we identified three mouse ESTs (GenBank Accession Nos. AI505249, AI507053, and AA237916). Sequence analyses of these clones indicated that they contained partial Abcg5 cDNA sequences. To obtain full-length cDNA sequence, we used a combination of RT-PCR and 5′ RACE. The complete sequence analysis of mouse Abcg5 cDNA demonstrated that it encoded an open reading frame of 652 amino acids with a calculated molecular mass of 75 kDa (Fig. 1). The deduced amino acid sequence of mouse Abcg5 showed a high degree of conservation, with a 92.8% and 80.1% homology match with rat and human ABCG5 proteins, respectively. Mouse Abcg5 has an extra amino acid, R35, compared human ABCG5. A poly(A)+ site was not identified in the 3′ UTR and 3′ RACE failed to extend the known 3′ end for this cDNA.

Fig. 1.

Fig. 1

Polypeptide sequence and secondary features of mouse Abcg5 and Abcg8 cDNAs. The sequence and secondary features for Abcg5 (A) (nucleotide sequence Genbank No. AF312713) and Abcg8 (B) (nucleotide sequence Genbank No. AF324495) are as shown. The positions of the Walker A motif (inverted triangles), B motif (alternating triangles), and C motif (upright triangles) are as indicated. The predicted transmembrane domains are underlined. Shaded boxes indicate complete homology, with conserved sequence changes indicated by the open boxes. The open circles indicate the positions of coding polymorphic variants detected in different mouse strains (see also Table 3).

Using human ABCG8 cDNA sequence information, we identified three homologous mouse cDNA sequences in the databases (IMAGE clones 1925061,1885393, and 1885458). Clone 1925061 contained an internal deletion and was not analyzed further. The remaining two clones were identical and contained an insert of 2.6 kb. Sequence analyses showed that these clones contained the full-length mouse Abcg8 cDNA with a poly(A)+ tail, located 13 bp down-stream of an AACAAA sequence motif. Two other poly(A) signals (AAUAAA) at 189 and 224 bp up-stream of the poly(A)+ tail were also present. Translation of the longest open reading frame revealed a protein of 673 amino acids, with calculated molecular mass of 76 kDa (Fig. 1). The predicted amino acid sequence of mouse Abcg8 showed a high degree of conservation, with 90.6% and 81.5% homology to rat and human ABCG8, respectively (14,16).

Mouse Abcg5 and Abcg8 show a 30.8% homology to each other at the peptide level. Both proteins have highly conserved ATP-binding cassette signature motifs (Walker A, B, and C) (Fig. 1) located at the N-terminal half, and a predicted 6-transmembrane domain located at the C-terminal end (underlined, Fig. 1).

Tissue distribution of mouse Abcg5 and Abcg8 mRNA

Expression patterns of mouse Abcg5 and Abcg8 genes were examined by Northern blot and RT-PCR methods (Fig. 2). Expression of Abcg5 (14) and Abcg8 appear to be restricted to the liver and the small intestine, with a low level of expression in the colon. Size heterogeneity for both Abcg5 mRNA (2.3 kb and 3.3 kb) (14) and Abcg8 mRNA (2.6 kb and 3.7 kb) (Fig. 2A) was observed. In order to examine the nature of the size difference of the two transcripts for Abcg8, we screened a mouse liver phage cDNA library using Abcg8 as probe. Two major lengths of cDNAs, 2.56 and 3.67 kb respectively, were identified. Sequence analyses showed the size differences arose from 3′ UTR variation (data not shown. See Genbank submission AF324494). Although a library screening for Abcg5 cDNA was not performed, we predict that a similar variation on the 3′ UTR may account for the larger sized message.

Fig. 2.

Fig. 2

Expression patterns of Abcg5 and Abcg8 genes. Northern blot analyses of poly(A)+ RNA from a variety of tissues (see Materials and Methods) showed that expression of Abcg8 was confined to the liver and small intestine (A). Data for mouse Abcg5 mRNA expression have been published previously (14). RT-PCR analyses confirmed the expression patterns and further showed that expression was predominantly limited to the small intestine and liver. As a control, β-actin was co-amplified as an internal control and the last tracks indicate no template (water-only) controls.

RT-PCR analyses of RNA from different mouse tissues confirmed that the expression of these genes was confined to the liver, jejunum, and ileum, with weak expression in the colon (Fig. 2B).

Gene organization, exon-intron boundary, and promoter analyses of Abcg5 and Abcg8 genes

In order to obtain genomic information for both genes, we screened a mouse BAC library (CitbCJ7) using primer sets designed from the first and last exon sequences of Abcg5 and Abcg8. Three BAC clones, 329B11, 329A22, and 376H16, were identified, all of which contained the full-length genomic fragments for both Abcg5 and Abcg8. Exon-intron boundaries were determined by direct sequencing of the BAC DNA and/or long PCR amplified products using exon specific primers. The two genes are organized in a head-to-head tandem configuration and each gene contains 13 exons and 12 introns (Fig. 3A and Table 2). Abcg5 spans about 23 kb of genomic DNA and Abcg8 spans ~16 kb of genomic DNA. The exon sizes range from 98 to 206 bp and intron sizes from 87 bp and to about 5 kb. All exon-intron boundaries show canonical sequences with initial GT at the splice-donor and terminal AG at splice-acceptor sites.

Fig. 3.

Fig. 3

Gene organization and promoter sequence comparison of Abcg5 and Abcg8. The gene organization of Abcg5 and Abcg8 is shown (A). The two genes span an approximate distance of 37 kb. The exact size of exon 13 for Abcg5 has not been determined. B: Shows the nucleotide sequence homology between mouse and human promoter region separating the two ATG initiation codons of Abcg5 and Abcg8. A number of transcriptional motifs were identified in the murine sequence by computer-based analyses, indicated by the arrows, but only two of these, GATA1 and GATA2, are conserved between mouse and human. The human sequence shows a remarkable paucity for transcriptional factor motifs (14). The start transcription sites (indicated by the thick arrows) for mouse Abcg8 is located at −133. For Abcg5, the precise start transcription site has not been precisely defined, but may lie in the region defined by the thick shaded line, based upon oligonucleotide-based PCR amplification of cDNA (see Materials and Methods).

TABLE 2.

Exon-intron boundaries and organization of Mouse Abcg5 and Abcg8

Abcg5
Exon
Intron
Exon
GenBank Accession No. No. Size(bp) 3′ End Sequence Splice Donor No. Size(bp) Splice Acceptor 5′ End Sequence No.
AF351786 1 146 GTCCGACAGCGTCAG gtaaggggacc 1 596 tttcctttaaag CAACCGTGTCGGGCC 2
AF351787 2 122 TCTTAGGCAGCTCAG gtaagtgcctgg 2 ~5000 gtcgccccctag GCTCAGGGAAGACCA 3
AF351788 3 136 TCCTACGTCCTGCAG gtgggcgtgtcc 3 87 tctggcccctag AGCGACGTTTTTCTG 4
AF351789 4 98 TTCTACAACAAGAAG gtacttgaaagtt 4 ~2000 gtgtctcttacag GTAGAGGCAGTCATG 5
AF351790 5 133 TCCTTCAGGACCCCA gtaagtgggaca 5 1228 ctttgccggcag AGGTCATGATGCTAG 6
AF351791 6 140 TCTGAGCTCTTCCAA gtaagggaatgc 6 900 ttgtccaagcag CACTTCGACAAAATT 7
AF351792 7 129 CCTTTGATTTTTACA gtaagtgtattct 7 688 gaaaacttttag TGGACTTGACATCAG 8
AF351793 8 214 TGGTGTCCTGCTGAG gtaagagccttg 8 99 tttggttttcag GCGAGTAACAAGAAA 9
AF351794 9 206 ATGCTGTGAATCTGT gtaagtgcctgt 9 836 cttctatgccag TTCCCATGCTGAGAG 10
AF351795 10 139 CAGTGTGTGTTATTG gtaaggcggtgt 10 3000 atgtttttctag GACTCTGGGCTTGTA 11
AF351796 11 186 ATCTGGATTTATCAG gtaagaagaaa 11 ~4500 tttttcttaag AAACATACAAGAGAT 12
AF351797 12 113 TGAACTTCACTTGTG gtaagtgttctat 12 1495 ttttccttgcag GTGGATCCAACACCT 13
AF351798 13 258
Abcg8
Exon
Intron
Exon
GenBank Accession No. No. Size(bp) 3′ End Sequence Splice Donor No. Size(bp) Splice Acceptor 5′ End Sequence No.
AF351799 1 257 CTTCAGGATGCTTCG gtgagtgagctctgc 1 ~2500 ctacatgtctcccag CAGGGCCTCCAGGAC 2
AF351800 2 105 GATCTCACCTACCAG gtaggggcacgtgc 2 1818 acctctccccacag GTGGACATCGCCTCT 3
AF351801 3 157 TCATAGGGAGCTCAG gtaccaacagaggct 3 3102 tctctggatttgcag GCTGCGGGAGAGCCT 4
AF351802 4 239 CAGCGTGACAAACGG gtaacagttggctgg 4 442 tccgcctgtcctcag GTGGAAGACGTAATC 5
AF351803 5 133 TCCTGTGGAACCCAG gtgaggcctgggaa 5 94 gtgcaatatccccag GAATCCTCATTCTGG 6
AF351804 6 270 CTGCGGACTTCTACG gtgagtggtaaaggc 6 1995 atcttctgcttacag TGGACTTGACCAGCA 7
AF351805 7 163 CACCCACACAGTCAG gtacgggaagcccg 7 87 actgctcccaacag CCTGACCCTCACACA 8
AF351806 8 81 TTCCACCCTGATCCG gtaaatctccctccc 8 600 gaggctttctttcag TCGTCAGATTTCCAA 9
AF351807 9 201 ATGTCGTCTCCAAAT gtgagtgtcacctgc 9 522 ctcccccatctttag GTCACTCGGAGAGGT 10
AF351808 10 77 TATTTCTTTGCCAAG gtcagggctgggag 10 545 agctgtgctttgcag ATCCTAGGAGAATTG 11
AF351809 11 268 ACAACCTGTGGATAG gtgaggcctgctgcc 11 887 tcttgctgtcttcag TGCCTGCATGGATCT 12
AF351810 12 128 ATCCTCGGAGACACG gtacgtagcgaagg 12 84 aatgtctgtccgcag ATGATCAGTGCCATG 13
AF351811 13 605

By chromosome primer walking using reverse oligonucleotide primers located in exons 1 of mouse Abcg5 and Abcg8, we discovered that both genes were in a head-to-head configuration, with a distance of 358 bp separating their respective initiation codons (Fig. 3B). Analysis of this region using TESS software (http://www.cbil.upenn.edu/tess/index.html) revealed a number of potential binding sites for transcription factors (Fig. 3B), but these were restricted to the mouse sequence only, the human sequence showing variances at almost all of these potential sites. A single TATA box was identified 232 bp up-stream of the mouse Abcg5 ATG initiation codon, as well as two GATA motifs (Fig. 3B).

This putative promoter region sequence has a 63% homology to the corresponding human sequence (Fig. 3B), though the human sequence does not contain the TATA motif. Preliminary results, using a minimal murine 358 bp fragment located between the two ATGs of Abcg5 and Abcg8 did not result in expression of a reporter gene when transfected into HEK 293 cells in either orientation (data not shown). Thus this sequence does not appear to function as a ‘minimal’ promoter element.

Chromosome localization

The human STSL locus, containing the ABCG5 and ABCG8 genes, maps to chromosome 2p21 between micro-satellite markers D2S2298 and D2S2294 (11,12). Human chromosome 2p21 is syntenic with mouse chromosome 17 (http://www.ncbi.nlm.nih.gov/Homology/). However, there are small regions within human 2p21-2p16 region that share synteny with mouse chromosome 11. To localize mouse Abcg5 and Abcg8, we searched the Mouse Genome Informatics (http://www.informatics.jax.org/) and Whitehead Institute Center for Genome Research Mouse RH Map (http://www.genome.wi.mit.edu/mouse_rh/index.html) databases and performed electronic PCR using mouse Abcg5 and Abcg8 genomic sequences. Two murine sequences were identified (GenBank No. AI114946 and AW112016), that share complete sequence identity with exon 13 of mouse Abcg8 and exon 10–13 of mouse Abcg5 respectively. Both of these murine sequences have been mapped between chromosomal markers D17Mit41 and D17Mit189 (Fig. 4). Thus, based upon both synteny, as well as identification of Abcg5 and Abcg8 sequences, these genes map to between markers D17Mit41 and D17Mit75 at about 53 cM on the genetic map.

Fig. 4.

Fig. 4

Chromosomal localization of Abcg5 and Abcg8. Based upon both synteny, as well as identification of expressed sequence tag (EST)s identical to Abcg5 or Abcg8, these genes can be firmly placed on mouse chromosome 17, at 53 cM between microsatellite markers D17mit41 and D17mit189.

Alternative splicing for Abcg8 cDNA

Compared with the human ABCG8 polypeptide, the mouse Abcg8 polypeptide sequence has one more glutamine located at the beginning of exon 2 (16). By sequencing genomic BAC DNA and RT-PCR products using mouse liver RNA, we identified two alternatively spliced forms of mRNA (Fig. 5A). The 5′-end of intron 1 contains a motif (GTGAGT), which conforms to the consensus splice donor site (GURAGU). However, at the 3′-end of intron 1 there is a tandem CAG repeat (Fig. 5). This tandem CAG repeat at the splice acceptor site can result in a choice of splice acceptor sites; either the proximal or the distal AG can be utilized. Mouse liver cDNA library screening resulted in two sets of clones differing in the presence or absence of a CAG triplet codon encoding a glutamine residue at this position (nucleotides 165–167, Genbank No. AF324495). Since the presence or absence of this CAG alters the nucleotide recognition site in the cDNA for the restriction endonuclease EcoO109I, we used this to quantitate the prevalence and tissue expression patterns of the two different isoforms. RT-PCR was performed using 32P-labeled primer and total RNA from mouse liver jejunum, ileum, or colon (Fig. 5B). The PCR products were digested by EcoO109I and separated by non-denaturing acrylamide gel electrophoresis. Undigested PCR products showed the presence of three bands, two of which migrated slower than the expected size. Upon digestion, the expected size bands were diminished in amounts, though the slower migrating bands were unchanged (digestion results in a 50 bp fragment size that is not shown, Fig. 5B). The two slower migrating bands were shown to be heterodimers of the two forms of spliced products produced during the PCR reaction, and were verified by direct sequencing. These heterodimers would be expected to be resistant to endonuclease digestion. Quantitative analysis of bands (Fig. 5B) showed that these two spliced forms were present in almost equal abundance in mouse liver, jejunum, or colon; 41% of the cDNAs contained the CAG and 59% had the CAG deleted. This is comparable to the equal numbers of such clones identified in the liver cDNA library.

Fig. 5.

Fig. 5

Alternative splicing for Abcg8 results from a CAG repeat at splice-acceptor boundary region in intron 1. Two sets of cDNAs were identified from the library screening that differed in the presence or absence of a CAG triplet codon at the boundary of exon 1 and exon 2 in the cDNA (A, electropherograms). Examination of the exon-intron boundary suggested that two possible choices for splice-acceptor sites existed (B, underlined). To confirm that alternatively spliced forms of Abcg8 were present in RNA isolated from normal mouse liver, jejunum, ileum, or colon, we exploited the loss of EcoO19I recognition sequence, if the CAG was deleted. Labeled PCR products (see Materials and Methods) were digested and separated by acrylamide gel electrophoresis (C). Although a PCR product size of 325 or 322 bp was expected, two other slower bands were identified, resulting from heteroduplexes formation (confirmed by direct sequencing of these bands). About 40% of the mRNA utilized the first CAG for splice-site identification and 60% utilized the next CAG (underlined). No significant differences between the various tissues in RNA alternative splice forms was detected.

Genetic variations in inbred mouse strains

Inbred mouse strains have been used to identify genes whose genetic variations may be important determinants of atherosclerosis, gall stone formation, or biliary cholesterol secretion (2832). Some of these inbred mouse strains have been screened for differences in dietary cholesterol absorption (2326). To identify whether genetic variations in Abcg5 and Abcg8 may be responsible for some of these phenotypes, we screened 20 strains. These strains were selected based upon documentation of either cholesterol absorption rates, or having very high levels of plasma cholesterol levels. The latter phenotype was chosen because some sitosterolemia patients presented with elevated levels of plasma cholesterol and were initially diagnosed as pseudohomozygous familial hypercholesterolemia. Table 3 shows a compilation of both coding and non-coding alterations detected for Abcg5 and Abcg8. There were 21 polymorphisms (14 in Abcg5 and 7 in Abcg8) that altered amino acid coding, 77 intronic nucleotide changes, 7 single nucleotide changes (SNPs) in the 5′UTR, 21 SNPs in the 3′UTR, and 50 single nucleotide changes in exonic regions that did not alter amino acid coding. All of these changes were present as homozygous changes, compatible with the inbreeding of these lines. The greatest variations were noted in DBA/1J and DBA/2J (both identical to each other), with CAST/Ei, SPRET/Ei, MOLF/Ei, RBF/DnJ, CBA/J, CE/J, and FVB/NJ all containing some genetic variations (Table 3). Remarkably, 129/SvEv, C57BL/6J, SM/J, BALB/cJ, C3H/HeJ, AKR/J, and C57L/J were all essentially identical at these loci.

TABLE 3.

Polymorphisms identified Abcg5 and Abcg8 in 17 mouse strains

Gene Position Nucleotide changea Effectb DBA/1J Cast/Ei MOLF/Ei RBF/DnJ SPRET/Ei CBA/J CE/J FVB/NJ RIIIS/J C58/J 129/SvFv ARK/J C57L/J C3H/HeJ C57BL/6Jc
Abcg5 5′-UTR −39C>T
exon 1 230C>T T31M
237A>G
intron 2 IVS2−8C>T
IVS2−49A>G
IVS2−50A>G
IVS2−101T>C
exon 3 454C>T R106C
477G>C E113D
489T>C
exon 4 624G>A
intron 4 IVS4−35C>T
IVS4−48G>A
exon 5 684G>A
691A>G M185V
intron 5 IVS5+19G>A
IVS5+35A>G
IVS5−24T>C
intron 6 IVS6+19G>A
IVS6+35A>C
IVS6−29C>A
IVS6−37T>G
IVS6−44C>T
exon 7 987C>A
1041T>C
intron 7 IVS+12C>T
IVS7+34G>A
IVS7+44C>T
IVS7+66 67delT
IVS7−45 56delTG
IVS7−50G>A
IVS7−88C>A
exon 8 1095G>A
1121G>C C328S
1199C>T T354M
1164C>T
1209C>T
1224T>G
1236C>T
intron 8 IVS816A>C
IVS8+19C>T
IVS8+23C>T
IVS8+86C>G
IVS8−16C>G
exon 9 1359C>A
1374G>A
1383C>T
intron 9 IVS9−13G>A
exon 10 1553T>C V472A
1567G>A V477I
1581C>G
intron 10 IVS10−77A>G
exon 11 1701G>A
1764G>A
intron 11 IVS11+43A>G
IVS11+45G>A
IVS11−8 15delTTTTCTCT
IVS11−73 74delCT
IVS11−149T>C
IVS11−165C>T
IVS11−179G>A
intron 12 IVS12−24G>A
IVS12−34 40delGAAGTCC
exon 13 1905T>C ?Splicing
1907G>A G590E
1909T>C S591P
1918T>A S594T
1950C>T
1960C>G Q608E
1960C>A Q608K
1972A>G K612E
1977C>T
1988T>C
2055A>G
Abcg8 5′-UTR −67A>T
−135G>A
−148G>A
−168C>T
−212C>T
−225C>A
intron 1 IVS1+33C>G
exon 2 277T>C
intron 2 IVS2+25G>T
IVS2+59A>G
IVS2−44T>A
exon3 284T>C
intron 3 IVS3+80C>T
IVS3+96G>C
exon 4 449C>T
509T>C
548T>G
551G>A
620G>A
intron 4 IVS4+39A>T
IVS4−22T>G
IVS4−24C>T
exon 5 683C>T
intron 5 IVS5+28A>G
IVS5+62G>A
IVS5+80G>A
IVS5−16G>A
IVS5−34G>A
exon 6 854T>C
991C>A A297E
1058G>A
intron 6 IVS6+5T>C
IVS6+6T>C
exon7 1163A>G
intron7 IVS7+25G>A
IVS7−35G>A
IVS7−40C>T
exon8 1254G>A D385N
1270T>C V390A
intron 8 IVS8−18A>T
IVS8−52T>C
exon 9 1394T>C
1400C>T
1445A>G
1472G>A
intron9 IVS9+46A>G
IVS9+64T>A
IVS9+79G>C
IVS9+81 81delC
IVS9+93T>C
intron 10 IVS10+24G>A
IVS10−24A>G
IVS10−39A>G
IVS10−43T>C
IVS10−62G>A
IVS10−79A>G
exon 11 1697C>A F532L
1703C>T
1820T>C
1835A>C
intron 11 IVS11+12T>C
IVS11+16 18delCCA
IVS11+24C>T
IVS11+27C>T
IVS11−27A>G
exon 12 1871C>A
1883G>T
1892G>C
1904G>C
1946A>G
1950A>G I617V
1971A>G I624V
intron 12 IVS12−17C>T
IVS12−28T>C
IVS12−34C>T
exon 13 1989A>G I630V
1997C>T
2045C>T
2057C>T
2075C>T
2081A>G
3′-UTR *25G>A
*29G>A
*37 37delT
*41G>A
*65C>G
*86A>G
*108 110delACCinsTAG
*193A>G
*199G>A
*208G>A
*297G>A
*361A>G
*362C>T
*412C>T
*413G>A
*417T>C
*451C>T
*461T>C
*510T>C
*513T>C
*559T>A
a

Numbering based on Genbank No. AF312713 (Abcg5) and Genbank No. AF324495 (Abcg8)

b

Numbering based upon the first ATG as + 1

c

Strains C57/L, C57BL/6ByJ, BALB/c, and PERA/Ei were identical to C57BL/6J.

All sequences are wild-type, unless indicated by the check mark. Sequences in bold indicate known coding changes.

We selected strains that showed genetic variations (C57BL/6J, CAST/Ei, MOLF/Ei, RBF/DnJ, DBA/1J, DBA/2J, FVB/NJ, and CE/J) and measured their plasma sitosterol levels after a week on standard rodent chow (61 mg/kg cholesterol, 31 mg/kg campesterol, and 105 mg/kg sitosterol). Despite the large genetic variations identified between some of the strains, none of these mice showed any detectable levels of plasma sitosterol (lower limit of detection ~0.1 mg/dl). A small amount of campesterol was detectable in most samples (<1.5 mg/dl), but these did not show any significant patterns. Given the small sample size (n = 2 per group), we do not draw any conclusions about the campesterol values.

DISCUSSION

We report here the identification and characterization of the murine genes Abcg5 and Abcg8, which are the homologs of human ABCG5 and ABCG8 genes, mutations of which are now known to cause the autosomal recessive disorder sitosterolemia (1416). The protein products of these genes, sterolin-1 and sterolin-2 respectively, are predicted to play a crucial role in the exclusion of dietary non-cholesterol sterols from the body, as well as sterol excretion by the liver into bile. The identification of these genes as defective in sitosterolemia has given credence to the hypothesis that dietary cholesterol absorption is regulated by a molecular mechanism(s) specific for cholesterol (1). Under normal circumstances, although our normal diets consist of almost equal amounts of plant sterols and cholesterol, on average about 55% of dietary cholesterol and less than 1% of plant sterols are retained by the body (33). This selectivity of intestinal absorption appears to extend to other sterols such as shellfish sterols and this selectivity is also exhibited by the normal liver for its preferential excretion of these sterols into bile (non-cholesterol sterols > cholesterol) (5). These processes are disrupted in sitosterolemia (2,3).

Murine Abcg5 and Abcg8 are located on chromosome 17 at ~53 cM, a region syntenic to human chromosome 2p21, where ABCG5 and ABCG8 have been mapped previously. Both the polypeptides as well as the genes show a high degree of conservation between human and mouse (14). The intron-exon organization is well preserved between the two species (data not shown) and at the protein level there is greater than 80% identity, with more than 90% conservation, suggesting that these genes play an important and highly conserved function in both species. Abcg5 and Abcg8 are also related to each other and appear to have arisen by a mechanism of gene duplication and inversion. Both proteins belong to the ABCG family, containing highly conserved ABC motifs, and overall organization; both proteins contain an ATP binding domain located in a larger N-terminal domain, predicted to be cytoplasmic, with a 6-transmembrane domain located at the carboxyl termini. The genes are arranged in a head-to-head configuration in the genome and are unusual as they are separated from each other by no more than 140 bases (4,1416). Abcg5 and Abcg8 are expressed in the intestine and liver only and we have previously presented genetic evidence that their two protein products, sterolin-1 and -2, may function as obligate heterodimers (4,16). Given their proximity to each other, the lack of an obvious promoter (except for a TATA box identified in the mouse sequence only), a head-to-head configuration, and the very short distance separating them, the transcriptional control of these genes is likely to be unusual. Berge et al. identified Abcg5 and Abcg8 cDNAs as transcripts that were induced after rexinoid exposure, suggesting that liver X receptor (LXR)-retinoid X receptor may be involved in their regulation (15). Repa et al. showed that LXR deficiency affected cholesterol absorption (34). Thus LXR is a strong candidate as a regulatory transcriptional factor. If these proteins act as obligate heterodimers for normal function, they would be predicted to be transcriptionally active in the same cell and transcribed in a co-ordinate manner. Given the very short distance separating the two transcriptional start-sites, some interference of transcription would be predicted, unless transcription of each gene was temporally separated. Preliminary data suggests that the minimal region, encompassing about 358 bases between the two respective ATGs, is not sufficient to drive expression of a reporter gene when transfected in human HEK 293 cells (data not shown).

Both the human ABCG8 and the mouse Abcg8 exhibit alternative splicing caused by a CAGCAG repeat at a splice-acceptor site. However, for the human, this repeat occurs in intron 7 (16), whereas such a repeat is present in intron 1 for the mouse. Additionally, almost 50% of the mouse transcripts show alternative splicing, whereas only about 10% of the human ABCG8 transcripts are alternatively spliced (16). Alternative splicing results in the deletion of a single amino acid, the functional consequence of which is not known at present. Amino-acid residues that are affected by the splicing variants are not highly conserved between human, rat, and mouse, thus alternative splicing may not have any functional significance (4,16). Although relatively rare, three other genes that harbor such repeats at the splice-acceptor boundaries and have been shown to lead to alternatively spliced mRNAs have been previously reported (3537).

Finally, the murine genes are located at ~53 cM on chromosome 17, and although a large number of quantitative trait loci (QTL) for obesity, atherosclerosis, cholesterol metabolism, or lithogenic bile in mouse have been reported (2832, 38), none of these QTLs map to this region. Thus whether genetic variation at these loci is an important determinant of any of these phenotypes remains to be determined. Direct comparison between different mouse strains has also been reported (23, 25, 26, 39). We screened 20 mouse strains and identified a very large number of polymorphic variants in both Abcg5 and Abcg8. However, in a sub-set of these mice fed a standard rodent chow diet, plasma sitosterol levels were essentially undetectable (our lower limit of detection is 0.1 mg/dl). Standard rodent chow is rich in plant sterols, containing more than 1.5-fold more plant sterol relative to cholesterol. Based upon an average daily food consumption, a 25 g mouse may consume 5 g of rodent chow, representing a daily intake of ~20 mg/kg body weight of sitosterol. In comparison, an average daily diet in humans contains ~400 mg of sitosterol per day, equivalent to ~6 mg/kg body weight. Thus the rodent chow results in ~3-fold more plant sterol daily intake in the mouse relative to the human. One explanation for a lack of detection of an effect of polymorphisms in Abcg5 or Abcg8 on plasma sitosterol levels is that none of the variations identified have any consequences, despite some of these polymorphisms altering amino acids. These data are limited in that cholesterol, campesterol, and sitosterol absorption rates were not directly measured and none of these strains were placed on a challenge diet (such as the atherogenic or lithogenic diets). However, the information presented should allow a more direct approach in examining whether any of the genetic changes identified lead to physiological consequences. Another important observation is that some of the inbred mouse strains were highly variant, such as DBA, CAST/Ei, and MOLF/Ei, while others were almost completely identical at these loci, such as AKR/J, FVB/NJ, 129/SvEv, and C57BL/6J. The most variant strain is the DBA, a mouse that has been reported to have hyporesponsiveness to high fat and cholesterol diets (23). In this context, it is important to note that many investigators maintain their own long term mouse colonies and these animals may show genetic variations not seen in our animals obtained from the Jackson Laboratories. Nevertheless, accurate measurements of dietary sterol absorption rates, as well as biliary excretion could now be undertaken to test the hypothesis that Abcg5 and Abcg8 are involved in dietary sterol balance.

Acknowledgments

The authors are grateful to the General Clinical Research Center, Medical University of South Carolina (MUSC) and the Bio-molecular Resource Center, MUSC for assistance with sequencing, and the members of the Patel lab for helpful discussions. This work was funded by a Scientist Development Award from the American Heart Association grant 9730087N (S.B.P.), the National Institutes of Health grants HL60616 (S.B.P.), MO1 RR01070-25 (MUSC GCRC), the Summer Undergraduate Program, MUSC (S.A.S.), and by an intramural award from the University Research Committee, MUSC (S.B.P.).

References

  • 1.Bhattacharyya AK, Connor WE. Beta-sitosterolemia and xanthomatosis. A newly described lipid storage disease in two sisters. J Clin Invest. 1974;53:1033–1043. doi: 10.1172/JCI107640. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Bjorkhem, I., and K. M. Boberg. 1995. Inborn errors in bile acid biosynthesis and storage of sterols other than cholesterol. In The Metabolic Basis of Inherited Disease. C. R. Scriver, A. L. Beaudet, W. S. Sly, and D. Valle, editors. McGraw-Hill Inc., New York. 2073–2102.
  • 3.Salen G, Shefer S, Nguyen L, Ness GC, Tint GS, Shore V. Sitosterolemia. J Lipid Res. 1992;33:945–955. [PubMed] [Google Scholar]
  • 4.Lee MH, Lu K, Patel SB. Genetic basis of sitosterolemia. Curr Opin Lipid. 2001;12:141–149. doi: 10.1097/00041433-200104000-00007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Gregg RE, Connor WE, Lin DS, Brewer, Jr H. Abnormal metabolism of shellfish sterols in a patient with sitosterolemia and xanthomatosis. J Clin Invest. 1986;77:1864–1872. doi: 10.1172/JCI112513. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Salen G, Shore V, Tint GS, Forte T, Shefer S, Horak I, Horak E, Dayal B, Nguyen L, Batta AK. Increased sitosterol absorption, decreased removal, and expanded body pools compensate for reduced cholesterol synthesis in sitosterolemia with xanthomatosis. J Lipid Res. 1989;30:1319–1330. [PubMed] [Google Scholar]
  • 7.Bhattacharyya AK, Connor WE, Lin DS, McMurry MM, Shulman RS. Sluggish sitosterol turnover and hepatic failure to excrete sitosterol into bile cause expansion of body pool of sitosterol in patients with sitosterolemia and xanthomatosis. Arterioscler Thromb. 1991;11:1287–1294. doi: 10.1161/01.atv.11.5.1287. [DOI] [PubMed] [Google Scholar]
  • 8.Salen G, Ahrens, Jr E, Grundy SM. Metabolism of beta-sitosterol in man. J Clin Invest. 1970;49:952–967. doi: 10.1172/JCI106315. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Nguyen LB, Salen G, Shefer S, Tint GS, Shore V, Ness GC. Decreased cholesterol biosynthesis in sitosterolemia with xanthomatosis: diminished mononuclear leukocyte 3-hydroxy-3-methylglutaryl coenzyme A reductase activity and enzyme protein associated with increased low-density lipoprotein receptor function. Metab: Clin & Exp. 1990;39:436–443. doi: 10.1016/0026-0495(90)90260-j. [DOI] [PubMed] [Google Scholar]
  • 10.Patel SB, Honda A, Salen G. Sitosterolemia: exclusion of genes involved in reduced cholesterol biosynthesis. J Lipid Res. 1998;39:1055–1061. [PubMed] [Google Scholar]
  • 11.Patel SB, Salen G, Hidaka H, Kwiterovich PO, Stalenhoef AF, Miettinen TA, Grundy SM, Lee MH, Rubenstein JS, Polymeropoulos MH, Brownstein MJ. Mapping a gene involved in regulating dietary cholesterol absorption. The sitosterolemia locus is found at chromosome 2p21. J Clin Invest. 1998;102:1041–1044. doi: 10.1172/JCI3963. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Lee MH, Gordon D, Ott J, Lu K, Ose L, Miettinen T, Gylling H, Stalenhoef AF, Pandya A, Hidaka H, Brewer J, Kojima BH, Sakuma N, Pegoraro R, Salen G, Patel SB. Fine mapping of a gene responsible for regulating dietary cholesterol absorption; founder effects underlie cases of phytosterolemia in multiple communities. Eur Hum Genet. 2001;9:375–384. doi: 10.1038/sj.ejhg.5200628. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Lu K, Lee MH, Carpten JD, Sekhon M, Patel SB. High-Resolution Physical and Transcript Map of Human Chromosome 2p21 Containing the Sitosterolemia Locus. Eur Hum Genet. 2001;9:364–374. doi: 10.1038/sj.ejhg.5200627. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Lee MH, Lu K, Hazard S, Yu H, Shulenin S, Hidaka H, Kojima H, Allikmets R, Sakuma N, Pegoraro R, Srivastava AK, Salen G, Dean M, Patel SB. Identification of a gene, ABCG5, important in the regulation of dietary cholesterol absorption. Nat Gen. 2001;27:79–83. doi: 10.1038/83799. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Berge KE, Tian H, Graf GA, Yu L, Grishin NV, Schultz J, Kwiterovich P, Shan B, Barnes R, Hobbs HH. Accumulation of dietary cholesterol in sitosterolemia caused by mutations in adjacent ABC transporters. Science. 2000;290:1771–1775. doi: 10.1126/science.290.5497.1771. [DOI] [PubMed] [Google Scholar]
  • 16.Lu K, Lee MH, Hazard S, Brooks-Wilson A, Hidaka H, Kojima H, Ose L, Stanlenhoef AFH, Mietinnen T, Bjorkhem I, Brukert E, Pandya A, Brewer, Jr HB, Salen G, Dean M, Srivastava A, Patel SB. Two genes that map to the STSL locus cause sitosterolemia: Genomic structure and spectrum of mutations involving sterolin-1 and sterolin-2, encoded by ABCG5 and ABCG8 respectively. Am J Hum Genet. 2001;69:278–290. doi: 10.1086/321294. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Lu K, Lee M, Patel SB. Dietary cholesterol absorption; more than just bile. Trends Endo Metab. 2001;12:314–320. doi: 10.1016/s1043-2760(01)00433-7. [DOI] [PubMed] [Google Scholar]
  • 18.Altschul SF, Gish W, Miller W, Myers EW, Lipman DJ. Basic local alignment search tool. J Molec Biol. 1990;215:403–410. doi: 10.1016/S0022-2836(05)80360-2. [DOI] [PubMed] [Google Scholar]
  • 19.Altschul SF, Madden TL, Schaffer AA, Zhang J, Zhang Z, Miller W, Lipman DJ. Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res. 1997;25:3389–3402. doi: 10.1093/nar/25.17.3389. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Lennon G, Auffray C, Polymeropoulos M, Soares MB. The I.M.A.G.E. Consortium: an integrated molecular analysis of genomes and their expression. Genomics. 1996;33:151–152. doi: 10.1006/geno.1996.0177. [DOI] [PubMed] [Google Scholar]
  • 21.Sambrook, J., E. F. Fritsch, and T. Maniatis. 1989. Molecular Cloning: A Laboratory Manual. Laboratory Press, Cold Spring Harbor.
  • 22.Paigen K, Eppig JT. A mouse phenome project. Mamm Genome. 2000;11:715–717. doi: 10.1007/s003350010152. [DOI] [PubMed] [Google Scholar]
  • 23.Kirk EA, Moe GL, Caldwell MT, Lernmark JA, Wilson DL, LeBoeuf RC. Hyper- and hypo-responsiveness to dietary fat and cholesterol among inbred mice: searching for level and variability genes. J Lipid Res. 1995;36:1522–1532. [PubMed] [Google Scholar]
  • 24.Howles PN, Carter CP, Hui DY. Dietary free and esterified cholesterol absorption in cholesterol esterase (bile salt-stimulated lipase) gene-targeted mice. J Biol Chem. 1996;271:7196–7202. doi: 10.1074/jbc.271.12.7196. [DOI] [PubMed] [Google Scholar]
  • 25.Carter CP, Howles PN, Hui DY. Genetic variation in cholesterol absorption efficiency among inbred strains of mice. J Nutr. 1997;127:1344–1348. doi: 10.1093/jn/127.7.1344. [DOI] [PubMed] [Google Scholar]
  • 26.Jolley CD, Dietschy JM, Turley SD. Genetic differences in cholesterol absorption in 129/Sv and C57BL/6 mice: effect on cholesterol responsiveness. Am J Physiol. 1999;276:G1117–G1124. doi: 10.1152/ajpgi.1999.276.5.G1117. [DOI] [PubMed] [Google Scholar]
  • 27.Salen G, Kwiterovich, Jr P, Shefer S, Tint GS, Horak I, Shore V, Dayal B, Horak E. Increased plasma cholestanol and 5 alpha-saturated plant sterol derivatives in subjects with sitosterolemia and xanthomatosis. J Lipid Res. 1985;26:203–209. [PubMed] [Google Scholar]
  • 28.Nishina PM, Wang J, Toyofuku W, Kuypers FA, Ishida BY, Paigen B. Atherosclerosis and plasma and liver lipids in nine inbred strains of mice. Lipids. 1993;28:599–605. doi: 10.1007/BF02536053. [DOI] [PubMed] [Google Scholar]
  • 29.Purcell-Huynh DA, Weinreb A, Castellani LW, Mehrabian M, Doolittle MH, Lusis AJ. Genetic factors in lipoprotein metabolism. Analysis of a genetic cross between inbred mouse strains NZB/BINJ and SM/J using a complete linkage map approach. J Clin Invest. 1995;96:1845–1858. doi: 10.1172/JCI118230. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Mehrabian M, Castellani LW, Wen PZ, Wong J, Rithaporn T, Hama SY, Hough GP, Johnson D, Albers JJ, Mottino GA, Frank JS, Navab M, Fogelman AM, Lusis AJ. Genetic control of HDL levels and composition in an interspecific mouse cross (CAST/Ei x C57BL/6J) J Lipid Res. 2000;41:1936–1946. [PubMed] [Google Scholar]
  • 31.Paigen B, Schork NJ, Svenson KL, Cheah YC, Mu JL, Lammert F, Wang DQ, Bouchard G, Carey MC. Quantitative trait loci mapping for cholesterol gallstones in AKR/J and C57L/J strains of mice. Physiol Genomics. 2000;4:59–65. doi: 10.1152/physiolgenomics.2000.4.1.59. [DOI] [PubMed] [Google Scholar]
  • 32.Perusse L, Chagnon YC, Weisnagel SJ, Rankinen T, Snyder E, Sands J, Bouchard C. The human obesity gene map: the 2000 update. Obes Res. 2001;9:135–169. doi: 10.1038/oby.2001.17. [DOI] [PubMed] [Google Scholar]
  • 33.Gould RG, Jones RJ, LeRoy GV, Wissler RW, Taylor CB. Absorbability of beta-sitosterol in humans. Metabolism. 1969;18:652–662. doi: 10.1016/0026-0495(69)90078-x. [DOI] [PubMed] [Google Scholar]
  • 34.Repa JJ, Turley SD, Lobaccaro JA, Medina J, Li L, Lustig K, Shan B, Heyman RA, Dietschy JM, Mangelsdorf DJ. Regulation of absorption and ABC1-mediated efflux of cholesterol by RXR heterodimers. Science. 2000;289:1524–1529. doi: 10.1126/science.289.5484.1524. [DOI] [PubMed] [Google Scholar]
  • 35.Yee D, Lebovic GS, Marcus RR, Rosen N. Identification of an alternate type I insulin-like growth factor receptor beta subunit mRNA transcript. J Biol Chem. 1989;264:21439–21441. [PubMed] [Google Scholar]
  • 36.Moingeon P, Stebbins CC, D’Adamio L, Lucich J, Reinherz EL. Human natural killer cells and mature T lymphocytes express identical CD3 zeta subunits as defined by cDNA cloning and sequence analysis. Eur J Immunol. 1990;20:1741–1745. doi: 10.1002/eji.1830200818. [DOI] [PubMed] [Google Scholar]
  • 37.Rogina B, Upholt WB. The chicken homeobox gene Hoxd-11 encodes two alternatively spliced RNA species. Biochem Mol Biol Int. 1995;35:825–831. [PubMed] [Google Scholar]
  • 38.Paigen B. Genetics of responsiveness to high-fat and high-cholesterol diets in the mouse. Am J Clin Nutr. 1995;62:458S–462S. doi: 10.1093/ajcn/62.2.458S. [DOI] [PubMed] [Google Scholar]
  • 39.Keelan M, Hui DY, Wild G, Clandinin MT, Thomson AB. Variability of the intestinal uptake of lipids is genetically determined in mice. Lipids. 2000;35:833–837. doi: 10.1007/s11745-000-0592-0. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES