Entry - *300747 - STEROID SULFATASE; STS - OMIM
* 300747

STEROID SULFATASE; STS


Alternative titles; symbols

ARYLSULFATASE C; ARSC
ESTRONE SULFATASE


Other entities represented in this entry:

STEROID SULFATASE, ISOZYME S, INCLUDED

HGNC Approved Gene Symbol: STS

Cytogenetic location: Xp22.31     Genomic coordinates (GRCh38): X:7,147,290-7,354,641 (from NCBI)


Gene-Phenotype Relationships
Location Phenotype Phenotype
MIM number
Inheritance Phenotype
mapping key
Xp22.31 Ichthyosis, X-linked 308100 XLR 3

TEXT

Description

The STS gene encodes steroid sulfatase (EC 3.1.6.2), a membrane-bound microsomal enzyme that is ubiquitously expressed and hydrolyzes several 3-beta-hydroxysteroid sulfates, which serve as metabolic precursors for estrogens, androgens, and cholesterol (Stein et al., 1989; Alperin and Shapiro, 1997).


Cloning and Expression

Ballabio et al. (1987) and Bonifas et al. (1987) independently isolated cDNA clones corresponding to the STS gene from human placenta cDNA libraries. Northern blot analysis detected 1.5- and 2.7-kb isoforms in cultured keratinocytes (Bonifas et al., 1987). Yen et al. (1987) isolated and sequenced cDNA clones containing the entire coding sequence of STS.

Stein et al. (1989) isolated a 2.4-kb clone corresponding to the human STS gene from a human placenta cDNA library. The deduced 583-residue protein has a molecular mass of 63 kD and contains a 21- or 23-residue signal peptide, 4 possible N-linked glycosylation sites, and 2 potential membrane-spanning domains. Expression in hamster kidney cells (BHK-21) showed localization predominantly in the endoplasmic reticulum, with smaller fractions found in the Golgi, at the cell surface, and in endosomes and lysosomes. The nucleotide sequence and numbering differed from that reported by Yen et al. (1987) and is used preferentially.

Alperin and Shapiro (1997) noted the changes in the numbering of the nucleotides and amino acids in the human STS cDNA. Nucleotide +1 was positioned at the major STS transcription start site instead of at the beginning of the STS cDNA, as had been reported by Yen et al. (1987). The major transcription start site was at position -221 with respect to the AUG translation initiation codon, and the amino acids were renumbered to make the initiating methionine +1. The initiating methionine had previously been numbered -22 to reflect the predicted signal sequence. These changes agreed with the numbering of some of the other sulfatases.

Cloning of the Mouse and Rat Sts Genes

Several attempts to clone the mouse homolog of the human STS gene failed, suggesting a substantial divergence between these genes. However, Kawano et al. (1989) showed that partial N-terminal sequence from purified rat liver STS was very similar to its human counterpart, and sequence comparisons revealed several domains that are conserved among all the sulfatases characterized to date.

Salido et al. (1996) used a degenerate-primer RT-PCR approach to amplify a conserved fragment of a rat STS cDNA that was then used to clone a mouse Sts cDNA. This 2.3-kb cDNA showed 75% similarity with the rat STS, whereas it was only 63% similar to human STS. Li et al. (1996) also used a degenerate-primer RT-PCR approach to amplify a 321-bp fragment from rat liver cDNA, which was used as a probe to clone and characterize the complete cDNA. Comparison of the complete coding region between the rat and human genes showed 66% homology both at the DNA and the protein levels. Authenticity of the cloned cDNA was indicated by the fact that STS activity was conferred to STS(-) A9 cells upon transfection with a rat Sts expression construct. The rat gene is only 8.2 kb long, whereas the human gene spans over 146 kb.


Gene Structure

The STS gene contains 10 exons and spans about 146 kb (Yen et al., 1987; Alperin and Shapiro, 1997).


Mapping

In studies of X-chromosome anomalies in man-mouse hybrids, Mohandas et al. (1979) concluded that the steroid sulfatase locus is on Xpter-p22. Because of the linkage of X-linked ichthyosis to Xg (314700), somatic cell hybridization indicated that Xg is also on Xp.

Curry et al. (1984) found that the steroid sulfatase, Xg, and MIC2X (313470) loci, as well as the locus for X-linked chondrodysplasia punctata (302950), were apparently absent in males with deletion of Xp22.32.

Ballabio et al. (1987) mapped the STS gene to chromosome Xp22.3. No homologous sequences were detected on the Y chromosome.

By several approaches, Mohandas et al. (1990) demonstrated that in XX males in whom the testis-determining factor gene (TDF; 480000) was transferred from a Y chromatid to an X chromatid, the breakpoint was distal to the STS locus, which was retained on the TDF-bearing X chromosome.

Pseudogene

Yen et al. (1987) identified an STS pseudogene on Yq in humans, suggesting a recent pericentric inversion.

Mapping of the Mouse and Rat Sts Genes

Gartler and Rivest (1983) confirmed X-linkage of STS in the mouse by the study of oocytes of XX and XO mice. Assays of STS in kidney tissue of these mice indicated dosage compensation for the gene, which is different from the situation in man.

Keitges et al. (1985) resolved the question of whether STS is autosomal or X-linked in the mouse. They showed that it is X-linked or pseudoautosomal. Their results indirectly indicated the existence of a functional STS allele on the Y chromosome which undergoes obligatory recombination during meiosis with the X-linked allele. Their experiments consisted of crosses between STS-deficient C3H/An male mice and STS-normal XO animals. STS should map to the same region as Sxr ('sex-reversed'), which, from its equal transmission to male and female offspring, appears also to be on a homologous pairing segment of X and Y.

As an explanation for the fact that the STS locus is in the pseudoautosomal segment of the X and Y in the mouse but not in man, Yen et al. (1988) suggested that a pericentric inversion of the Y chromosome occurred during primate evolution, disrupting the former pseudoautosomal arrangement of these genes. In man, an STS pseudogene is present on the long arm of the Y chromosome. Obligatory recombination of pairing segments of the human X and Y appears to be excluded by results from study of a polymorphism of red cell antigen 12E7 (encoded by MIC2; 313470) which showed a complex sex-limited expression of variation in 12E7 levels (Goodfellow and Tippett, 1981).

Keitges and Gartler (1986) confirmed the existence of functional Y-linked and X-linked alleles for Sts in the mouse by dosage studies in XO, XX, and XY mice and by clonal analysis in fibroblast cell cultures from mice heterozygous for steroid sulfatase deficiency and for the X-linked electrophoretic variant of phosphoglycerate kinase (PGK1; 311800).

The mouse Sts gene was mapped physically by Salido et al. (1996) to the distal end of the mouse sex chromosomes (X and Y). Backcross studies placed Sts distal to the 'obligatory' crossover in male meiosis. The fact that the gene is pseudoautosomal in the mouse and not in the human suggested divergence. While Sts is located in the mouse pseudoautosomal region, both physical and genetic mapping demonstrated that STS is not pseudoautosomal in the rat (Li et al., 1996).

Kipling et al. (1996) reported physical linkage of 3 mouse pseudoautosomal region (PAR) probes: DXYHgu1, DXYMov15, and the telomeric sequence (TTAGGG)n. They found that in the mouse the Sts locus maps distal to these 3 probes, indicating that there is an internal array of the telomere sequence in the PAR. They found that pseudoautosomal PacI restriction fragments, up to 2 Mb in size, are unstable in C57BL/6 x C57BL/6 crosses. New alleles, often several hundred kilobases different in size, occurred at a sex-averaged rate of approximately 30% per allele. Kipling et al. (1996) noted that such frequent large-scale germline genome rearrangements were without precedent in mammals.

Also see EVOLUTION section.


Gene Function

Gant et al. (1977) showed that steroid sulfatase resided in the chorion laeve; the amnion was totally devoid of this activity.

Chang et al. (1986) demonstrated that arylsulfatase C consisted of 2 isozymes, 's' and 'f' (ARSC2; 301780), that were biochemically and immunologically distinct. Only the s form had steroid sulfatase activity when acting on 3-beta-hydroxysteroid sulfates. The f form acted on 4-methylumbelliferyl sulfate. The findings indicated that 'arylsulfatase C' per se was not necessarily identical with steroid sulfatase and suggesting genetic heterogeneity. Chang et al. (1990) found that although both isozymes were linked to the human X chromosome and both escaped X inactivition, they were not related by posttranslational modification of the same gene product. The findings were consistent with the 2 isozymes resulting from separate genes. Shankaran et al. (1991) also found that only the placental s form of arylsulfatase C had steroid sulfatase activity and hydrolyzes estrone sulfate, dehydroepiandrosterone sulfate, and cholesterol sulfate. The liver f form had barely detectable activity towards these sterol sulfates. With the artificial substrate 4-methylumbelliferyl sulfate, both forms demonstrated a similar activity, but had different optimum pH. Polyclonal antibodies raised against the placental form reacted specifically against the s and not the f form. Shankaran et al. (1991) concluded that the 2 isozymes of arylsulfatase C in humans purified from placenta and liver, respectively, are distinct proteins with different substrate specificity, pH optima, heat lability, and antigenic properties.

Non-X-Inactivation (Non-Lyonization) of the STS Gene

Shapiro et al. (1979) found that the STS locus did not lyonize, despite its location on the X chromosome. In fibroblasts doubly heterozygous for steroid sulfatase and G6PD (305900), steroid sulfatase was expressed in all clones regardless of whether the X chromosome was active or not, as indicated by the G6PD activity of the clone. Conflicting results were obtained by Balazs et al. (1979), who concluded that the STS locus was situated between Xq13 and Xq24 and that it lyonized regularly. Tracing back from an STS-deficient mouse cell line, they showed that STS was X-linked in the mouse from which the line was derived.

Muller et al. (1980) demonstrated that steroid sulfatase activity was higher in normal females than in normal males, a finding consistent with nonlyonization. Furthermore, they demonstrated that heterozygotes could clearly be distinguished both from normal females and from hemizygous affected males. Thus, heterozygote detection was not impeded by the usual vagaries of lyonization. The authors noted that it was curious that the STS locus has maintained its X chromosomal localization.

Ropers et al. (1981) found that although the STS locus is not normally inactivated, it may be when located on an aberrant X chromosome. Similar inactivation patterns had been reported for the Xg locus.

Migeon et al. (1982) pursued the possibility of incomplete inactivation, or incomplete escape from inactivation, by examining STS levels in fibroblast clones from women heterozygous for deficiency of both STS and G6PD. Their study demonstrated that the wildtype STS allele was expressed at about half the level from the inactive chromosome compared with its expression from the active homolog. Partial inactivation was the most satisfactory explanation for this observation and would account for the deviation from a strict dosage relationship observed for the female:male ratio. XO persons and STS-deficiency heterozygotes had enzyme levels below the normal male range.

Vogel et al. (1984) confirmed the non-inactivation of the STS gene as well as of the ARSC2 gene. The authors studied gene dosage in fibroblasts from a 45,X/47,XXX mosaic and from a 69,XXY triploidy with 1 or 2 active X chromosomes. The comparison of the 47,XXX with 45,X clones showed an incomplete gene dosage effect (1.8 for STS and 2.0 for ARSC ). This was not the case for the triploid clones with different X-inactivation patterns. These results confirmed previous reports on the non-inactivation of the STS gene, and established X linkage and non-inactivation for the ARSC gene as well.

Craig and Tolley (1986) reviewed the relationship between STS and mammalian X-chromosome conservation. They surveyed information available on the female:male ratio for STS levels in fibroblasts and placenta and found the average value of 1.6 rather than the expected 2.0 if no lyonization occurs. Ratios observed for peripheral white cells were even lower. Furthermore, individuals with multiple X chromosomes did not exhibit the proportionately high levels of enzymes anticipated.


Molecular Genetics

More than 85% of patients with STS deficiency (308100) have a large deletion involving the entire STS gene and its flanking sequences (Basler et al., 1992). In 12 unrelated Italian patients with STS deficiency, including 8 with classic X-linked ichthyosis, Ballabio et al. (1987) found a deletion of the STS gene using a cDNA clone. Although 1 patient had an X/Y translocation with a deletion of the Xpter-p22 region, the others did not show karyotypic abnormalities. Similarly, Bonifas et al. (1987) found gross deletions of the STS gene in 14 of 15 apparently unrelated families with X-linked ichthyosis. Yen et al. (1987) identified complete STS gene deletions in 8 of 10 patients with inherited STS deficiency.

Conary et al. (1987) found that DNA from 2 patients with STS deficiency showed lack of hybridization with an STS clone. DNA from a third patient showed a normal hybridization pattern.

Wirth et al. (1988) found deletion involving the STS locus in 8 of 9 unrelated families with X-linked ichthyosis. Three patients in the ninth family had no evident deletion when studied with 2 probes. Since approximately 90% of STS-deficient persons have large deletions at the STS locus, Shapiro et al. (1989) investigated the breakpoints to identify potential sequences prone to undergo either intrachromosomal or interchromosomal nonhomologous recombination. Most of the breakpoints occurred at a distance from the STS gene itself and were, therefore, difficult to characterize. They found 1 subject who had an entirely intragenic deletion of 40 kb, permitting cloning and sequencing of the deletion junction. Bernatowicz et al. (1992) stated that only 2 patients with partial deletions of the STS gene had been reported. One was the person studied by Shapiro et al. (1989) who had an intragenic deletion extending from intron 1 to intron 5. The breakpoints contained no detectable secondary structure or repetitive elements except an 8-bp direct repeat located 8 bp 5-prime of the deletion junction. Bernatowicz et al. (1992) characterized the breakpoints in the other patient who had a deletion of the 3-prime end of the STS gene. The deletion started within intron 7 of the gene and extended over 150 kb downstream toward the centromere. Analysis of sequences flanking the deletion breakpoints revealed 3 bp of homology. (The 5-prime end of the STS gene is oriented toward the telomere.)

Basler et al. (1990, 1992) identified 3 different point mutations in the STS gene (300747.0001-300747.0003) in 3 unrelated patients with X-linked ichthyosis. Alperin and Shapiro (1997) identified 3 additional point mutations in the STS gene (300747.0004-300747.0006) in patients with X-linked ichthyosis and reviewed the point mutations reported by Basler et al. (1990, 1992). All 6 mutations were located in a 105-amino acid region of the C-terminal half of the polypeptide. Five of the 6 mutations were missense, whereas 1 resulted in a frameshift and premature protein termination. In vitro functional expression studies showed that all 6 mutants lacked STS enzymatic activity.


Evolution

Yen et al. (1987) raised the possibility that STS deficiency results from aberrant X-Y interchange. Comparative in situ hybridization in various primate species demonstrated a pseudoautosomal location of the human ANT3 gene (300151) and an X-specific location for the STS gene throughout the higher primate species up to the New World monkeys. However, Toder et al. (1995) found that ANT3 and STS map together on an autosome of 2 prosimian species of the genera Lemur and Eulemur. These results suggested an autosome-to-X/Y translocation after the simians radiated from the prosimians, resulting in a pseudoautosomal location of genes, such as ANT3 and STS. In simian primates, STS then became X-specific by a pericentric inversion in the Y chromosome followed by mutational inactivation of the Y allele.


Animal Model

Eicher (1974) speculated that the 'scurfy' (sf) mutation in the mouse may be homologous to X-linked ichthyosis of man. Buckle et al. (1985) alluded to ichthyosis with male hypogonadism (see 308200) as an entity separate from ichthyosis with steroid sulfatase deficiency and homologous to 'scurfy' in the mouse. From comparative mapping of the X chromosomes of mouse and man, they predicted that this possibly separate human condition may be determined by a mutation on Xp near OTC (300461).

Lyon et al. (1990) described hematologic abnormalities in the 'scurfy' mouse and raised a question of homology to Wiskott-Aldrich syndrome (WAS; 301000) rather than STS deficiency. In comparing gene order in mouse and human, Laval and Boyd (1993) found evidence for a partial inversion of gene order within a homologous segment of the X chromosome between DXS255 and TIMP (305370). In the 2 species, the scurfy/WAS phenotypes and the GATA1 (305371)/Gf-1 loci mapped to the same region of the X chromosome. The findings supported the possibility that scurfy and WAS are indeed homologous.

Ropers and Wiberg (1982) demonstrated that STS is also X-linked and noninactivated in the wood lemming, Myopus schisticolor. According to the work of Cooper et al. (1984), steroid sulfatase is not X-linked in Australian marsupials. Correlated with this are the facts that the 'basic' marsupial X is smaller than the 'basic' eutherian X, and the X and Y of Australian marsupials lack a pairing segment.


ALLELIC VARIANTS ( 6 Selected Examples):

.0001 ICHTHYOSIS, X-LINKED

STS, TRP372ARG
  
RCV000011298

In a male patient with X-linked ichthyosis (308100) resulting from STS deficiency, Basler et al. (1990, 1992) identified a 1320T-A transversion in the STS gene in the numbering system reported by Yen et al. (1987). This mutation corresponds to a trp372-to-arg (W372R) substitution in the revised numbering system of Stein et al. (1989) (Alperin and Shapiro, 1997).


.0002 ICHTHYOSIS, X-LINKED

STS, CYS446TYR
  
RCV000011299

In a patient with X-linked ichthyosis (308100), Basler et al. (1990, 1992) identified a 1543G-A transition in the STS gene in the numbering system reported by Yen et al. (1987). This mutation corresponds to a cys446-to-tyr (C446Y) substitution in the revised numbering system of Stein et al. (1989) (Alperin and Shapiro, 1997).


.0003 ICHTHYOSIS, X-LINKED

STS, SER341LEU
  
RCV000011300...

In a patient with X-linked ichthyosis (308100), Basler et al. (1992) identified a 1226C-T transition in the STS gene in the numbering system reported by Yen et al. (1987). This mutation corresponds to a ser341-to-leu (S341L) substitution in the revised numbering system of Stein et al. (1989) (Alperin and Shapiro, 1997).


.0004 ICHTHYOSIS, X-LINKED

STS, TRP372PRO
  
RCV000011301

In a patient with X-linked ichthyosis (308100), Alperin and Shapiro (1997) observed a 1336G-C transversion in the STS gene, resulting in a trp372-to-pro (W372P) substitution. This is the same residue as that involved in W372R (300747.0001).


.0005 ICHTHYOSIS, X-LINKED

STS, HIS444ARG
  
RCV000011302

In a patient with X-linked ichthyosis (308100), Alperin and Shapiro (1997) identified a 1552A-G transition in the STS gene, resulting in a his444-to-arg (H444R) substitution.


.0006 ICHTHYOSIS, X-LINKED

STS, IVS8DS, G-T, +1
  
RCV000011303

In a patient with X-linked ichthyosis (308100), Alperin and Shapiro (1997) identified a 19-bp insertion starting at nucleotide 1477 of the STS gene. A G-to-T transversion was identified at the exon 8/intron 8 splice donor site and confirmed in genomic DNA. This splice junction mutation results in the addition of 19 bp from intron 8 to the STS mRNA, changing the reading frame. The predicted polypeptide prematurely terminates at residue 427, 8 amino acids after the frameshift. As a result, the mutant STS polypeptide was predicted to lose 156 residues from its C terminus.


REFERENCES

  1. Alperin, E. S., Shapiro, L. J. Characterization of point mutations with X-linked ichthyosis: effects on the structure and function of the steroid sulfatase protein. J. Biol. Chem. 272: 20756-20763, 1997. [PubMed: 9252398, related citations] [Full Text]

  2. Balazs, I., Filippi, G., Rinaldi, A., Grzeschik, K.-H., Siniscalco, M. Studies on X-linked ichthyosis and steroid sulfatase in man, mice and their hybrids. (Abstract) Cytogenet. Cell Genet. 25: 133, 1979.

  3. Ballabio, A., Parenti, G., Carrozzo, R., Sebastio, G., Andria, G., Buckle, V., Fraser, N., Craig, I., Rocchi, M., Romeo, G., Jobsis, A. C., Persico, M. G. Isolation and characterization of a steroid sulfatase cDNA clone: genomic deletions in patients with X-chromosome-linked ichthyosis. Proc. Nat. Acad. Sci. 84: 4519-4523, 1987. [PubMed: 3474618, related citations] [Full Text]

  4. Basler, E., Grompe, M., Parenti, G., Caskey, C. T., Ballabio, A. Identification of point mutations in three patients with steroid sulfatase deficiency. (Abstract) Am. J. Hum. Genet. 47 (suppl.): A150, 1990.

  5. Basler, E., Grompe, M., Parenti, G., Yates, J., Ballabio, A. Identification of point mutations in the steroid sulfatase gene of three patients with X-linked ichthyosis. Am. J. Hum. Genet. 50: 483-491, 1992. [PubMed: 1539590, related citations]

  6. Bernatowicz, L. F., Li, X.-M., Carrozzo, R., Ballabio, A., Mohandas, T., Yen, P. H., Shapiro, L. J. Sequence analysis of a partial deletion of the human steroid sulfatase gene reveals 3 bp of homology at deletion breakpoints. Genomics 13: 892-893, 1992. [PubMed: 1639422, related citations] [Full Text]

  7. Bonifas, J. M., Morley, B. J., Oakey, R. E., Kan, Y. W., Epstein, E. H., Jr. Cloning of a cDNA for steroid sulfatase: frequent occurrence of gene deletions in patients with recessive X chromosome-linked ichthyosis. Proc. Nat. Acad. Sci. 84: 9248-9251, 1987. [PubMed: 3480541, related citations] [Full Text]

  8. Buckle, V. J., Edwards, J. H., Evans, E. P., Jonasson, J. A., Lyon, M. F., Peters, J., Searle, A. G. Comparative maps of human and mouse X chromosomes. (Abstract) Cytogenet. Cell Genet. 40: 594-595, 1985.

  9. Burns, G. On the identity of arylsulphatase C and steroid sulphatase. Hum. Genet. 65: 189, 1983. [PubMed: 6581118, related citations] [Full Text]

  10. Chance, P. F., Gartler, S. M. Evidence for a dosage effect at the X-linked steroid sulfatase locus in human tissues. Am. J. Hum. Genet. 35: 234-240, 1983. [PubMed: 6573129, related citations]

  11. Chang, P. L., Lafferty, K. I., Rosa, N. E., Davidson, R. G. Arylsulfatase-C isozymes in human tissues. (Abstract) Am. J. Hum. Genet. 33: 38A, 1981.

  12. Chang, P. L., Mueller, O. T., Lafrenie, R. M., Varey, P. A., Rosa, N. E., Davidson, R. G., Henry, W. M., Shows, T. B. The human arylsulfatase-C isoenzymes: two distinct genes that escape from X inactivation. Am. J. Hum. Genet. 46: 729-737, 1990. [PubMed: 1690506, related citations]

  13. Chang, P. L., Varey, P. A., Rosa, N. E., Ameen, M., Davidson, R. G. Association of steroid sulfatase with one of the arylsulfatase C isozymes in human fibroblasts. J. Biol. Chem. 261: 14443-14447, 1986. [PubMed: 3464600, related citations]

  14. Conary, J. T., Lorkowski, G., Schmidt, B., Pohlmann, R., Nagel, G., Meyer, H. E., Krentler, C., Cully, J., Hasilik, A., von Figura, K. Genetic heterogeneity of steroid sulfatase deficiency revealed with cDNA for human steroid sulfatase. Biochem. Biophys. Res. Commun. 144: 1010-1017, 1987. [PubMed: 3034252, related citations] [Full Text]

  15. Cooper, D. W., McAllan, B. M., Donald, J. A., Dawson, G., Dobrovic, A., Marshall Graves, J. A. Steroid sulphatase is not detected on the X chromosome of Australian marsupials. (Abstract) Cytogenet. Cell Genet. 37: 439, 1984.

  16. Craig, I. W., Tolley, E. Steroid sulphatase and the conservation of mammalian X chromosomes. Trends Genet. 2: 201-204, 1986.

  17. Curry, C. J. R., Magenis, R. E., Brown, M., Lanman, J. T., Jr., Tsai, J., O'Lague, P., Goodfellow, P., Mohandas, T., Bergner, E. A., Shapiro, L. J. Inherited chondrodysplasia punctata due to a deletion of the terminal short arm of an X chromosome. New Eng. J. Med. 311: 1010-1015, 1984. [PubMed: 6482910, related citations] [Full Text]

  18. Eicher, E. M. Personal Communication. Bar Harbor, Me. 1974.

  19. Gant, N. F., Milewich, L., Calvert, M. E., MacDonald, P. C. Steroid sulfatase activity in human fetal membranes. J. Clin. Endocr. 45: 965-972, 1977. [PubMed: 144740, related citations] [Full Text]

  20. Gartler, S. M., Andina, R. J. Mammalian X-chromosome inactivation. Adv. Hum. Genet. 7: 99-140, 1976. [PubMed: 797247, related citations] [Full Text]

  21. Gartler, S. M., Rivest, M. Evidence for X-linkage of steroid sulfatase in the mouse: steroid sulfatase levels in oocytes of XX and XO mice. Genetics 103: 137-141, 1983. [PubMed: 6573292, related citations] [Full Text]

  22. Goodfellow, P. N., Tippett, P. A human quantitative polymorphism related to Xg blood groups. Nature 289: 404-409, 1981. [PubMed: 7464908, related citations] [Full Text]

  23. Kawano, J., Kotani, T., Ohtaki, S., Mimamino, N., Matsuo, H., Oinuma, T., Aikawa, E. Characterization of rat and human steroid sulfatases. Biochim. Biophys. Acta 997: 199-205, 1989. [PubMed: 2765556, related citations] [Full Text]

  24. Keitges, E., Gartler, S. M. Dosage of the Sts gene in the mouse. Am. J. Hum. Genet. 39: 470-476, 1986. [PubMed: 3464202, related citations]

  25. Keitges, E., Rivest, M., Siniscalco, M., Gartler, S. M. X-linkage of steroid sulphatase in the mouse is evidence for a functional Y-linked allele. Nature 315: 226-227, 1985. [PubMed: 3858674, related citations] [Full Text]

  26. Kipling, D., Salido, E. C., Shapiro, L. J., Cooke, H. J. High frequency de novo alterations in the long-range genomic structure of the mouse pseudoautosomal region. Nature Genet. 13: 78-82, 1996. [PubMed: 8673108, related citations] [Full Text]

  27. Laval, S. H., Boyd, Y. Partial inversion of gene order within a homologous segment on the X chromosome. Mammalian Genome 4: 119-123, 1993. [PubMed: 8431636, related citations] [Full Text]

  28. Li, X. M., Salido, E. C., Gong, Y., Kitada, K., Serikawa, T., Yen, P. H., Shapiro, L. J. Cloning of the rat steroid sulfatase gene (Sts), a non-pseudoautosomal X-linked gene that undergoes X inactivation. Mammalian Genome 7: 420-424, 1996. [PubMed: 8662223, related citations] [Full Text]

  29. Lyon, M. F., Peters, J., Glenister, P. H., Ball, S., Wright, E. The scurfy mouse mutant has previously unrecognized hematological abnormalities and resembles Wiskott-Aldrich syndrome. Proc. Nat. Acad. Sci. 87: 2433-2437, 1990. [PubMed: 2320565, related citations] [Full Text]

  30. Migeon, B. R., Shapiro, L. J., Norum, R. A., Mohandas, T., Axelman, J., Dabora, R. L. Differential expression of steroid sulphatase locus on active and inactive human X chromosome. Nature 299: 838-840, 1982. [PubMed: 6957717, related citations] [Full Text]

  31. Mohandas, T. K., Stern, H. J., Meeker, C. A., Passage, M. B., Muller, U., Page, D. C., Yen, P. H., Shapiro, L. J. Steroid sulfatase gene in XX males. Am. J. Hum. Genet. 46: 369-376, 1990. [PubMed: 2301402, related citations]

  32. Mohandas, T., Shapiro, L. J., Sparkes, R. S., Sparkes, M. C. Regional assignment of the steroid sulfatase--X-linked ichthyosis locus: implications for a non-inactivated region on the short arm of the human X-chromosome. Proc. Nat. Acad. Sci. 76: 5779-5783, 1979. [PubMed: 293682, related citations] [Full Text]

  33. Mohandas, T., Sparkes, R. S., Hellkuhl, B., Grzeschik, K. H., Shapiro, L. J. Expression of an X-linked gene from an inactive human X chromosome in mouse-human hybrid cells: further evidence for the noninactivation of the steroid sulfatase locus in man. Proc. Nat. Acad. Sci. 77: 6759-6763, 1980. [PubMed: 6935682, related citations] [Full Text]

  34. Monroe, D. G., Chang, P. L. Tissue-specific expression of human arylsulfatase-C isozymes and steroid sulfatase. Am. J. Hum. Genet. 40: 102-114, 1987. [PubMed: 3471087, related citations]

  35. Muller, C. R., Migl, B., Ropers, H.-H. X-linked steroid sulfatase: evidence for different gene-dosage in males and females. Hum. Genet. 54: 197-199, 1980. [PubMed: 6930360, related citations] [Full Text]

  36. Muller, C. R., Wahlstrom, J., Ropers, H.-H. Further evidence for the assignment of the steroid sulfatase X-linked ichthyosis locus to the telomer of Xp. (Letter) Hum. Genet. 58: 446, 1981. [PubMed: 6948769, related citations] [Full Text]

  37. Nomura, K., Nakano, H., Umeki, K., Harada, K., Kon, A., Tamai, K., Sawamura, D., Hashimoto, I. A study of the steroid sulfatase gene in families with X-linked ichthyosis using polymerase chain reaction. Acta Derm. Venereol. 75: 340-342, 1995. [PubMed: 8615047, related citations] [Full Text]

  38. Ropers, H.-H., Migl, B., Zimmer, J., Fraccaro, M., Maraschio, P. P., Westerveld, A. Activity of steroid sulfatase in fibroblasts with numerical and structural X chromosome aberrations. Hum. Genet. 57: 354-356, 1981. [PubMed: 6945285, related citations] [Full Text]

  39. Ropers, H.-H., Wiberg, U. Evidence for X-linkage and non-inactivation of steroid sulphatase locus in wood lemming. Nature 296: 766-767, 1982. [PubMed: 7040982, related citations] [Full Text]

  40. Salido, E. C., Li, X. M., Yen, P. H., Martin, N., Mohandas, T. K., Shapiro, L. J. Cloning and expression of the mouse pseudoautosomal steroid sulphatase gene (Sts). Nature Genet. 13: 83-86, 1996. [PubMed: 8673109, related citations] [Full Text]

  41. Shankaran, R., Ameen, M., Daniel, W. L., Davidson, R. G., Chang, P. L. Characterization of arylsulfatase C isozymes from human liver and placenta. Biochim. Biophys. Acta 1078: 251-257, 1991. [PubMed: 2065092, related citations] [Full Text]

  42. Shapiro, L. J., Mohandas, T., Weiss, R., Romeo, G. Non-activation of a X-chromosome locus in man. Science 204: 1224-1226, 1979. [PubMed: 156396, related citations] [Full Text]

  43. Shapiro, L. J., Yen, P., Pomerantz, D., Martin, E., Rolewic, L., Mohandas, T. Molecular studies of deletions at the human steroid sulfatase locus. Proc. Nat. Acad. Sci. 86: 8477-8481, 1989. [PubMed: 2813406, related citations] [Full Text]

  44. Stein, C., Hille, A., Seidel, J., Rijnbout, S., Waheed, A., Schmidt, B., Geuze, H., von Figura, K. Cloning and expression of human steroid-sulfatase: membrane topology, glycosylation, and subcellular distribution in BHK-21 cells. J. Biol. Chem. 264: 13865-13872, 1989. [PubMed: 2668275, related citations]

  45. Toder, R., Rappold, G. A., Schiebel, K., Schempp, W. ANT3 and STS are autosomal in prosimian lemurs: implications for the evolution of the pseudoautosomal region. Hum. Genet. 95: 22-28, 1995. [PubMed: 7814020, related citations] [Full Text]

  46. Vogel, W., Grompe, M., Storz, R., Pentz, S. A comparative study on steroid sulfatase and arylsulfatase C in fibroblast clones from 45,X/47,XXX and 69,XXY. Hum. Genet. 66: 367-369, 1984. [PubMed: 6586638, related citations] [Full Text]

  47. Willard, H. F., Holmes, M. T. A sensitive and dependable assay for distinguishing hamster and human X-linked steroid sulfatase activity in somatic cell hybrids. Hum. Genet. 66: 272-275, 1984. [PubMed: 6585346, related citations] [Full Text]

  48. Wirth, B., Herrmann, F. H., Neugebauer, M., Gillard, E. F., Wulff, K., Stein, C., von Figura, K., Ferguson-Smith, M. A., Gal, A. Linkage analysis in X-linked ichthyosis (steroid sulfatase deficiency). Hum. Genet. 80: 191-192, 1988. [PubMed: 3169744, related citations] [Full Text]

  49. Yen, P. H., Allen, E., Marsh, B., Mohandas, T., Wang, N., Taggart, R. T., Shapiro, L. J. Cloning and expression of steroid sulfatase cDNA and the frequent occurrence of deletions in STS deficiency: implications for X-Y interchange. Cell 49: 443-454, 1987. [PubMed: 3032454, related citations] [Full Text]

  50. Yen, P. H., Marsh, B., Allen, E., Tsai, S. P., Ellison, J., Connolly, L., Neiswanger, K., Shapiro, L. J. The human X-linked steroid sulfatase gene and a Y-encoded pseudogene: evidence for an inversion of the Y chromosome during primate evolution. Cell 55: 1123-1135, 1988. [PubMed: 3203382, related citations] [Full Text]


Creation Date:
Cassandra L. Kniffin : 10/28/2008
carol : 08/04/2016
carol : 11/03/2008
ckniffin : 11/3/2008
carol : 10/31/2008
ckniffin : 10/31/2008

* 300747

STEROID SULFATASE; STS


Alternative titles; symbols

ARYLSULFATASE C; ARSC
ESTRONE SULFATASE


Other entities represented in this entry:

STEROID SULFATASE, ISOZYME S, INCLUDED

HGNC Approved Gene Symbol: STS

SNOMEDCT: 72523005;   ICD10CM: Q80.1;  


Cytogenetic location: Xp22.31     Genomic coordinates (GRCh38): X:7,147,290-7,354,641 (from NCBI)


Gene-Phenotype Relationships

Location Phenotype Phenotype
MIM number
Inheritance Phenotype
mapping key
Xp22.31 Ichthyosis, X-linked 308100 X-linked recessive 3

TEXT

Description

The STS gene encodes steroid sulfatase (EC 3.1.6.2), a membrane-bound microsomal enzyme that is ubiquitously expressed and hydrolyzes several 3-beta-hydroxysteroid sulfates, which serve as metabolic precursors for estrogens, androgens, and cholesterol (Stein et al., 1989; Alperin and Shapiro, 1997).


Cloning and Expression

Ballabio et al. (1987) and Bonifas et al. (1987) independently isolated cDNA clones corresponding to the STS gene from human placenta cDNA libraries. Northern blot analysis detected 1.5- and 2.7-kb isoforms in cultured keratinocytes (Bonifas et al., 1987). Yen et al. (1987) isolated and sequenced cDNA clones containing the entire coding sequence of STS.

Stein et al. (1989) isolated a 2.4-kb clone corresponding to the human STS gene from a human placenta cDNA library. The deduced 583-residue protein has a molecular mass of 63 kD and contains a 21- or 23-residue signal peptide, 4 possible N-linked glycosylation sites, and 2 potential membrane-spanning domains. Expression in hamster kidney cells (BHK-21) showed localization predominantly in the endoplasmic reticulum, with smaller fractions found in the Golgi, at the cell surface, and in endosomes and lysosomes. The nucleotide sequence and numbering differed from that reported by Yen et al. (1987) and is used preferentially.

Alperin and Shapiro (1997) noted the changes in the numbering of the nucleotides and amino acids in the human STS cDNA. Nucleotide +1 was positioned at the major STS transcription start site instead of at the beginning of the STS cDNA, as had been reported by Yen et al. (1987). The major transcription start site was at position -221 with respect to the AUG translation initiation codon, and the amino acids were renumbered to make the initiating methionine +1. The initiating methionine had previously been numbered -22 to reflect the predicted signal sequence. These changes agreed with the numbering of some of the other sulfatases.

Cloning of the Mouse and Rat Sts Genes

Several attempts to clone the mouse homolog of the human STS gene failed, suggesting a substantial divergence between these genes. However, Kawano et al. (1989) showed that partial N-terminal sequence from purified rat liver STS was very similar to its human counterpart, and sequence comparisons revealed several domains that are conserved among all the sulfatases characterized to date.

Salido et al. (1996) used a degenerate-primer RT-PCR approach to amplify a conserved fragment of a rat STS cDNA that was then used to clone a mouse Sts cDNA. This 2.3-kb cDNA showed 75% similarity with the rat STS, whereas it was only 63% similar to human STS. Li et al. (1996) also used a degenerate-primer RT-PCR approach to amplify a 321-bp fragment from rat liver cDNA, which was used as a probe to clone and characterize the complete cDNA. Comparison of the complete coding region between the rat and human genes showed 66% homology both at the DNA and the protein levels. Authenticity of the cloned cDNA was indicated by the fact that STS activity was conferred to STS(-) A9 cells upon transfection with a rat Sts expression construct. The rat gene is only 8.2 kb long, whereas the human gene spans over 146 kb.


Gene Structure

The STS gene contains 10 exons and spans about 146 kb (Yen et al., 1987; Alperin and Shapiro, 1997).


Mapping

In studies of X-chromosome anomalies in man-mouse hybrids, Mohandas et al. (1979) concluded that the steroid sulfatase locus is on Xpter-p22. Because of the linkage of X-linked ichthyosis to Xg (314700), somatic cell hybridization indicated that Xg is also on Xp.

Curry et al. (1984) found that the steroid sulfatase, Xg, and MIC2X (313470) loci, as well as the locus for X-linked chondrodysplasia punctata (302950), were apparently absent in males with deletion of Xp22.32.

Ballabio et al. (1987) mapped the STS gene to chromosome Xp22.3. No homologous sequences were detected on the Y chromosome.

By several approaches, Mohandas et al. (1990) demonstrated that in XX males in whom the testis-determining factor gene (TDF; 480000) was transferred from a Y chromatid to an X chromatid, the breakpoint was distal to the STS locus, which was retained on the TDF-bearing X chromosome.

Pseudogene

Yen et al. (1987) identified an STS pseudogene on Yq in humans, suggesting a recent pericentric inversion.

Mapping of the Mouse and Rat Sts Genes

Gartler and Rivest (1983) confirmed X-linkage of STS in the mouse by the study of oocytes of XX and XO mice. Assays of STS in kidney tissue of these mice indicated dosage compensation for the gene, which is different from the situation in man.

Keitges et al. (1985) resolved the question of whether STS is autosomal or X-linked in the mouse. They showed that it is X-linked or pseudoautosomal. Their results indirectly indicated the existence of a functional STS allele on the Y chromosome which undergoes obligatory recombination during meiosis with the X-linked allele. Their experiments consisted of crosses between STS-deficient C3H/An male mice and STS-normal XO animals. STS should map to the same region as Sxr ('sex-reversed'), which, from its equal transmission to male and female offspring, appears also to be on a homologous pairing segment of X and Y.

As an explanation for the fact that the STS locus is in the pseudoautosomal segment of the X and Y in the mouse but not in man, Yen et al. (1988) suggested that a pericentric inversion of the Y chromosome occurred during primate evolution, disrupting the former pseudoautosomal arrangement of these genes. In man, an STS pseudogene is present on the long arm of the Y chromosome. Obligatory recombination of pairing segments of the human X and Y appears to be excluded by results from study of a polymorphism of red cell antigen 12E7 (encoded by MIC2; 313470) which showed a complex sex-limited expression of variation in 12E7 levels (Goodfellow and Tippett, 1981).

Keitges and Gartler (1986) confirmed the existence of functional Y-linked and X-linked alleles for Sts in the mouse by dosage studies in XO, XX, and XY mice and by clonal analysis in fibroblast cell cultures from mice heterozygous for steroid sulfatase deficiency and for the X-linked electrophoretic variant of phosphoglycerate kinase (PGK1; 311800).

The mouse Sts gene was mapped physically by Salido et al. (1996) to the distal end of the mouse sex chromosomes (X and Y). Backcross studies placed Sts distal to the 'obligatory' crossover in male meiosis. The fact that the gene is pseudoautosomal in the mouse and not in the human suggested divergence. While Sts is located in the mouse pseudoautosomal region, both physical and genetic mapping demonstrated that STS is not pseudoautosomal in the rat (Li et al., 1996).

Kipling et al. (1996) reported physical linkage of 3 mouse pseudoautosomal region (PAR) probes: DXYHgu1, DXYMov15, and the telomeric sequence (TTAGGG)n. They found that in the mouse the Sts locus maps distal to these 3 probes, indicating that there is an internal array of the telomere sequence in the PAR. They found that pseudoautosomal PacI restriction fragments, up to 2 Mb in size, are unstable in C57BL/6 x C57BL/6 crosses. New alleles, often several hundred kilobases different in size, occurred at a sex-averaged rate of approximately 30% per allele. Kipling et al. (1996) noted that such frequent large-scale germline genome rearrangements were without precedent in mammals.

Also see EVOLUTION section.


Gene Function

Gant et al. (1977) showed that steroid sulfatase resided in the chorion laeve; the amnion was totally devoid of this activity.

Chang et al. (1986) demonstrated that arylsulfatase C consisted of 2 isozymes, 's' and 'f' (ARSC2; 301780), that were biochemically and immunologically distinct. Only the s form had steroid sulfatase activity when acting on 3-beta-hydroxysteroid sulfates. The f form acted on 4-methylumbelliferyl sulfate. The findings indicated that 'arylsulfatase C' per se was not necessarily identical with steroid sulfatase and suggesting genetic heterogeneity. Chang et al. (1990) found that although both isozymes were linked to the human X chromosome and both escaped X inactivition, they were not related by posttranslational modification of the same gene product. The findings were consistent with the 2 isozymes resulting from separate genes. Shankaran et al. (1991) also found that only the placental s form of arylsulfatase C had steroid sulfatase activity and hydrolyzes estrone sulfate, dehydroepiandrosterone sulfate, and cholesterol sulfate. The liver f form had barely detectable activity towards these sterol sulfates. With the artificial substrate 4-methylumbelliferyl sulfate, both forms demonstrated a similar activity, but had different optimum pH. Polyclonal antibodies raised against the placental form reacted specifically against the s and not the f form. Shankaran et al. (1991) concluded that the 2 isozymes of arylsulfatase C in humans purified from placenta and liver, respectively, are distinct proteins with different substrate specificity, pH optima, heat lability, and antigenic properties.

Non-X-Inactivation (Non-Lyonization) of the STS Gene

Shapiro et al. (1979) found that the STS locus did not lyonize, despite its location on the X chromosome. In fibroblasts doubly heterozygous for steroid sulfatase and G6PD (305900), steroid sulfatase was expressed in all clones regardless of whether the X chromosome was active or not, as indicated by the G6PD activity of the clone. Conflicting results were obtained by Balazs et al. (1979), who concluded that the STS locus was situated between Xq13 and Xq24 and that it lyonized regularly. Tracing back from an STS-deficient mouse cell line, they showed that STS was X-linked in the mouse from which the line was derived.

Muller et al. (1980) demonstrated that steroid sulfatase activity was higher in normal females than in normal males, a finding consistent with nonlyonization. Furthermore, they demonstrated that heterozygotes could clearly be distinguished both from normal females and from hemizygous affected males. Thus, heterozygote detection was not impeded by the usual vagaries of lyonization. The authors noted that it was curious that the STS locus has maintained its X chromosomal localization.

Ropers et al. (1981) found that although the STS locus is not normally inactivated, it may be when located on an aberrant X chromosome. Similar inactivation patterns had been reported for the Xg locus.

Migeon et al. (1982) pursued the possibility of incomplete inactivation, or incomplete escape from inactivation, by examining STS levels in fibroblast clones from women heterozygous for deficiency of both STS and G6PD. Their study demonstrated that the wildtype STS allele was expressed at about half the level from the inactive chromosome compared with its expression from the active homolog. Partial inactivation was the most satisfactory explanation for this observation and would account for the deviation from a strict dosage relationship observed for the female:male ratio. XO persons and STS-deficiency heterozygotes had enzyme levels below the normal male range.

Vogel et al. (1984) confirmed the non-inactivation of the STS gene as well as of the ARSC2 gene. The authors studied gene dosage in fibroblasts from a 45,X/47,XXX mosaic and from a 69,XXY triploidy with 1 or 2 active X chromosomes. The comparison of the 47,XXX with 45,X clones showed an incomplete gene dosage effect (1.8 for STS and 2.0 for ARSC ). This was not the case for the triploid clones with different X-inactivation patterns. These results confirmed previous reports on the non-inactivation of the STS gene, and established X linkage and non-inactivation for the ARSC gene as well.

Craig and Tolley (1986) reviewed the relationship between STS and mammalian X-chromosome conservation. They surveyed information available on the female:male ratio for STS levels in fibroblasts and placenta and found the average value of 1.6 rather than the expected 2.0 if no lyonization occurs. Ratios observed for peripheral white cells were even lower. Furthermore, individuals with multiple X chromosomes did not exhibit the proportionately high levels of enzymes anticipated.


Molecular Genetics

More than 85% of patients with STS deficiency (308100) have a large deletion involving the entire STS gene and its flanking sequences (Basler et al., 1992). In 12 unrelated Italian patients with STS deficiency, including 8 with classic X-linked ichthyosis, Ballabio et al. (1987) found a deletion of the STS gene using a cDNA clone. Although 1 patient had an X/Y translocation with a deletion of the Xpter-p22 region, the others did not show karyotypic abnormalities. Similarly, Bonifas et al. (1987) found gross deletions of the STS gene in 14 of 15 apparently unrelated families with X-linked ichthyosis. Yen et al. (1987) identified complete STS gene deletions in 8 of 10 patients with inherited STS deficiency.

Conary et al. (1987) found that DNA from 2 patients with STS deficiency showed lack of hybridization with an STS clone. DNA from a third patient showed a normal hybridization pattern.

Wirth et al. (1988) found deletion involving the STS locus in 8 of 9 unrelated families with X-linked ichthyosis. Three patients in the ninth family had no evident deletion when studied with 2 probes. Since approximately 90% of STS-deficient persons have large deletions at the STS locus, Shapiro et al. (1989) investigated the breakpoints to identify potential sequences prone to undergo either intrachromosomal or interchromosomal nonhomologous recombination. Most of the breakpoints occurred at a distance from the STS gene itself and were, therefore, difficult to characterize. They found 1 subject who had an entirely intragenic deletion of 40 kb, permitting cloning and sequencing of the deletion junction. Bernatowicz et al. (1992) stated that only 2 patients with partial deletions of the STS gene had been reported. One was the person studied by Shapiro et al. (1989) who had an intragenic deletion extending from intron 1 to intron 5. The breakpoints contained no detectable secondary structure or repetitive elements except an 8-bp direct repeat located 8 bp 5-prime of the deletion junction. Bernatowicz et al. (1992) characterized the breakpoints in the other patient who had a deletion of the 3-prime end of the STS gene. The deletion started within intron 7 of the gene and extended over 150 kb downstream toward the centromere. Analysis of sequences flanking the deletion breakpoints revealed 3 bp of homology. (The 5-prime end of the STS gene is oriented toward the telomere.)

Basler et al. (1990, 1992) identified 3 different point mutations in the STS gene (300747.0001-300747.0003) in 3 unrelated patients with X-linked ichthyosis. Alperin and Shapiro (1997) identified 3 additional point mutations in the STS gene (300747.0004-300747.0006) in patients with X-linked ichthyosis and reviewed the point mutations reported by Basler et al. (1990, 1992). All 6 mutations were located in a 105-amino acid region of the C-terminal half of the polypeptide. Five of the 6 mutations were missense, whereas 1 resulted in a frameshift and premature protein termination. In vitro functional expression studies showed that all 6 mutants lacked STS enzymatic activity.


Evolution

Yen et al. (1987) raised the possibility that STS deficiency results from aberrant X-Y interchange. Comparative in situ hybridization in various primate species demonstrated a pseudoautosomal location of the human ANT3 gene (300151) and an X-specific location for the STS gene throughout the higher primate species up to the New World monkeys. However, Toder et al. (1995) found that ANT3 and STS map together on an autosome of 2 prosimian species of the genera Lemur and Eulemur. These results suggested an autosome-to-X/Y translocation after the simians radiated from the prosimians, resulting in a pseudoautosomal location of genes, such as ANT3 and STS. In simian primates, STS then became X-specific by a pericentric inversion in the Y chromosome followed by mutational inactivation of the Y allele.


Animal Model

Eicher (1974) speculated that the 'scurfy' (sf) mutation in the mouse may be homologous to X-linked ichthyosis of man. Buckle et al. (1985) alluded to ichthyosis with male hypogonadism (see 308200) as an entity separate from ichthyosis with steroid sulfatase deficiency and homologous to 'scurfy' in the mouse. From comparative mapping of the X chromosomes of mouse and man, they predicted that this possibly separate human condition may be determined by a mutation on Xp near OTC (300461).

Lyon et al. (1990) described hematologic abnormalities in the 'scurfy' mouse and raised a question of homology to Wiskott-Aldrich syndrome (WAS; 301000) rather than STS deficiency. In comparing gene order in mouse and human, Laval and Boyd (1993) found evidence for a partial inversion of gene order within a homologous segment of the X chromosome between DXS255 and TIMP (305370). In the 2 species, the scurfy/WAS phenotypes and the GATA1 (305371)/Gf-1 loci mapped to the same region of the X chromosome. The findings supported the possibility that scurfy and WAS are indeed homologous.

Ropers and Wiberg (1982) demonstrated that STS is also X-linked and noninactivated in the wood lemming, Myopus schisticolor. According to the work of Cooper et al. (1984), steroid sulfatase is not X-linked in Australian marsupials. Correlated with this are the facts that the 'basic' marsupial X is smaller than the 'basic' eutherian X, and the X and Y of Australian marsupials lack a pairing segment.


ALLELIC VARIANTS 6 Selected Examples):

.0001   ICHTHYOSIS, X-LINKED

STS, TRP372ARG
SNP: rs137853165, ClinVar: RCV000011298

In a male patient with X-linked ichthyosis (308100) resulting from STS deficiency, Basler et al. (1990, 1992) identified a 1320T-A transversion in the STS gene in the numbering system reported by Yen et al. (1987). This mutation corresponds to a trp372-to-arg (W372R) substitution in the revised numbering system of Stein et al. (1989) (Alperin and Shapiro, 1997).


.0002   ICHTHYOSIS, X-LINKED

STS, CYS446TYR
SNP: rs137853166, ClinVar: RCV000011299

In a patient with X-linked ichthyosis (308100), Basler et al. (1990, 1992) identified a 1543G-A transition in the STS gene in the numbering system reported by Yen et al. (1987). This mutation corresponds to a cys446-to-tyr (C446Y) substitution in the revised numbering system of Stein et al. (1989) (Alperin and Shapiro, 1997).


.0003   ICHTHYOSIS, X-LINKED

STS, SER341LEU
SNP: rs137853167, ClinVar: RCV000011300, RCV001093287

In a patient with X-linked ichthyosis (308100), Basler et al. (1992) identified a 1226C-T transition in the STS gene in the numbering system reported by Yen et al. (1987). This mutation corresponds to a ser341-to-leu (S341L) substitution in the revised numbering system of Stein et al. (1989) (Alperin and Shapiro, 1997).


.0004   ICHTHYOSIS, X-LINKED

STS, TRP372PRO
SNP: rs137853168, ClinVar: RCV000011301

In a patient with X-linked ichthyosis (308100), Alperin and Shapiro (1997) observed a 1336G-C transversion in the STS gene, resulting in a trp372-to-pro (W372P) substitution. This is the same residue as that involved in W372R (300747.0001).


.0005   ICHTHYOSIS, X-LINKED

STS, HIS444ARG
SNP: rs137853169, ClinVar: RCV000011302

In a patient with X-linked ichthyosis (308100), Alperin and Shapiro (1997) identified a 1552A-G transition in the STS gene, resulting in a his444-to-arg (H444R) substitution.


.0006   ICHTHYOSIS, X-LINKED

STS, IVS8DS, G-T, +1
SNP: rs1601748137, ClinVar: RCV000011303

In a patient with X-linked ichthyosis (308100), Alperin and Shapiro (1997) identified a 19-bp insertion starting at nucleotide 1477 of the STS gene. A G-to-T transversion was identified at the exon 8/intron 8 splice donor site and confirmed in genomic DNA. This splice junction mutation results in the addition of 19 bp from intron 8 to the STS mRNA, changing the reading frame. The predicted polypeptide prematurely terminates at residue 427, 8 amino acids after the frameshift. As a result, the mutant STS polypeptide was predicted to lose 156 residues from its C terminus.


See Also:

Burns (1983); Chance and Gartler (1983); Chang et al. (1981); Gartler and Andina (1976); Goodfellow and Tippett (1981); Mohandas et al. (1980); Monroe and Chang (1987); Muller et al. (1981); Nomura et al. (1995); Willard and Holmes (1984)

REFERENCES

  1. Alperin, E. S., Shapiro, L. J. Characterization of point mutations with X-linked ichthyosis: effects on the structure and function of the steroid sulfatase protein. J. Biol. Chem. 272: 20756-20763, 1997. [PubMed: 9252398] [Full Text: https://doi.org/10.1074/jbc.272.33.20756]

  2. Balazs, I., Filippi, G., Rinaldi, A., Grzeschik, K.-H., Siniscalco, M. Studies on X-linked ichthyosis and steroid sulfatase in man, mice and their hybrids. (Abstract) Cytogenet. Cell Genet. 25: 133, 1979.

  3. Ballabio, A., Parenti, G., Carrozzo, R., Sebastio, G., Andria, G., Buckle, V., Fraser, N., Craig, I., Rocchi, M., Romeo, G., Jobsis, A. C., Persico, M. G. Isolation and characterization of a steroid sulfatase cDNA clone: genomic deletions in patients with X-chromosome-linked ichthyosis. Proc. Nat. Acad. Sci. 84: 4519-4523, 1987. [PubMed: 3474618] [Full Text: https://doi.org/10.1073/pnas.84.13.4519]

  4. Basler, E., Grompe, M., Parenti, G., Caskey, C. T., Ballabio, A. Identification of point mutations in three patients with steroid sulfatase deficiency. (Abstract) Am. J. Hum. Genet. 47 (suppl.): A150, 1990.

  5. Basler, E., Grompe, M., Parenti, G., Yates, J., Ballabio, A. Identification of point mutations in the steroid sulfatase gene of three patients with X-linked ichthyosis. Am. J. Hum. Genet. 50: 483-491, 1992. [PubMed: 1539590]

  6. Bernatowicz, L. F., Li, X.-M., Carrozzo, R., Ballabio, A., Mohandas, T., Yen, P. H., Shapiro, L. J. Sequence analysis of a partial deletion of the human steroid sulfatase gene reveals 3 bp of homology at deletion breakpoints. Genomics 13: 892-893, 1992. [PubMed: 1639422] [Full Text: https://doi.org/10.1016/0888-7543(92)90179-v]

  7. Bonifas, J. M., Morley, B. J., Oakey, R. E., Kan, Y. W., Epstein, E. H., Jr. Cloning of a cDNA for steroid sulfatase: frequent occurrence of gene deletions in patients with recessive X chromosome-linked ichthyosis. Proc. Nat. Acad. Sci. 84: 9248-9251, 1987. [PubMed: 3480541] [Full Text: https://doi.org/10.1073/pnas.84.24.9248]

  8. Buckle, V. J., Edwards, J. H., Evans, E. P., Jonasson, J. A., Lyon, M. F., Peters, J., Searle, A. G. Comparative maps of human and mouse X chromosomes. (Abstract) Cytogenet. Cell Genet. 40: 594-595, 1985.

  9. Burns, G. On the identity of arylsulphatase C and steroid sulphatase. Hum. Genet. 65: 189, 1983. [PubMed: 6581118] [Full Text: https://doi.org/10.1007/BF00286661]

  10. Chance, P. F., Gartler, S. M. Evidence for a dosage effect at the X-linked steroid sulfatase locus in human tissues. Am. J. Hum. Genet. 35: 234-240, 1983. [PubMed: 6573129]

  11. Chang, P. L., Lafferty, K. I., Rosa, N. E., Davidson, R. G. Arylsulfatase-C isozymes in human tissues. (Abstract) Am. J. Hum. Genet. 33: 38A, 1981.

  12. Chang, P. L., Mueller, O. T., Lafrenie, R. M., Varey, P. A., Rosa, N. E., Davidson, R. G., Henry, W. M., Shows, T. B. The human arylsulfatase-C isoenzymes: two distinct genes that escape from X inactivation. Am. J. Hum. Genet. 46: 729-737, 1990. [PubMed: 1690506]

  13. Chang, P. L., Varey, P. A., Rosa, N. E., Ameen, M., Davidson, R. G. Association of steroid sulfatase with one of the arylsulfatase C isozymes in human fibroblasts. J. Biol. Chem. 261: 14443-14447, 1986. [PubMed: 3464600]

  14. Conary, J. T., Lorkowski, G., Schmidt, B., Pohlmann, R., Nagel, G., Meyer, H. E., Krentler, C., Cully, J., Hasilik, A., von Figura, K. Genetic heterogeneity of steroid sulfatase deficiency revealed with cDNA for human steroid sulfatase. Biochem. Biophys. Res. Commun. 144: 1010-1017, 1987. [PubMed: 3034252] [Full Text: https://doi.org/10.1016/s0006-291x(87)80064-5]

  15. Cooper, D. W., McAllan, B. M., Donald, J. A., Dawson, G., Dobrovic, A., Marshall Graves, J. A. Steroid sulphatase is not detected on the X chromosome of Australian marsupials. (Abstract) Cytogenet. Cell Genet. 37: 439, 1984.

  16. Craig, I. W., Tolley, E. Steroid sulphatase and the conservation of mammalian X chromosomes. Trends Genet. 2: 201-204, 1986.

  17. Curry, C. J. R., Magenis, R. E., Brown, M., Lanman, J. T., Jr., Tsai, J., O'Lague, P., Goodfellow, P., Mohandas, T., Bergner, E. A., Shapiro, L. J. Inherited chondrodysplasia punctata due to a deletion of the terminal short arm of an X chromosome. New Eng. J. Med. 311: 1010-1015, 1984. [PubMed: 6482910] [Full Text: https://doi.org/10.1056/NEJM198410183111603]

  18. Eicher, E. M. Personal Communication. Bar Harbor, Me. 1974.

  19. Gant, N. F., Milewich, L., Calvert, M. E., MacDonald, P. C. Steroid sulfatase activity in human fetal membranes. J. Clin. Endocr. 45: 965-972, 1977. [PubMed: 144740] [Full Text: https://doi.org/10.1210/jcem-45-5-965]

  20. Gartler, S. M., Andina, R. J. Mammalian X-chromosome inactivation. Adv. Hum. Genet. 7: 99-140, 1976. [PubMed: 797247] [Full Text: https://doi.org/10.1007/978-1-4757-0659-8_3]

  21. Gartler, S. M., Rivest, M. Evidence for X-linkage of steroid sulfatase in the mouse: steroid sulfatase levels in oocytes of XX and XO mice. Genetics 103: 137-141, 1983. [PubMed: 6573292] [Full Text: https://doi.org/10.1093/genetics/103.1.137]

  22. Goodfellow, P. N., Tippett, P. A human quantitative polymorphism related to Xg blood groups. Nature 289: 404-409, 1981. [PubMed: 7464908] [Full Text: https://doi.org/10.1038/289404a0]

  23. Kawano, J., Kotani, T., Ohtaki, S., Mimamino, N., Matsuo, H., Oinuma, T., Aikawa, E. Characterization of rat and human steroid sulfatases. Biochim. Biophys. Acta 997: 199-205, 1989. [PubMed: 2765556] [Full Text: https://doi.org/10.1016/0167-4838(89)90187-8]

  24. Keitges, E., Gartler, S. M. Dosage of the Sts gene in the mouse. Am. J. Hum. Genet. 39: 470-476, 1986. [PubMed: 3464202]

  25. Keitges, E., Rivest, M., Siniscalco, M., Gartler, S. M. X-linkage of steroid sulphatase in the mouse is evidence for a functional Y-linked allele. Nature 315: 226-227, 1985. [PubMed: 3858674] [Full Text: https://doi.org/10.1038/315226a0]

  26. Kipling, D., Salido, E. C., Shapiro, L. J., Cooke, H. J. High frequency de novo alterations in the long-range genomic structure of the mouse pseudoautosomal region. Nature Genet. 13: 78-82, 1996. [PubMed: 8673108] [Full Text: https://doi.org/10.1038/ng0596-78]

  27. Laval, S. H., Boyd, Y. Partial inversion of gene order within a homologous segment on the X chromosome. Mammalian Genome 4: 119-123, 1993. [PubMed: 8431636] [Full Text: https://doi.org/10.1007/BF00290437]

  28. Li, X. M., Salido, E. C., Gong, Y., Kitada, K., Serikawa, T., Yen, P. H., Shapiro, L. J. Cloning of the rat steroid sulfatase gene (Sts), a non-pseudoautosomal X-linked gene that undergoes X inactivation. Mammalian Genome 7: 420-424, 1996. [PubMed: 8662223] [Full Text: https://doi.org/10.1007/s003359900125]

  29. Lyon, M. F., Peters, J., Glenister, P. H., Ball, S., Wright, E. The scurfy mouse mutant has previously unrecognized hematological abnormalities and resembles Wiskott-Aldrich syndrome. Proc. Nat. Acad. Sci. 87: 2433-2437, 1990. [PubMed: 2320565] [Full Text: https://doi.org/10.1073/pnas.87.7.2433]

  30. Migeon, B. R., Shapiro, L. J., Norum, R. A., Mohandas, T., Axelman, J., Dabora, R. L. Differential expression of steroid sulphatase locus on active and inactive human X chromosome. Nature 299: 838-840, 1982. [PubMed: 6957717] [Full Text: https://doi.org/10.1038/299838a0]

  31. Mohandas, T. K., Stern, H. J., Meeker, C. A., Passage, M. B., Muller, U., Page, D. C., Yen, P. H., Shapiro, L. J. Steroid sulfatase gene in XX males. Am. J. Hum. Genet. 46: 369-376, 1990. [PubMed: 2301402]

  32. Mohandas, T., Shapiro, L. J., Sparkes, R. S., Sparkes, M. C. Regional assignment of the steroid sulfatase--X-linked ichthyosis locus: implications for a non-inactivated region on the short arm of the human X-chromosome. Proc. Nat. Acad. Sci. 76: 5779-5783, 1979. [PubMed: 293682] [Full Text: https://doi.org/10.1073/pnas.76.11.5779]

  33. Mohandas, T., Sparkes, R. S., Hellkuhl, B., Grzeschik, K. H., Shapiro, L. J. Expression of an X-linked gene from an inactive human X chromosome in mouse-human hybrid cells: further evidence for the noninactivation of the steroid sulfatase locus in man. Proc. Nat. Acad. Sci. 77: 6759-6763, 1980. [PubMed: 6935682] [Full Text: https://doi.org/10.1073/pnas.77.11.6759]

  34. Monroe, D. G., Chang, P. L. Tissue-specific expression of human arylsulfatase-C isozymes and steroid sulfatase. Am. J. Hum. Genet. 40: 102-114, 1987. [PubMed: 3471087]

  35. Muller, C. R., Migl, B., Ropers, H.-H. X-linked steroid sulfatase: evidence for different gene-dosage in males and females. Hum. Genet. 54: 197-199, 1980. [PubMed: 6930360] [Full Text: https://doi.org/10.1007/BF00278971]

  36. Muller, C. R., Wahlstrom, J., Ropers, H.-H. Further evidence for the assignment of the steroid sulfatase X-linked ichthyosis locus to the telomer of Xp. (Letter) Hum. Genet. 58: 446, 1981. [PubMed: 6948769] [Full Text: https://doi.org/10.1007/BF00282842]

  37. Nomura, K., Nakano, H., Umeki, K., Harada, K., Kon, A., Tamai, K., Sawamura, D., Hashimoto, I. A study of the steroid sulfatase gene in families with X-linked ichthyosis using polymerase chain reaction. Acta Derm. Venereol. 75: 340-342, 1995. [PubMed: 8615047] [Full Text: https://doi.org/10.2340/0001555575340342]

  38. Ropers, H.-H., Migl, B., Zimmer, J., Fraccaro, M., Maraschio, P. P., Westerveld, A. Activity of steroid sulfatase in fibroblasts with numerical and structural X chromosome aberrations. Hum. Genet. 57: 354-356, 1981. [PubMed: 6945285] [Full Text: https://doi.org/10.1007/BF00281683]

  39. Ropers, H.-H., Wiberg, U. Evidence for X-linkage and non-inactivation of steroid sulphatase locus in wood lemming. Nature 296: 766-767, 1982. [PubMed: 7040982] [Full Text: https://doi.org/10.1038/296766a0]

  40. Salido, E. C., Li, X. M., Yen, P. H., Martin, N., Mohandas, T. K., Shapiro, L. J. Cloning and expression of the mouse pseudoautosomal steroid sulphatase gene (Sts). Nature Genet. 13: 83-86, 1996. [PubMed: 8673109] [Full Text: https://doi.org/10.1038/ng0596-83]

  41. Shankaran, R., Ameen, M., Daniel, W. L., Davidson, R. G., Chang, P. L. Characterization of arylsulfatase C isozymes from human liver and placenta. Biochim. Biophys. Acta 1078: 251-257, 1991. [PubMed: 2065092] [Full Text: https://doi.org/10.1016/0167-4838(91)90566-i]

  42. Shapiro, L. J., Mohandas, T., Weiss, R., Romeo, G. Non-activation of a X-chromosome locus in man. Science 204: 1224-1226, 1979. [PubMed: 156396] [Full Text: https://doi.org/10.1126/science.156396]

  43. Shapiro, L. J., Yen, P., Pomerantz, D., Martin, E., Rolewic, L., Mohandas, T. Molecular studies of deletions at the human steroid sulfatase locus. Proc. Nat. Acad. Sci. 86: 8477-8481, 1989. [PubMed: 2813406] [Full Text: https://doi.org/10.1073/pnas.86.21.8477]

  44. Stein, C., Hille, A., Seidel, J., Rijnbout, S., Waheed, A., Schmidt, B., Geuze, H., von Figura, K. Cloning and expression of human steroid-sulfatase: membrane topology, glycosylation, and subcellular distribution in BHK-21 cells. J. Biol. Chem. 264: 13865-13872, 1989. [PubMed: 2668275]

  45. Toder, R., Rappold, G. A., Schiebel, K., Schempp, W. ANT3 and STS are autosomal in prosimian lemurs: implications for the evolution of the pseudoautosomal region. Hum. Genet. 95: 22-28, 1995. [PubMed: 7814020] [Full Text: https://doi.org/10.1007/BF00225068]

  46. Vogel, W., Grompe, M., Storz, R., Pentz, S. A comparative study on steroid sulfatase and arylsulfatase C in fibroblast clones from 45,X/47,XXX and 69,XXY. Hum. Genet. 66: 367-369, 1984. [PubMed: 6586638] [Full Text: https://doi.org/10.1007/BF00287644]

  47. Willard, H. F., Holmes, M. T. A sensitive and dependable assay for distinguishing hamster and human X-linked steroid sulfatase activity in somatic cell hybrids. Hum. Genet. 66: 272-275, 1984. [PubMed: 6585346] [Full Text: https://doi.org/10.1007/BF00286615]

  48. Wirth, B., Herrmann, F. H., Neugebauer, M., Gillard, E. F., Wulff, K., Stein, C., von Figura, K., Ferguson-Smith, M. A., Gal, A. Linkage analysis in X-linked ichthyosis (steroid sulfatase deficiency). Hum. Genet. 80: 191-192, 1988. [PubMed: 3169744] [Full Text: https://doi.org/10.1007/BF00702868]

  49. Yen, P. H., Allen, E., Marsh, B., Mohandas, T., Wang, N., Taggart, R. T., Shapiro, L. J. Cloning and expression of steroid sulfatase cDNA and the frequent occurrence of deletions in STS deficiency: implications for X-Y interchange. Cell 49: 443-454, 1987. [PubMed: 3032454] [Full Text: https://doi.org/10.1016/0092-8674(87)90447-8]

  50. Yen, P. H., Marsh, B., Allen, E., Tsai, S. P., Ellison, J., Connolly, L., Neiswanger, K., Shapiro, L. J. The human X-linked steroid sulfatase gene and a Y-encoded pseudogene: evidence for an inversion of the Y chromosome during primate evolution. Cell 55: 1123-1135, 1988. [PubMed: 3203382] [Full Text: https://doi.org/10.1016/0092-8674(88)90257-7]


Creation Date:
Cassandra L. Kniffin : 10/28/2008

Edit History:
carol : 08/04/2016
carol : 11/03/2008
ckniffin : 11/3/2008
carol : 10/31/2008
ckniffin : 10/31/2008