Structures important in mammalian 11β- and 17β-hydroxysteroid dehydrogenases

Abstract We have used the X-ray crystallographic structures of rat and human dihydropteridine reductase and Streptomyces hydrogenans 20β-hydroxysteroid dehydrogenase to model parts of the 3-dimensional structure of human 11β- and 17β-hydroxysteroid dehydrogenases. We use this information along with previous results from studies of Drosophila alcohol dehydrogenase mutants to analyze the structures in binding sites for NAD(H) and NADP(H) in 11β-hydroxysteroid dehydrogenase-types 1 and 2. We also examine the structure of an α-helix at catalytic site of 17β-hydroxysteroid dehydrogenase-types 1, 2, 3, and 4. This α-helix contains a highly conserved tyrosine and lysine. Adjacent to the carboxyl side of this lysine is a site proposed to be important in subunit association. We find that 11β- and 17β-hydroxysteroid dehydrogenases-type 1 have the same residues at the “anchor site” and conserve other stabilizing features, despite only 20% sequence identity between their entire sequences. Similar conservation of stabilizing structures is found in the 11β- and 17β-hydroxysteroid dehydrogenases-type 2. We suggest that interactions of the dimerization surface of α-helix F with proteins or membranes may be important in regulating activity of hydroxysteroid dehydrogenases.

[1]  I. Tsigelny,et al.  Adding a positive charge at residue 46 of Drosophila alcohol dehydrogenase increases cofactor specificity for NADP+ , 1994, FEBS letters.

[2]  Michael E. Baker,et al.  Sequence analysis of steroid- and prostaglandin-metabolizing enzymes: Application to understanding catalysis , 1994, Steroids.

[3]  Xenobiotic Ketone Reductase Purification and Properties of an NADPH-dependent Carbonyl Reductase from Human Brain , 1980 .

[4]  M. Baker,et al.  Site-specific mutagenesis of Drosophila alcohol dehydrogenase: evidence for involvement of tyrosine-152 and lysine-156 in catalysis. , 1993, Biochemistry.

[5]  D. Ghosh,et al.  The refined three-dimensional structure of 3α,20β-hydroxysteroid dehydrogenase and possible roles of the residues conserved in short-chain dehydrogenases , 1994 .

[6]  J. L. Napoli,et al.  Cloning of a cDNA for Liver Microsomal Retinol Dehydrogenase , 1995, The Journal of Biological Chemistry.

[7]  G M Maggiora,et al.  Energetics of the structure of the four-alpha-helix bundle in proteins. , 1988, Proceedings of the National Academy of Sciences of the United States of America.

[8]  N. Xuong,et al.  The crystallographic structure of a human dihydropteridine reductase NADH binary complex expressed in Escherichia coli by a cDNA constructed from its rat homologue. , 1994, The Journal of biological chemistry.

[9]  I. Tsigelny,et al.  Amino acids important in enzyme activity and dimer stability for Drosophila alcohol dehydrogenase. , 1995, The Biochemical journal.

[10]  B. Wermuth Purification and properties of an NADPH-dependent carbonyl reductase from human brain. Relationship to prostaglandin 9-ketoreductase and xenobiotic ketone reductase. , 1981, The Journal of biological chemistry.

[11]  Z. Chen,et al.  Role of aspartic acid 38 in the cofactor specificity of Drosophila alcohol dehydrogenase. , 1991, European journal of biochemistry.

[12]  H. Erdjument-Bromage,et al.  Molecular cloning and characterization of (R)-3-hydroxybutyrate dehydrogenase from human heart. , 1992, The Journal of biological chemistry.

[13]  P. Terpstra,et al.  Prediction of the Occurrence of the ADP-binding βαβ-fold in Proteins, Using an Amino Acid Sequence Fingerprint , 1986 .

[14]  C. Branden,et al.  Introduction to protein structure , 1991 .

[15]  F. R. Salemme,et al.  Structural and functional diversity in 4-α-helical proteins , 1980, Nature.

[16]  K. O. Elliston,et al.  Expression cloning and characterization of human 17 beta-hydroxysteroid dehydrogenase type 2, a microsomal enzyme possessing 20 alpha-hydroxysteroid dehydrogenase activity. , 1993, The Journal of biological chemistry.

[17]  Z. Krozowski,et al.  Cloning and tissue distribution of the human 1 lβ-hydroxysteroid dehydrogenase type 2 enzyme , 1994, Molecular and Cellular Endocrinology.

[18]  G. Marfany,et al.  Effect of site‐directed mutagenesis on conserved positions of Drosophila alcohol dehydrogenase , 1993, FEBS letters.

[19]  S. Beverley,et al.  A member of the aldoketo reductase family confers methotrexate resistance in Leishmania. , 1992, The Journal of biological chemistry.

[20]  L. Ribas de Pouplana,et al.  The active site architecture of a short-chain dehydrogenase defined by site-directed mutagenesis and structure modeling. , 1994, Biochemistry.

[21]  H. Jörnvall,et al.  Carboxyethyllysine in a protein: native carbonyl reductase/NADP(+)-dependent prostaglandin dehydrogenase. , 1993, Proceedings of the National Academy of Sciences of the United States of America.

[22]  M Krook,et al.  Characteristics of short-chain alcohol dehydrogenases and related enzymes. , 1991, European journal of biochemistry.

[23]  P. White,et al.  Tyr-179 and Lys-183 are essential for enzymatic activity of 11 beta-hydroxysteroid dehydrogenase. , 1992, Biochemical and biophysical research communications.

[24]  S. Atrian,et al.  Protein engineering of Drosophila alcohol dehydrogenase The hydroxyl group of Tyr152 is involved in the active site of the enzyme , 1992, FEBS letters.

[25]  Michael E. Baker,et al.  Genealogy of regulation of human sex and adrenal function, prostaglandin action, snapdragon and petunia flower colors, antibiotics, and nitrogen fixation: functional diversity from two ancestral dehydrogenases , 1991, Steroids.

[26]  U. Hellman,et al.  The Retinal Pigment Epithelial-specific 11-cis Retinol Dehydrogenase Belongs to the Family of Short Chain Alcohol Dehydrogenases (*) , 1995, The Journal of Biological Chemistry.

[27]  N. Xuong,et al.  Crystal structure of rat liver dihydropteridine reductase. , 1992, Proceedings of the National Academy of Sciences of the United States of America.

[28]  M. Ouellette,et al.  A novel antifolate resistance gene on the amplified H circle of Leishmania. , 1992, The EMBO journal.

[29]  N L Harris,et al.  Four helix bundle diversity in globular proteins. , 1994, Journal of molecular biology.

[30]  K. O. Elliston,et al.  Male pseudohermaphroditism caused by mutations of testicular 17β–hydroxysteroid dehydrogenase 3 , 1994, Nature Genetics.

[31]  A. Lesk,et al.  The relation between the divergence of sequence and structure in proteins. , 1986, The EMBO journal.

[32]  M. Poutanen,et al.  Site-directed mutagenesis of the putative active site of human 17 beta-hydroxysteroid dehydrogenase type 1. , 1994, The Biochemical journal.

[33]  J M Whiteley,et al.  Structural and mechanistic characteristics of dihydropteridine reductase: a member of the Tyr-(Xaa)3-Lys-containing family of reductases and dehydrogenases. , 1994, Proceedings of the National Academy of Sciences of the United States of America.

[34]  M. Baker,et al.  Protochlorophyllide reductase is homologous to human carbonyl reductase and pig 20 beta-hydroxysteroid dehydrogenase. , 1994, The Biochemical journal.

[35]  R. Vihko,et al.  Genomic organization and DNA sequences of human 17β‐hydroxysteroid dehydrogenase genes and flanking regions , 1992 .

[36]  F E Cohen,et al.  Modeling protein structures: construction and their applications , 1993, FASEB journal : official publication of the Federation of American Societies for Experimental Biology.

[37]  Z. Chen,et al.  Site-directed mutagenesis of glycine-14 and two "critical" cysteinyl residues in Drosophila alcohol dehydrogenase. , 1990, Biochemistry.

[38]  H. Tai,et al.  Site-directed mutagenesis of the conserved tyrosine 151 of human placental NAD(+)-dependent 15-hydroxyprostaglandin dehydrogenase yields a catalytically inactive enzyme. , 1991, Biochemical and biophysical research communications.

[39]  A. Place,et al.  Partial correction of structural defects in alcohol dehydrogenase through interallelic complementation in Drosophila melanogaster. , 1987, Genetics.

[40]  F. Leenders,et al.  Molecular cloning and amino acid sequence of the porcine 17β‐estradiol dehydrogenase , 1994 .

[41]  H. Jörnvall,et al.  Three‐dimensional model of NAD+‐dependent 15‐hydroxyprostaglandin dehydrogenase and relationships to the NADP+‐dependent enzyme (carbonyl reductase) , 1993, FEBS letters.

[42]  Z. Krozowski,et al.  11β-Hydroxysteroid dehydrogenase and the short-chain alcohol dehydrogenase (SCAD) superfamily , 1992, Molecular and Cellular Endocrinology.

[43]  J. Greer,et al.  Comparative modeling of homologous proteins. , 1991, Methods in enzymology.

[44]  D. Thatcher The complete amino acid sequence of three alcohol dehydrogenase alleloenzymes (AdhN-11, AdhS and AdhUF) from the fruitfly Drosophila melanogaster. , 1980, The Biochemical journal.

[45]  D. Ghosh,et al.  Three-dimensional structure of holo 3 alpha,20 beta-hydroxysteroid dehydrogenase: a member of a short-chain dehydrogenase family. , 1991, Proceedings of the National Academy of Sciences of the United States of America.

[46]  R. Wierenga,et al.  INTERACTION OF PYROPHOSPHATE MOIETIES WITH ALPHA-HELIXES IN DINUCLEOTIDE BINDING-PROTEINS , 1985 .

[47]  M. Baker,et al.  A common ancestor for human placental 17β‐hydroxysteroid dehydrogenase, Streptomyces coelicolor actIII protein, and Drosophila melanogaster alcohol dehydrogenase , 1990, FASEB journal : official publication of the Federation of American Societies for Experimental Biology.

[48]  W. Trommer,et al.  Coenzyme binding by 3-hydroxybutyrate dehydrogenase, a lipid-requiring enzyme: lecithin acts as an allosteric modulator to enhance the affinity for coenzyme. , 1989, Biochemistry.