1
|
Zwickl P, Fabry S, Bogedain C, Haas A, Hensel R. Glyceraldehyde-3-phosphate dehydrogenase from the hyperthermophilic archaebacterium Pyrococcus woesei: characterization of the enzyme, cloning and sequencing of the gene, and expression in Escherichia coli. J Bacteriol 1990; 172:4329-38. [PMID: 2165475 PMCID: PMC213258 DOI: 10.1128/jb.172.8.4329-4338.1990] [Citation(s) in RCA: 137] [Impact Index Per Article: 3.9] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 12/30/2022] Open
Abstract
The glyceraldehyde-3-phosphate dehydrogenase from the hyperthermophilic archaebacterium Pyrococcus woesei (optimal growth temperature, 100 to 103 degrees C) was purified to homogeneity. This enzyme was strictly phosphate dependent, utilized either NAD+ or NADP+, and was insensitive to pentalenolactone like the enzyme from the methanogenic archaebacterium Methanothermus fervidus. The enzyme exhibited a considerable thermostability, with a 44-min half-life at 100 degrees C. The amino acid sequence of the glyceraldehyde-3-phosphate dehydrogenase from P. woesei was deduced from the nucleotide sequence of the coding gene. Compared with the enzyme homologs from mesophilic archaebacteria (Methanobacterium bryantii, Methanobacterium formicicum) and an extremely thermophilic archaebacterium (Methanothermus fervidus), the primary structure of the P. woesei enzyme exhibited a strikingly high proportion of aromatic amino acid residues and a low proportion of sulfur-containing residues. The coding gene of P. woesei was expressed at a high level in Escherichia coli, thus providing an ideal basis for detailed structural and functional studies of that enzyme.
Collapse
|
research-article |
35 |
137 |
2
|
Scholz S, Sonnenbichler J, Schäfer W, Hensel R. Di-myo-inositol-1,1'-phosphate: a new inositol phosphate isolated from Pyrococcus woesei. FEBS Lett 1992; 306:239-42. [PMID: 1633880 DOI: 10.1016/0014-5793(92)81008-a] [Citation(s) in RCA: 125] [Impact Index Per Article: 3.8] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 12/28/2022]
Abstract
A new inositol derivative could be isolated from the Archaeum Pyrococcus woesei and identified as di-myo-inositol-1,1'-phosphate by 1H, 31P NMR spectroscopy, mass spectrometry and thin layer chromatography. In P. woesei, this inositol phosphate represents the dominant counterion of K+ which ranges from 500 to 600 mM. The role of the potassium salt of di-myo-inositol-1,1'-phosphate as thermostabilizer is discussed.
Collapse
|
|
33 |
125 |
3
|
Michalke K, Wickenheiser EB, Mehring M, Hirner AV, Hensel R. Production of volatile derivatives of metal(loid)s by microflora involved in anaerobic digestion of sewage sludge. Appl Environ Microbiol 2000; 66:2791-6. [PMID: 10877769 PMCID: PMC92074 DOI: 10.1128/aem.66.7.2791-2796.2000] [Citation(s) in RCA: 113] [Impact Index Per Article: 4.5] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/20/2022] Open
Abstract
Gases released from anaerobic wastewater treatment facilities contain considerable amounts of volatile methyl and hydride derivatives of metals and metalloids, such as arsine (AsH(3)), monomethylarsine, dimethylarsine, trimethylarsine, trimethylbismuth (TMBi), elemental mercury (Hg(0)), trimethylstibine, dimethyltellurium, and tetramethyltin. Most of these compounds could be shown to be produced by pure cultures of microorganisms which are representatives of the anaerobic sewage sludge microflora, i.e., methanogenic archaea (Methanobacterium formicicum, Methanosarcina barkeri, Methanobacterium thermoautotrophicum), sulfate-reducing bacteria (Desulfovibrio vulgaris, D. gigas), and a peptolytic bacterium (Clostridium collagenovorans). Additionally, dimethylselenium and dimethyldiselenium could be detected in the headspace of most of the pure cultures. This is the first report of the production of TMBi, stibine, monomethylstibine, and dimethylstibine by a pure culture of M. formicicum.
Collapse
|
research-article |
25 |
113 |
4
|
Walden H, Bell GS, Russell RJ, Siebers B, Hensel R, Taylor GL. Tiny TIM: a small, tetrameric, hyperthermostable triosephosphate isomerase. J Mol Biol 2001; 306:745-57. [PMID: 11243785 DOI: 10.1006/jmbi.2000.4433] [Citation(s) in RCA: 110] [Impact Index Per Article: 4.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/22/2022]
Abstract
Comparative structural studies on proteins derived from organisms with growth optima ranging from 15 to 100 degrees C are beginning to shed light on the mechanisms of protein thermoadaptation. One means of sustaining hyperthermostability is for proteins to exist in higher oligomeric forms than their mesophilic homologues. Triosephosphate isomerase (TIM) is one of the most studied enzymes, whose fold represents one of nature's most common protein architectures. Most TIMs are dimers of approximately 250 amino acid residues per monomer. Here, we report the 2.7 A resolution crystal structure of the extremely thermostable TIM from Pyrococcus woesei, a hyperthermophilic archaeon growing optimally at 100 degrees C, representing the first archaeal TIM structure. P. woesei TIM exists as a tetramer comprising monomers of only 228 amino acid residues. Structural comparisons with other less thermostable TIMs show that although the central beta-barrel is largely conserved, severe pruning of several helices and truncation of some loops give rise to a much more compact monomer in the small hyperthermophilic TIM. The classical TIM dimer formation is conserved in P. woesei TIM. The extreme thermostability of PwTIM appears to be achieved by the creation of a compact tetramer where two classical TIM dimers interact via an extensive hydrophobic interface. The tetramer is formed through largely hydrophobic interactions between some of the pruned helical regions. The equivalent helical regions in less thermostable dimeric TIMs represent regions of high average temperature factor. The PwTIM seems to have removed these regions of potential instability in the formation of the tetramer. This study of PwTIM provides further support for the role of higher oligomerisation states in extreme thermal stabilisation.
Collapse
|
|
24 |
110 |
5
|
Abstract
Triosephosphate isomerase (TIM) of the hyperthermophilic Archaea Pyrococcus woesei and Methanothermus fervidus have been purified to homogeneity. The enzymes from the two hyperthermophiles represent homo-tetramers of 100 kDa, contrary to all known bacterial and eukaryotic TIMs, which are dimers of 48-60 kDa. Molecular size determination of the TIM from the mesophilic methanogen Methanobacterium bryantii yielded the usual molecular mass of only 57 kDa, indicating that the tetrameric aggregation state does not represent an archaeal feature but rather correlates with thermoadaptation. A similar preference for higher protein aggregates in hyperthermophilic Archaea has previously been demonstrated for 3-phosphoglycerate kinases. The gene of the P. woesei TIM was cloned and sequenced. The archaeal TIM proved to be homologous to its bacterial and eukaryotic pendants. Most strikingly, the deduced protein sequence comprises only 224 residues and thus represents the shortest TIM sequence known as yet. Taking the three-dimensional structure of the eucaryal TIM as a basis, from the shortenings of the chain considerable rearrangements at the bottom of the alpha/beta barrel and at its functionally inactive flank are expected, which are interpreted in terms of the formation of new subunit contacts.
Collapse
|
Comparative Study |
29 |
75 |
6
|
Bröckl G, Behr M, Fabry S, Hensel R, Kaudewitz H, Biendl E, König H. Analysis and nucleotide sequence of the genes encoding the surface-layer glycoproteins of the hyperthermophilic methanogens Methanothermus fervidus and Methanothermus sociabilis. EUROPEAN JOURNAL OF BIOCHEMISTRY 1991; 199:147-52. [PMID: 1712296 DOI: 10.1111/j.1432-1033.1991.tb16102.x] [Citation(s) in RCA: 74] [Impact Index Per Article: 2.2] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 12/28/2022]
Abstract
The genes (slgA) encoding the surface-layer glycoproteins of the hyperthermophilic methanogens Methano-thermus pervidus and Methanothermus sociabilis were cloned and sequenced. The nucleotide sequences of these genes differ at only nine positions, resulting in three amino acid differences. In both organisms, the transcription start site was localized by primer extension analyses. The DNA sequence at this site conforms to the promotor box B motif for promotors of archaea. 24 nucleotides upstream of the transcription start is an A + T-rich region, which closely resembles the consensus box A motif of promoters of methanogens. Ribosome binding sites are exactly complementary to the 3' end of the 16S rRNA of these methanogens. Both slgA genes encode for a precursor of the mature surface-layer protein containing 593 amino acid residues with a putative N-terminal signal sequence of 22 amino acid residues. The deduced protein sequences contain 20 sequon structures representing possible carbohydrate-binding sites. In comparison with other surface-layer proteins, these obtained from the two hyperthermophilic methanogens contain unusually high amounts of isoleucine, asparagine and cysteine residues. Predicted secondary structures have a high content of beta-sheet structure (44%) and only 7% alpha-helix structures.
Collapse
|
|
34 |
74 |
7
|
Hensel R, Mayr U, Stetter KO, Kandler O. Comparative studies of lactic acid dehydrogenases in lactic acid bacteria. I. Purification and kinetics of the allosteric L-lactic acid dehydrogenase from Lactobacillus casei ssp. casei and Lactobacillus curvatus. Arch Microbiol 1977; 112:81-93. [PMID: 14601 DOI: 10.1007/bf00446658] [Citation(s) in RCA: 73] [Impact Index Per Article: 1.5] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 12/12/2022]
Abstract
The stability, pH-dependence and kinetic properties of the Mn2+ and FDP-activated NAD-dependent lactic acid dehydrogenases from Lactobacillus casei ssp. casei (ATCC 393) and L. curvatus (DSM 20010) were studied after the enzymes were purified to homogeneity by affinity chromatography. Both enzymes are virtually unidirectional, catalysing efficiency only the reduction of pyruvate. They are similar with respect to the effector requirement and pH-optimum. They differ, however, in their electrophoretic mobility, heat stability, pH-dependence of the Mn2+ requirement and several kinetic properties. It is suggested that most of these differences are caused by differences of the negative charges in the vicinity of the FDP-binding site or the site responsible for the interaction of the subunits of the enzymatically active oligomeres.
Collapse
|
Comparative Study |
48 |
73 |
8
|
Brunner NA, Brinkmann H, Siebers B, Hensel R. NAD+-dependent glyceraldehyde-3-phosphate dehydrogenase from Thermoproteus tenax. The first identified archaeal member of the aldehyde dehydrogenase superfamily is a glycolytic enzyme with unusual regulatory properties. J Biol Chem 1998; 273:6149-56. [PMID: 9497334 DOI: 10.1074/jbc.273.11.6149] [Citation(s) in RCA: 71] [Impact Index Per Article: 2.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/06/2022] Open
Abstract
The hyperthermophilic archaeum Thermoproteus tenax possesses two glyceraldehyde-3-phosphate dehydrogenases differing in cosubstrate specificity and phosphate dependence of the catalyzed reaction. NAD+-dependent glyceraldehyde-3-phosphate dehydrogenase catalyzes the phosphate-independent irreversible oxidation of D-glyceraldehyde 3-phosphate to 3-phosphoglycerate. The coding gene was cloned, sequenced, and expressed in Escherichia coli. Sequence comparisons showed no similarity to phosphorylating glyceraldehyde-3-phosphate dehydrogenases but revealed a relationship to aldehyde dehydrogenases, with the highest similarity to the subgroup of nonphosphorylating glyceraldehyde-3-phosphate dehydrogenases. The activity of the enzyme is affected by a series of metabolites. All effectors tested influence the affinity of the enzyme for its cosubstrate NAD+. Whereas NADP(H), NADH, and ATP reduce the affinity for the cosubstrate, AMP, ADP, glucose 1-phosphate, and fructose 6-phosphate increase the affinity for NAD+. Additionally, most of the effectors investigated induce cooperativity of NAD+ binding. The irreversible catabolic oxidation of glyceraldehyde 3-phosphate, the control of the enzyme by energy charge of the cell, and the regulation by intermediates of glycolysis and glucan degradation identify the NAD+-dependent glyceraldehyde-3-phosphate dehydrogenase as an integral constituent of glycolysis in T. tenax. Its regulatory properties substitute for those lacking in the reversible nonregulated pyrophosphate-dependent phosphofructokinase in this variant of the Embden-Meyerhof-Parnas pathway.
Collapse
|
Comparative Study |
27 |
71 |
9
|
Siebers B, Klenk HP, Hensel R. PPi-dependent phosphofructokinase from Thermoproteus tenax, an archaeal descendant of an ancient line in phosphofructokinase evolution. J Bacteriol 1998; 180:2137-43. [PMID: 9555897 PMCID: PMC107141 DOI: 10.1128/jb.180.8.2137-2143.1998] [Citation(s) in RCA: 68] [Impact Index Per Article: 2.5] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/07/2023] Open
Abstract
Flux into the glycolytic pathway of most cells is controlled via allosteric regulation of the irreversible, committing step catalyzed by ATP-dependent phosphofructokinase (PFK) (ATP-PFK; EC 2.7.1.11), the key enzyme of glycolysis. In some organisms, the step is catalyzed by PPi-dependent PFK (PPi-PFK; EC 2.7.1.90), which uses PPi instead of ATP as the phosphoryl donor, conserving ATP and rendering the reaction reversible under physiological conditions. We have determined the enzymic properties of PPi-PFK from the anaerobic, hyperthermophilic archaeon Thermoproteus tenax, purified the enzyme to homogeneity, and sequenced the gene. The approximately 100-kDa PPi-PFK from T. tenax consists of 37-kDa subunits; is not regulated by classical effectors of ATP-PFKs such as ATP, ADP, fructose 2,6-bisphosphate, or metabolic intermediates; and shares 20 to 50% sequence identity with known PFK enzymes. Phylogenetic analyses of biochemically characterized PFKs grouped the enzymes into three monophyletic clusters: PFK group I represents only classical ATP-PFKs from Bacteria and Eucarya; PFK group II contains only PPi-PFKs from the genus Propionibacterium, plants, and amitochondriate protists; whereas group III consists of PFKs with either cosubstrate specificity, i.e., the PPi-dependent enzymes from T. tenax and Amycolatopsis methanolica and the ATP-PFK from Streptomyces coelicolor. Comparative analyses of the pattern of conserved active-site residues strongly suggest that the group III PFKs originally bound PPi as a cosubstrate.
Collapse
|
research-article |
27 |
68 |
10
|
Fabry S, Hensel R. Purification and characterization of D-glyceraldehyde-3-phosphate dehydrogenase from the thermophilic archaebacterium Methanothermus fervidus. EUROPEAN JOURNAL OF BIOCHEMISTRY 1987; 165:147-55. [PMID: 3569291 DOI: 10.1111/j.1432-1033.1987.tb11205.x] [Citation(s) in RCA: 64] [Impact Index Per Article: 1.7] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 01/06/2023]
Abstract
The D-glyceraldehyde-3-phosphate dehydrogenase from the extremely thermophilic archaebacterium Methanothermus fervidus was purified and crystallized. The enzyme is a homomeric tetramer (molecular mass of subunits 45 kDa). Partial sequence analysis shows homology to the enzymes from eubacteria and from the cytoplasm of eukaryotes. Unlike these enzymes, the D-glyceraldehyde-3-phosphate dehydrogenase from Methanothermus fervidus reacts with both NAD+ and NADP+ and is not inhibited by pentalenolactone. The enzyme is intrinsically stable up to 75 degrees C. It is stabilized by the coenzyme NADP+ and at high ionic strength up to about 90 degrees C. Breaks in the Arrhenius and Van't Hoff plots indicate conformational changes of the enzyme at around 52 degrees C.
Collapse
|
Comparative Study |
38 |
64 |
11
|
Honka E, Fabry S, Niermann T, Palm P, Hensel R. Properties and primary structure of the L-malate dehydrogenase from the extremely thermophilic archaebacterium Methanothermus fervidus. EUROPEAN JOURNAL OF BIOCHEMISTRY 1990; 188:623-32. [PMID: 2110059 DOI: 10.1111/j.1432-1033.1990.tb15443.x] [Citation(s) in RCA: 62] [Impact Index Per Article: 1.8] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 12/30/2022]
Abstract
L-Malate dehydrogenase from the extremely thermophilic mathanogen Methanothermus fervidus was isolated and its phenotypic properties were characterized. The primary structure of the protein was deducted from the coding gene. The enzyme is a homomeric dimer with a molecular mass of 70 kDa, possesses low specificity for NAD+ or NADP+ and catalyzes preferentially the reduction of oxalacetate. The temperature dependence of the activity as depicted in the Arrhenius and van't Hoff plots shows discontinuities near 52 degrees C, as was found for glyceraldehyde-3-phosphate dehydrogenase from the same organism. With respect to the primary structure, the archaebacterial L-malate dehydrogenase deviates strikingly from the eubacterial and eukaryotic enzymes. The sequence similarity is even lower than that between the L-malate dehydrogenases and L-lactate dehydrogenases of eubacteria and eukaryotes. The phylogenetic meaning of this relationship is discussed.
Collapse
|
Comparative Study |
35 |
62 |
12
|
Hensel R, Zwickl P, Fabry S, Lang J, Palm P. Sequence comparison of glyceraldehyde-3-phosphate dehydrogenases from the three urkingdoms: evolutionary implication. Can J Microbiol 1989; 35:81-5. [PMID: 2497945 DOI: 10.1139/m89-012] [Citation(s) in RCA: 61] [Impact Index Per Article: 1.7] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/01/2023]
Abstract
The primary structure of the glyceraldehyde-3-phosphate dehydrogenase from the archaebacteria shows striking deviation from the known sequences of eubacterial and eukaryotic sequences, despite unequivocal homologies in functionally important regions. Thus, the structural similarity between the eubacterial and eukaryotic enzymes is significantly higher than that between the archaebacterial enzymes and the eubacterial and eukaryotic enzymes. This preferred similarity of eubacterial and eukaryotic glyceraldehyde-3-phosphate dehydrogenase structures does not correspond to the phylogenetic distances among the three urkingdoms as deduced from comparisons of ribosomal ribonucleic acid sequences. Indications will be presented that the closer relationship of the eubacterial and eukaryotic glyceraldehyde-3-phosphate dehydrogenase resulted from a gene transfer from eubacteria to eukaryotes after the segregation of the three urkingdoms.
Collapse
|
Comparative Study |
36 |
61 |
13
|
Hensel R, Laumann S, Lang J, Heumann H, Lottspeich F. Characterization of two D-glyceraldehyde-3-phosphate dehydrogenases from the extremely thermophilic archaebacterium Thermoproteus tenax. EUROPEAN JOURNAL OF BIOCHEMISTRY 1987; 170:325-33. [PMID: 3121324 DOI: 10.1111/j.1432-1033.1987.tb13703.x] [Citation(s) in RCA: 59] [Impact Index Per Article: 1.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 01/04/2023]
Abstract
Thermoproteus tenax possesses two different glyceraldehyde-3-phosphate dehydrogenases, one specific for NADP+ and the other for NAD+. NADP(H) inhibits the NAD+-specific enzyme competetively with respect to NAD+ whereas NAD(H) virtually does not interact with the NADP+-specific enzyme. Both enzymes represent homomeric tetramers with subunit molecular masses of 39 kDa (NADP+-specific enzyme) and 49 kDa (NAD+-specific enzyme), respectively. The NADP+-specific enzyme shows significant homology to the known glyceraldehyde-3-phosphate dehydrogenases from eubacteria and eukaryotes as indicated by partial sequencing. The enzymes are thermostable, the NADP+-specific enzyme with a half-life of 35 min at 100 degrees C, the NAD+-specific enzyme with a half-line of greater than or equal to 20 min at 100 degrees C, depending on the protein concentration. Both enzymes show conformational and functional changes at 60-70 degrees C.
Collapse
|
Comparative Study |
38 |
59 |
14
|
Schramm A, Siebers B, Tjaden B, Brinkmann H, Hensel R. Pyruvate kinase of the hyperthermophilic crenarchaeote Thermoproteus tenax: physiological role and phylogenetic aspects. J Bacteriol 2000; 182:2001-9. [PMID: 10715009 PMCID: PMC101911 DOI: 10.1128/jb.182.7.2001-2009.2000] [Citation(s) in RCA: 57] [Impact Index Per Article: 2.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/20/2022] Open
Abstract
Pyruvate kinase (PK; EC 2.7.1.40) of Thermoproteus tenax was purified to homogeneity, and its coding gene was cloned and expressed in Escherichia coli. It represents a homomeric tetramer with a molecular mass of 49 kDa per subunit. PK exhibits positive binding cooperativity with respect to phosphoenolpyruvate and metal ions such as Mg(2+) and Mn(2+). Heterotropic effects, as commonly found for PKs from bacterial and eucaryal sources, could not be detected. The enzyme does not depend on K(+) ions. Heterotrophically grown cells exhibit specific activity of PK four times higher than autotrophically grown cells. Since the mRNA level of the PK coding gene is also accordingly higher in heterotrophic cells, we conclude that the PK activity is adjusted to growth conditions mainly on the transcript level. The enzymic properties of the PK and the regulation of its expression are discussed with respect to the physiological framework given by the T. tenax-specific variant of the Embden-Meyerhof-Parnas pathway. T. tenax PK shows moderate overall sequence similarity (25 to 40% identity) to its bacterial and eucaryal pendants. Phylogenetic analyses of the known PK sequences result in a dichotomic tree topology that divides the enzymes into two major PK clusters, probably diverged by an early gene duplication event. The phylogenetic divergence is paralleled by a striking phenotypic differentiation of PKs: PKs of cluster I, which occur in eucaryal cytoplasm, some gamma proteobacteria, and low-GC gram-positive bacteria, are only active in the presence of fructose-1,6-bisphosphate or other phosphorylated sugars, whereas PKs of cluster II, found in various bacterial phyla, plastids, and in Archaea, show activity without effectors but are commonly regulated by the energy charge of the cell.
Collapse
|
research-article |
25 |
57 |
15
|
Fabry S, Lang J, Niermann T, Vingron M, Hensel R. Nucleotide sequence of the glyceraldehyde-3-phosphate dehydrogenase gene from the mesophilic methanogenic archaebacteria Methanobacterium bryantii and Methanobacterium formicicum. Comparison with the respective gene structure of the closely related extreme thermophile Methanothermus fervidus. EUROPEAN JOURNAL OF BIOCHEMISTRY 1989; 179:405-13. [PMID: 2492940 DOI: 10.1111/j.1432-1033.1989.tb14568.x] [Citation(s) in RCA: 50] [Impact Index Per Article: 1.4] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 01/01/2023]
Abstract
The genes for glyceraldehyde-3-phosphate dehydrogenase (gap genes) from the mesophilic methanogenic archaebacteria Methanobacterium formicicum and Methanobacterium bryantii were cloned and sequenced. The deduced amino acid sequences show 95% identity to each other and about 70% identity to the glyceraldehyde-3-phosphate dehydrogenase from the thermophilic methanogenic archaebacterium Methanothermus fervidus. Although the sequence similarity between the archaebacterial glyceraldehyde-3-phosphate dehydrogenase and the homologous enzyme of eubacteria and eukaryotes is low, an equivalent secondary-structural arrangement can be deduced from the profiles of the physical parameters hydropathy, chain flexibility and amphipathy. In order to find possible thermophile-specific structural features of the enzyme from M. fervidus, a comparative primary-sequence analysis was performed. Amino acid exchanges leading, to a stabilization of the main-chain conformation, could be found throughout the sequence of the thermophile enzyme. Striking features of the thermophile sequence are the preference for isoleucine, especially in beta-sheets, and a low arginine/lysine ratio of 0.54.
Collapse
|
Comparative Study |
36 |
50 |
16
|
Hess D, Krüger K, Knappik A, Palm P, Hensel R. Dimeric 3-phosphoglycerate kinases from hyperthermophilic Archaea. Cloning, sequencing and expression of the 3-phosphoglycerate kinase gene of Pyrococcus woesei in Escherichia coli and characterization of the protein. Structural and functional comparison with the 3-phosphoglycerate kinase of Methanothermus fervidus. EUROPEAN JOURNAL OF BIOCHEMISTRY 1995; 233:227-37. [PMID: 7588750 DOI: 10.1111/j.1432-1033.1995.227_1.x] [Citation(s) in RCA: 48] [Impact Index Per Article: 1.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 01/26/2023]
Abstract
The gene coding for the 3-phosphoglycerate kinase (EC 2.7.2.3) of Pyrococcus woesei was cloned and sequenced. The gene sequence comprises 1230 bp coding for a polypeptide with the theoretical M(r) of 46,195. The deduced protein sequence exhibits a high similarity (46.1% and 46.6% identity) to the other known archaeal 3-phosphoglycerate kinases of Methanobacterium bryantii and Methanothermus fervidus [Fabry, S., Heppner, P., Dietmaier, W. & Hensel, R. (1990) Gene 91, 19-25]. By comparing the 3-phosphoglycerate kinase sequences of the mesophilic and the two thermophilic Archaea, trends in thermoadaptation were confirmed that could be deduced from comparisons of glyceraldehyde-3-phosphate dehydrogenase sequences from the same organisms [Zwickl, P., Fabry, S., Bogedain, C., Haas, A. & Hensel, R. (1990) J. Bacteriol. 172, 4329-4338]. With increasing temperature the average hydrophobicity and the portion of aromatic residues increases, whereas the chain flexibility as well as the content in chemically labile residues (Asn, Cys) decreases. To study the phenotypic properties of the 3-phosphoglycerate kinases from thermophilic Archaea in more detail, the 3-phosphoglycerate kinase genes from P. woesei and M. fervidus were expressed in Escherichia coli. Comparisons of kinetic and molecular properties of the enzymes from the original organisms and from E. coli indicate that the proteins expressed in the mesophilic host are folded correctly. Besides their higher thermostability according to their origin from hyperthermophilic organisms, both enzymes differ from their bacterial and eucaryotic homologues mainly in two respects. (a) The 3-phosphoglycerate kinases from P. woesei and M. fervidus are homomeric dimers in their native state contrary to all other known 3-phosphoglycerate kinases, which are monomers including the enzyme from the mesophilic Archaeum M. bryantii. (b) Monovalent cations are essential for the activity of both archaeal enzymes with K+ being significantly more efficient than Na+. For the P. woesei enzyme, non-cooperative K+ binding with an apparent Kd (K+) of 88 mM could be determined by kinetic analysis, whereas for the M. fervidus 3-phosphoglycerate kinase the K+ binding is rather complex: from the fitting of the saturation data, non-cooperative binding sites with low selectivity for K+ and Na+ (apparent Kd = 270 mM) and at least three cooperative and highly specific K+ binding sites/subunit are deduced. At the optimum growth temperature of P. woesei (100 degrees C) and M. fervidus (83 degrees C), the 3-phosphoglycerate kinases show half-lives of inactivation of only 28 min and 44 min, respectively.(ABSTRACT TRUNCATED AT 400 WORDS)
Collapse
|
Comparative Study |
30 |
48 |
17
|
Siebers B, Wendisch VF, Hensel R. Carbohydrate metabolism in Thermoproteus tenax: in vivo utilization of the non-phosphorylative Entner-Doudoroff pathway and characterization of its first enzyme, glucose dehydrogenase. Arch Microbiol 1997; 168:120-7. [PMID: 9238103 DOI: 10.1007/s002030050477] [Citation(s) in RCA: 46] [Impact Index Per Article: 1.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/04/2023]
Abstract
Thermoproteus tenax is a hyperthermophilic, facultative heterotrophic archaeum. In this organism the utilization of the two catabolic pathways, a variant of the Embden-Meyerhof-Parnas (EMP) pathway and the modified (nonphosphorylative) Entner-Doudoroff (ED) pathway, was investigated and the first enzyme of the ED pathway, glucose dehydrogenase, was characterized. The distribution of the 13C label in alanine synthesized by cells grown with [1-13C]glucose indicated that in vivo the EMP pathway and the modified ED pathway operate parallel, with glucose metabolization via the EMP pathway being prominent. To initiate studies on the regulatory mechanisms governing carbon flux via these pathways, the first enzyme of the ED pathway, glucose dehydrogenase, was purified to homogeneity and its phenotypic properties were characterized. The pyridine-nucleotide-dependent enzyme used both NAD+ and NADP+ as cosubstrates, showing a 100-fold higher affinity for NADP+. Besides glucose, xylose was used as substrate, but with significantly lower affinity. These data suggest that the physiological function of the enzyme is the oxidation of glucose by NADP+. A striking feature was the influence of NADP+ and NAD+ on the quaternary structure and activity state of the enzyme. Without cosubstrate, the enzyme was highly aggregated (mol. mass > 600 kDa) but inactive, whereas in the presence of the cosubstrate the aggregates dissociated into enzymatically active, homomeric dimers with a mol. mass of 84 kDa (mol. mass of subunits: 41 kDa). The N-terminal amino acid sequence showed striking similarity to the respective partial sequences of alcohol dehydrogenases and sorbitol dehydrogenases, but no resemblance to the known pyridine-nucleotide-dependent archaeal and bacterial glucose dehydrogenases.
Collapse
|
Comparative Study |
28 |
46 |
18
|
Grümping R, Michalke K, Hirner AV, Hensel R. Microbial degradation of octamethylcyclotetrasiloxane. Appl Environ Microbiol 1999; 65:2276-8. [PMID: 10224038 PMCID: PMC91335 DOI: 10.1128/aem.65.5.2276-2278.1999] [Citation(s) in RCA: 45] [Impact Index Per Article: 1.7] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/20/2022] Open
Abstract
The microbial degradation of low-molecular-weight polydimethylsiloxanes was investigated through laboratory experiments. Octamethylcyclotetrasiloxane was found to be biodegraded under anaerobic conditions in composted sewage sludge, as monitored by the occurrence of the main polydimethylsiloxane degradation product, dimethylsilanediol, compared to that found in experiments with sterilized control samples.
Collapse
|
research-article |
26 |
45 |
19
|
Hensel R, Mayr U, Yang CY. The complete primary structure of the allosteric L-lactate dehydrogenase from Lactobacillus casei. EUROPEAN JOURNAL OF BIOCHEMISTRY 1983; 134:503-11. [PMID: 6411465 DOI: 10.1111/j.1432-1033.1983.tb07595.x] [Citation(s) in RCA: 42] [Impact Index Per Article: 1.0] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 01/20/2023]
Abstract
The polypeptide chain of the allosteric L-lactate dehydrogenase (EC 1.1.1.27) of Lactobacillus casei consists of 325 amino acid residues. Despite the strikingly different enzymatic characteristics of the allosteric L-lactate dehydrogenase of L. casei and of the non-allosteric vertebrate enzymes, the sequence of the allosteric enzyme shows a distinct homology with that of the non-allosteric vertebrate enzymes (average identity: 37%). An especially high sequence homology can be identified within the active center (average identity: 70%). A clear deviation of the L. casei enzyme from the vertebrate enzyme is the lack of the first 12 amino acid residues at the N terminus and an additional 7 amino acid residues at the C terminus. The localization of the binding site of the allosteric effector D-fructose 1,6-bisphosphate and pH and effector-induced changes of the spectroscopic properties are discussed on the basis of the primary structure.
Collapse
|
|
42 |
42 |
20
|
Fabry S, Hensel R. Primary structure of glyceraldehyde-3-phosphate dehydrogenase deduced from the nucleotide sequence of the thermophilic archaebacterium Methanothermus fervidus. Gene X 1988; 64:189-97. [PMID: 2841192 DOI: 10.1016/0378-1119(88)90334-4] [Citation(s) in RCA: 35] [Impact Index Per Article: 0.9] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/02/2023] Open
Abstract
The gene for the glycolytic enzyme glyceraldehyde-3-phosphate dehydrogenase (GAPDH) from the thermophilic methanogenic archaebacterium Methanothermus fervidus (growth optimum at 84 degrees C) was cloned in Escherichia coli and the nucleotide sequence was determined. A striking preference for adenine and thymidine bases was found in the gene, which is in agreement with the low G + C content of the M. fervidus DNA. The deduced amino acid sequence indicates an Mr of 37,500 for the protein subunit. Alignment with the amino acid sequences of GAPDHs from other organisms shows that the archaebacterial GAPDH is homologous to the respective eubacterial and eukaryotic enzymes, but the similarity between the archaebacterial enzyme and the eubacterial or eukaryotic GAPDHs is much less than that between the latter two.
Collapse
|
|
37 |
35 |
21
|
Brunner NA, Siebers B, Hensel R. Role of two different glyceraldehyde-3-phosphate dehydrogenases in controlling the reversible Embden-Meyerhof-Parnas pathway in Thermoproteus tenax: regulation on protein and transcript level. Extremophiles 2001; 5:101-9. [PMID: 11354453 DOI: 10.1007/s007920100181] [Citation(s) in RCA: 33] [Impact Index Per Article: 1.4] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 10/27/2022]
Abstract
The hyperthermophilic archaeum Thermoproteus tenax uses a variant of the Embden-Meyerhof-Parnas (EMP) pathway as the main route for carbohydrate metabolism. This variant is characterized by a reversible nonallosteric PPi-dependent phosphofructokinase and two glyceraldehyde-3-phosphate dehydrogenases differing in cosubstrate specificity, phosphate dependence, and allosteric behavior. Although the nonphosphorylating NAD+-dependent glyceraldehyde-3-phosphate dehydrogenase (GAPN; E.C. 1.2.1.8) fulfills exclusively catabolic purposes, the phosphorylating NADP+-dependent glyceraldehyde-3-phosphate dehydrogenase (NADP+-GAPDH; E.C. 1.2.1.13) exhibits anabolic features. The gene encoding the NADP+-GAPDH was cloned, sequenced, and expressed in Escherichia coli. The deduced protein sequence displayed 47%-53% sequence identity to archaeal phosphorylating GAPDHs. The kinetic parameters of the NADP+-GAPDH showed a clear preference for the reductive reaction with a 5-fold-higher specific activity in the reductive reaction as compared to the oxidative reaction and a 20-fold-lower Km for 1,3-bisphosphoglycerate as compared to glyceraldehyde-3-phosphate. Contrary to GAPN, the enzyme is not allosterically regulated. The coding gene overlaps by 1 bp with a preceding open reading frame coding for 3-phosphoglycerate kinase (PGK; E.C. 2.7.2.3). Northern analyses identified mono- and bicistronic messages of both genes in an equimolar ratio. Transcript levels and specific activity of NADP+-GAPDH and PGK were 3- to 4-fold higher under autotrophic conditions as compared to heterotrophic conditions, whereas transcript abundance and specific activity of GAPN remained constant in autotrophically and heterotrophically grown cells. The different regulation of the two counteracting glyceraldehyde-3-phosphate dehydrogenases is discussed with respect to the flux control of the T. tenax-specific EMP variant.
Collapse
|
Research Support, Non-U.S. Gov't |
24 |
33 |
22
|
Hensel R, Mayr U, Fujiki H, Kandler O. Comparative studies of lactate dehydrogenases in lactic acid bacteria. Amino-acid composition of an active-site region and chemical properties of the L-lactate dehydrogenase of Lactobacillus casei, Lactobacillus curvatus, Lactobacillus plantarum, and Lactobacillus acidophilus. EUROPEAN JOURNAL OF BIOCHEMISTRY 1977; 80:83-92. [PMID: 411654 DOI: 10.1111/j.1432-1033.1977.tb11859.x] [Citation(s) in RCA: 31] [Impact Index Per Article: 0.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 12/15/2022]
Abstract
The molecular weight, the amino acid composition and the N-terminal and C-terminal amino acids of two allosteric (Lactobacillus casei, L. curvatus) and two non-allosteric (L. plantarum, L. acidophilus) L-lactate dehydrogenases, purified to homogeneity by affinity chromatography, were determined. The amino acid composition of the only tryptic peptide unequivocally common to the fingerprints of the 4 enzymes is virtually identical with that of the arginine peptide, called Arg6 of the the substratebinding site of the L-lactate dehydrogenase dehydrogenase of several animals. However, the 'essential' cysteine residue 165 is replaced by threonine, as it is in the L-lactate dehydrogenase of lobster. In addition, the 4 bacterial peptides differ by one or two changes in single amino acid residues from each other as well as from those of animals. The data indicate that not only the animal L-lactate dehydrogenases, but also the allosteric and lactate dehydrogenases from bacterial sources may have evolved from a common gene.
Collapse
|
Comparative Study |
48 |
31 |
23
|
Argenti D, Jensen BK, Hensel R, Bordeaux K, Schleimer R, Bickel C, Heald D. A mass balance study to evaluate the biotransformation and excretion of [14C]-triamcinolone acetonide following oral administration. J Clin Pharmacol 2000; 40:770-80. [PMID: 10883419 DOI: 10.1177/00912700022009413] [Citation(s) in RCA: 31] [Impact Index Per Article: 1.2] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/15/2022]
Abstract
The principle objective of this study was to characterize the absorption, metabolism, and disposition of orally administered [14C]-triamcinolone acetonide. Six healthy male subjects each received a single 100 microCi (approximately 800 micrograms) oral dose of [14C]-triamcinolone acetonide. Plasma, urine, and fecal samples were collected at selected times and analyzed for triamcinolone acetonide and [14C]-derived radioactivity. Plasma protein binding of triamcinolone acetonide was also determined. Metabolite profiling and identification were carried out in plasma and excreta. Principle metabolites were assessed for activity with in vitro anti-inflammatory models. [14C]-triamcinolone acetonide was found to be systemically absorbed following oral administration. The presystemic metabolism and clearance of triamcinolone acetonide were extensive, with only a small fraction of the total plasma radioactivity being made up of triamcinolone acetonide. Little to no parent compound was detected in the plasma 24 hours after administration. Most of the urinary and fecally [14C]-derived radioactivity was also excreted within 24 and 72 hours postdose, respectively. Mean plasma protein binding of triamcinolone acetonide was constant, predictable, and a relatively low 68% over a 24-fold range of plasma concentrations. Three principle metabolites of triamcinolone acetonide were profiled in plasma, urine, and feces. These metabolites were identified as 6 beta-hydroxy triamcinolone, 21-carboxylic acid triamcinolone acetonide, and 6 beta-hydroxy-21-oic triamcinolone acetonide. All three metabolites failed to show any concentration-dependent effects in anti-inflammatory models evaluating IL-5-sustained eosinophil viability and IgE-induced basophil histamine release.
Collapse
|
Clinical Trial |
25 |
31 |
24
|
Mayr U, Hensel R, Kandler O. Factors affecting the quaternary structure of the allosteric L-lactate dehydrogenase from Lactobacillus casei and Lactobacillus curvatus as investigated by hybridization and ultracentrifugation. EUROPEAN JOURNAL OF BIOCHEMISTRY 1980; 110:527-38. [PMID: 6777160 DOI: 10.1111/j.1432-1033.1980.tb04895.x] [Citation(s) in RCA: 27] [Impact Index Per Article: 0.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 01/21/2023]
Abstract
The allosteric L-lactate dehydrogenases of Lactobacillus curvatus and Lactobacillus casei exist in the tetrameric from (molecular weight about 145 000) at pH 5.0--5.5 even in the absence of the effectors Mn2+ and Fru(1,6)P2 (fructose 1,6-bisphosphate), but undergo reversible dissociation to monomers (molecular weight about 35 000) at higher pH values or in the presence of urea. In the range between pH 5.5 (tetrameric state) and pH 7.4 (monomeric state) the L. curvatus L-lactate dehydrogenase exists in a dissociation-association equilibrium comprising tetramers, dimers and monomers as indicated by the Sc20,w values and the results of hybridization experiments. The simultaneous addition of both effectors [Mn2+ and Fru(1,6)P2] at pH 7.4, however, resulted in the stabilization of the tetrameric form. The addition of Fru(1,6)P2 alone at pH 7.4 had almost no influence on the quarternary structure, whereas the addition of Mn2+, as well as that of NADH, largely prevented dissociation. The L-lactate dehydrogenase of L. casei showed similar properties, although the enzyme dissociates only at about pH greater than or equal to 7.8. As in the case of the L. curvatus enzyme, Fru(1,6)P2 has no influence on the pH-dependent dissociation of the L. casei enzyme, whereas Mn2+ stabilizes the tetrameric structure. Reconstitution of a mixture of the two dissociated enzymes results in the formation of all statistically possible, enzymatically active hybrids. No hybridization between the allosteric enzymes from L. casei and L. curvatus and the non-allosteric ones of Lactobacillus plantarum and Lactobacillus acidophilus was observed.
Collapse
|
|
45 |
27 |
25
|
Lorentzen E, Siebers B, Hensel R, Pohl E. Structure, function and evolution of the Archaeal class I fructose-1,6-bisphosphate aldolase. Biochem Soc Trans 2004; 32:259-63. [PMID: 15046584 DOI: 10.1042/bst0320259] [Citation(s) in RCA: 27] [Impact Index Per Article: 1.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/17/2022]
Abstract
FBPA (fructose-1,6-bisphosphate aldolase) catalyses the reversible aldol condensation of glyceraldehyde 3-phosphate and dihydroxyacetone phosphate to form fructose 1,6-bisphosphate. Two classes of FBPA, which rely on different reaction mechanisms, have so far been discovered, class I mainly found in Eucarya and class II mainly in Bacteria. Only recently were genes encoding proteins with FBPA activity identified in Archaea. Archaeal FBPAs do not share any significant overall sequence identity with members of the traditional classes of FBPAs, raising the interesting question of whether they have evolved independently by convergent evolution or diverged from a common ancestor. Biochemical characterization of FBPAs of the two hyperthermophilic Archaea Thermoproteus tenax and Pyrococcus furiosus showed that the enzymes use a Schiff-base mechanism and thus belong to the class I aldolases. The crystal structure of the archaeal FBPA from T. tenax revealed that the protein fold, as for the classical FBPA I and II, is that of a parallel (betaalpha)(8) barrel. A substrate-bound crystal structure allowed detailed active-site comparisons which showed the conservation of six important catalytic and substrate-binding residues between the archaeal and the classical FBPA I. This observation provides further evidence that the two sequence families of proteins share a common evolutionary origin. Furthermore, structure and sequence analysis indicate that the class I FBPA shares a common evolutionary origin with several other enzyme superfamilies of the (betaalpha)(8) barrel fold.
Collapse
|
Review |
21 |
27 |