Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 20 de 141
Filter
1.
Appl Microbiol Biotechnol ; 65(3): 344-8, 2004 Aug.
Article in English | MEDLINE | ID: mdl-14740194

ABSTRACT

When aluminum (Al) was added to a culture, growth of Rhodotorula glutinis IFO1125 was temporarily arrested, showing longer lag phases, depending on the Al concentrations (50-300 microM) added, but the growth rates were not affected at all. Resistant strains obtained by one round of plate treatment containing Al reverted the resistance level to the wild-type level when cultivated without Al. Repeated Al treatments, however, induced heritable and stable Al resistance, the level of which was increased up to 4,000 microM by stepwise increments in Al concentrations. Thus, the heritable Al resistance adaptively acquired was due neither to adaptation nor to mutation, but to a mechanism which has yet to be studied. Heritable Al resistance seemed to release the Al inhibition of magnesium uptake.


Subject(s)
Aluminum Compounds/pharmacology , Chlorides/pharmacology , Drug Resistance, Fungal , Rhodotorula/drug effects , Adaptation, Physiological , Aluminum Chloride , Culture Media , Magnesium/metabolism , Mutation , Rhodotorula/genetics , Rhodotorula/growth & development
2.
Cell Mol Life Sci ; 60(5): 999-1006, 2003 May.
Article in English | MEDLINE | ID: mdl-12827287

ABSTRACT

The primary structure of nicotinoprotein alcohol dehydrogenase (ADH) from Amycolatopsis methanolica was determined and used for modelling against known ADH structures, and for evaluation of the coenzyme binding. The results establish the medium-chain dehydrogenase/reductase nature of the nicotinoprotein ADH. Its subunit model and that of the human class Ibeta ADH subunit structure are similar, with mean a carbon deviations of 0.95 A, but they differ in seven loops. Nicotinoprotein ADH occupies a phylogenetic position intermediate between the dimeric and tetrameric ADH families. Two of the differing loops are important for coenzyme binding in the nicotinoprotein model, where one (with a Thr271Arg exchange towards the traditional enzyme) may suggest a slight rotation of the coenzyme adenine ring in the nicotinoprotein, and the other, with an Asn288 insertion, may suggest an extra hydrogen bond to its nicotinamide ribose, favouring stronger binding of the coenzyme. Combined with previous data, this suggests differences in the details of the tight coenzyme binding in different nicotinoproteins, but a common mode for this binding by loop differences.


Subject(s)
Actinomycetales/enzymology , Alcohol Dehydrogenase/chemistry , Bacterial Proteins/chemistry , Alcohol Dehydrogenase/physiology , Amino Acid Sequence , Binding Sites , Models, Molecular , Molecular Sequence Data , Phylogeny , Sequence Alignment
3.
J Bacteriol ; 183(22): 6694-8, 2001 Nov.
Article in English | MEDLINE | ID: mdl-11673442

ABSTRACT

Several Sphingomonas spp. utilize polyethylene glycols (PEGs) as a sole carbon and energy source, oxidative PEG degradation being initiated by a dye-linked dehydrogenase (PEG-DH) that oxidizes the terminal alcohol groups of the polymer chain. Purification and characterization of PEG-DH from Sphingomonas terrae revealed that the enzyme is membrane bound. The gene encoding this enzyme (pegA) was cloned, sequenced, and expressed in Escherichia coli. The purified recombinant enzyme was vulnerable to aggregation and inactivation, but this could be prevented by addition of detergent. It is as a homodimeric protein with a subunit molecular mass of 58.8 kDa, each subunit containing 1 noncovalently bound flavin adenine dinucleotide but not Fe or Zn. PEG-DH recognizes a broad variety of primary aliphatic and aromatic alcohols as substrates. Comparison with known sequences revealed that PEG-DH belongs to the group of glucose-methanol-choline (GMC) flavoprotein oxidoreductases and that it is a novel type of flavoprotein alcohol dehydrogenase related (percent identical amino acids) to other, so far uncharacterized bacterial, membrane-bound, dye-linked dehydrogenases: alcohol dehydrogenase from Pseudomonas oleovorans (46%); choline dehydrogenase from E. coli (40%); L-sorbose dehydrogenase from Gluconobacter oxydans (38%); and 4-nitrobenzyl alcohol dehydrogenase from a Pseudomonas species (35%).


Subject(s)
Alcohol Dehydrogenase/metabolism , Alcohol Oxidoreductases/metabolism , Polyethylene Glycols/metabolism , Sphingomonas/enzymology , Alcohol Dehydrogenase/chemistry , Alcohol Dehydrogenase/genetics , Alcohol Oxidoreductases/chemistry , Alcohol Oxidoreductases/genetics , Amino Acid Sequence , Cloning, Molecular , Flavin-Adenine Dinucleotide/chemistry , Molecular Sequence Data , Molecular Weight , Sequence Alignment , Sequence Homology , Sphingomonas/genetics
4.
Acta Crystallogr D Biol Crystallogr ; 57(Pt 11): 1732-4, 2001 Nov.
Article in English | MEDLINE | ID: mdl-11679760

ABSTRACT

Quinohaemoprotein alcohol dehydrogenase from Comamonas testosteroni is a functional electron-transfer protein containing both a haem c and a pyrroloquinoline quinone cofactor. The enzyme has been crystallized at 277 K using polyethylene glycol 6000 as precipitant. The crystals belong to space group C2, with unit-cell parameters a = 98.1, b = 74.3, c = 92.2 A, beta = 105.9 degrees. A native data set with a resolution of 2.44 A resolution has been collected. The approximate orientation of the haem group with respect to the unit-cell axes has been determined from the optical properties of the crystals.


Subject(s)
Alcohol Oxidoreductases/chemistry , Comamonas testosteroni/enzymology , Crystallization , Crystallography, X-Ray , Protein Conformation
5.
Arch Biochem Biophys ; 394(2): 245-54, 2001 Oct 15.
Article in English | MEDLINE | ID: mdl-11594739

ABSTRACT

Benzoate-para-hydroxylase (CYP51A or BpH) and NADPH:cytochrome P450 reductase from the filamentous fungus Aspergillus niger were purified to apparent homogeneity, using an overproducing A. niger strain. This is the first membrane-bound fungal cytochrome P450 to be isolated and characterized. Combining BpH with NADPH:cytochrome P450 oxidoreductase in the presence of the phospholipid dilauryl phosphatidylcholine restored the BpH activity, although to only a minor extent. Spectral analysis of BpH showed characteristic spectra for a cytochrome P450. Substrate binding studies with purified BpH as a function of temperature and as a function of pH were performed. Temperature-dependent studies, at pH 8.0, showed that the simplified spin equilibrium model originally proposed for camphor binding to cytochrome P450cam (M. T. Fisher and S. G. Sligar, 1987, Biochemistry 26, 4797-4803) also applies to the benzoate-BpH system. Two equilibrium constants were determined, K(1) for substrate binding without a spin change and K(2) for the spin change of the benzoate-BpH complex. pH-dependent binding studies showed that both K(1) and K(2) increase with pH, indicative of a higher affinity. As K(1) decreases more strongly with pH than K(2), we suggest that benzoate first binds to a binding site on the outside of the protein in a pH-dependent way, followed by transfer to the inside of the protein causing a spin change at the heme iron. The strong pH dependence of K(1) could be the result of the need to break salt bridges at the binding site on the outside of the protein. pH-dependent kinetic studies with microsomes showed that the apparent K(M) values followed the trend observed for benzoate binding to purified BpH, while k(cat) values were virtually constant between pH 6.6 and 8.0 and decreased above pH 8, probably due to loss of productive interaction between BpH and NADPH:cytochrome P450 oxidoreductase. Research into the substrate specificity of BpH showed that BpH can only use benzoic acid and some of its derivatives. Monosubstitution on the phenyl ring is allowed but only at certain positions with specific, not too large groups. Substitution always leads to a lower affinity of the substrate. With one exception, all substrates were converted to their 4-hydroxy derivative. The exception, 3-methoxybenzoate, was demethylated to yield 3-hydroxybenzoate only. The restricted number of substrates and the specificity in catalysis suggest that BpH is not a general-purpose hydroxylase but that its role is confined to benzoate hydroxylation in the beta-ketoadipate pathway of A. niger.


Subject(s)
Aspergillus niger/enzymology , Cytochrome P-450 Enzyme System/chemistry , Mixed Function Oxygenases/chemistry , Aspergillus niger/chemistry , Benzoate 4-Monooxygenase , Benzoic Acid/metabolism , Binding, Competitive , Cell Membrane/chemistry , Cell Membrane/enzymology , Cytochrome P-450 Enzyme System/isolation & purification , Cytochrome P-450 Enzyme System/metabolism , Enzyme Activation , Hydrogen-Ion Concentration , Isoenzymes/chemistry , Isoenzymes/isolation & purification , Isoenzymes/metabolism , Microsomes/chemistry , Microsomes/enzymology , Mixed Function Oxygenases/isolation & purification , Mixed Function Oxygenases/metabolism , NADPH-Ferrihemoprotein Reductase/chemistry , NADPH-Ferrihemoprotein Reductase/isolation & purification , Phosphatidylcholines/chemistry , Spectrophotometry , Substrate Specificity , Temperature
6.
J Biol Chem ; 276(46): 42923-31, 2001 Nov 16.
Article in English | MEDLINE | ID: mdl-11555656

ABSTRACT

Pseudomonas putida contains an amine dehydrogenase that is called a quinohemoprotein as it contains a quinone and two hemes c as redox active groups. Amino acid sequence analysis of the smallest (8.5 kDa), quinone-cofactor-bearing subunit of this heterotrimeric enzyme encountered difficulties in the interpretation of the results at several sites of the polypeptide chain. As this suggested posttranslational modifications of the subunit, the structural genes for this enzyme were determined and mass spectrometric de novo sequencing was applied to several peptides obtained by chemical or enzymatic cleavage. In agreement with the interpretation of the X-ray electronic densities in the diffraction data for the holoenzyme, our results show that the polypeptide of the small subunit contains four intrachain cross-linkages in which the sulfur atom of a cysteine residue is involved. Two of these cross-linkages occur with the beta-carbon atom of an aspartic acid, one with the gamma-carbon atom of a glutamic acid and the fourth with a tryptophanquinone residue, this adduct constituting the enzyme's quinone cofactor, CTQ. The thioether type bond in all four of these adducts has never been found in other proteins. CTQ is a novel cofactor in the series of the recently discovered quinone cofactors.


Subject(s)
Cysteine/chemistry , Indolequinones , Oxidoreductases Acting on CH-NH Group Donors/chemistry , Pseudomonas putida/enzymology , Sulfides/chemistry , Tryptophan/analogs & derivatives , Amino Acid Sequence , Amino Acids/chemistry , Cloning, Molecular , Cross-Linking Reagents/pharmacology , Glutamic Acid/chemistry , Heme/chemistry , Mass Spectrometry , Models, Chemical , Models, Genetic , Molecular Sequence Data , Open Reading Frames , Oxidation-Reduction , Oxidoreductases Acting on CH-NH Group Donors/genetics , Oxidoreductases Acting on CH-NH Group Donors/metabolism , Peptides/chemistry , Protein Binding , Protein Processing, Post-Translational , Quinones/chemistry , Sequence Analysis, DNA , Sequence Homology, Amino Acid , Spectrometry, Mass, Electrospray Ionization , Tryptophan/chemistry , X-Rays
7.
Biochem J ; 357(Pt 3): 893-8, 2001 Aug 01.
Article in English | MEDLINE | ID: mdl-11463363

ABSTRACT

Spectroelectrochemical studies were performed on the interaction between Ca(2+) and pyrroloquinoline quinone (PQQ) in soluble glucose dehydrogenase (sGDH) and in the free state by applying a mediated continuous-flow column electrolytic spectroelectrochemical technique. The enzyme forms used were holo-sGDH (the holo-form of sGDH from Acinetobacter calcoaceticus) and an incompletely reconstituted form of this, holo-X, in which the PQQ-activating Ca(2+) is lacking. The spectroelectrochemical and ESR data clearly demonstrated the generation of the semiquinone radical of PQQ in holo-sGDH and in the free state in the presence of Ca(2+). In contrast, in the absence of Ca(2+) no semiquinone was observed, either for PQQ in the free state (at pH 7.0) or in the enzyme (holo-X). Incorporation of Ca(2+) into the active site of holo-X, yielding holo-sGDH, caused not only stabilization of the semiquinone form of PQQ but also a negative shift (of 26.5 mV) of the two-electron redox potential, indicating that the effect of Ca(2+) is stronger on the oxidized than on the reduced PQQ. Combining these data with the observations on the kinetic and chemical mechanisms, it was concluded that the strong stimulating effect of Ca(2+) on the activity of sGDH can be attributed to facilitation of certain kinetic steps, and not to improvement of the thermodynamics of substrate oxidation. The consequences of this conclusion are discussed for the oxidative as well as for the reductive part of the reaction of sGDH.


Subject(s)
Benzoquinones/metabolism , Calcium/metabolism , Glucose Dehydrogenases/metabolism , Quinolones/metabolism , Quinones/metabolism , Acinetobacter calcoaceticus/enzymology , Glucose 1-Dehydrogenase , Oxidation-Reduction , PQQ Cofactor , Recombinant Proteins/metabolism
8.
Appl Microbiol Biotechnol ; 55(4): 486-91, 2001 May.
Article in English | MEDLINE | ID: mdl-11398932

ABSTRACT

The non-ligninolytic fungus Cunninghamella elegans was investigated for its ability to produce naphthalene hydroxylase (NAH) and naphthol hydroxylase (NOH) activities under various conditions. When the organism was cultivated on a rich growth medium, the mycelia exhibited significant constitutive NAH activity in the late exponential growth phase, but not in the early-exponential-growth-phase. On incubating the early-exponential-growth-phase mycelia with naphthalene, NAH activity was increased five-fold; however, this increase did not occur in the presence of the protein synthesis inhibitor cycloheximide. Since incubation of the late-phase mycelia with naphthalene did not lead to a higher degradation rate of naphthalene, mycelia in this physiological state have apparently lost the ability to induce synthesis of the enzyme exhibiting NAH activity. This is not due to an overall inability to perform de novo protein synthesis, since NOH activity, non-constitutive at all growth phases, could be induced by incubating late-phase mycelia with naphthalene. Whether inducible and constitutive NAH activity originate from one and the same enzyme remains to be elucidated. It is suggested that naphthalene oxidizing enzyme(s) may also oxidize pyrene, but not anthracene or benzo[a]pyrene, although the latter are degradable by C. elegans.


Subject(s)
Cunninghamella/metabolism , Mixed Function Oxygenases/metabolism , Naphthalenes/metabolism , Biodegradation, Environmental , Chromatography, High Pressure Liquid , Cunninghamella/enzymology , Enzyme Induction , Mixed Function Oxygenases/biosynthesis , Polycyclic Compounds/pharmacology
9.
Enzyme Microb Technol ; 28(2-3): 233-239, 2001 Feb 01.
Article in English | MEDLINE | ID: mdl-11166817

ABSTRACT

It is already known that kinetic resolution of racemic glycidol (2,3-epoxy-1-propanol) takes place when Acetobacter pasteurianus oxidizes the compound to glycidic acid (2,3-epoxy-propionic acid) with glycidaldehyde (2,3-epoxy-propanal) proposed to be the transient seen in this conversion. Since inhibition affects the feasibility of a process based on this conversion in a negative sense, and the chemical reactivity of glycidaldehyde predicts that it could be the cause for the phenomena observed, it is important to know which enzyme(s) oxidise(s) this compound. To study this, rac.- as well as (R)-glycidaldehyde were prepared by chemical synthesis and analytical methods developed for their determination. It appears that purified quinohemoprotein alcohol dehydrogenase (QH-ADH type II), the enzyme responsible for the kinetic resolution of rac.-glycidol, also catalyses the oxidation of glycidaldehyde. In addition, a preparation exhibiting dye-linked aldehyde dehydrogenase activity for acetaldehyde, most probably originating from molybdohemoprotein aldehyde dehydrogenase (ALDH), which has been described for other Acetic acid bacteria, oxidised glycidaldehyde as well with a preference for the (R)-enantiomer, the selectivity quantified by an enantiomeric ratio (E) value of 7. From a comparison of the apparent kinetic parameter values of QH-ADH and ALDH, it is concluded that ALDH is mainly responsible for the removal of glycidaldehyde in conversions of glycidol catalysed by A. pasteurianus cells. It is shown that the transient observed in rac.-glycidol conversion by whole cells, is indeed (R)-glycidaldehyde. Since both QH-ADH and ALDH are responsible for vinegar production from ethanol by Acetobacters, growth and induction conditions optimal for this process seem also suited to yield cells with high catalytic performance with respect to kinetic resolution of glycidol and prevention of formation of inhibitory concentrations glycidaldehyde.

10.
Chem Rec ; 1(1): 74-83, 2001.
Article in English | MEDLINE | ID: mdl-11893060

ABSTRACT

Until recently, it was generally believed that enzymatic oxidation and reduction requires the participation of either a nicotinamide (NAD(P)+) or a flavin (FAD, FMN), in agreement with the existence of NAD(P)/H-dependent dehydrogenases/reductases and flavoprotein dehydrogenases/reductases/oxidases. However, during the past 20 years, the unraveling of the enzymology of the oxidation and reduction of C1-compounds by bacteria has led to the discovery of many new redox cofactors, some of them discussed here as they have a wider physiological significance than just enabling enzymatic C1-conversions to occur. A good example is the quinone cofactors, encompassing PQQ (2,7,9-tricarboxy-1H-pyrrolo[2,3-f]-quinoline-4,5-dione), TTQ (tryptophyl tryptophanquinone), TPQ (topaquinone), LTQ (lysyl topaquinone), and several others whose structures have still to be elucidated. Another example is mycothiol (1-O-(2'-[N-acetyl-L-cysteinyl]amido-2'-deoxy-alpha-D-glucopyranosyl)-D-myo-inosoitol), the counterpart of glutathione, once thought to be a universal coenzyme. Because these novel cofactors assist in reactions that can also be catalyzed by already known enzyme "classic cofactor" combinations, and first indications suggest that the chemistry of the reactions is not unique, one may wonder about the evolutionary background for this cofactor diversity. However, as will be illustrated by examples, from a practical point of view the diversity is beneficial, as it has increased the arsenal of enzymes suitable for application.


Subject(s)
Coenzymes/chemistry , Bacteria/chemistry , Bacteria/enzymology , Glutathione/chemistry , Niacinamide/chemistry , Oxidation-Reduction , PQQ Cofactor , Quinolones/chemistry , Quinones/chemistry
11.
Biochemistry ; 39(31): 9384-92, 2000 Aug 08.
Article in English | MEDLINE | ID: mdl-10924133

ABSTRACT

Spectral and kinetic studies were performed on enzyme forms of soluble glucose dehydrogenase of the bacterium Acinetobacter calcoaceticus (sGDH) in which the PQQ-activating Ca(2+) was absent (Holo X) or was replaced with Ba(2+) (Ba-E) or in which PQQ was replaced with an analogue or a derivative called "nitroPQQ" (E-NPQ). Although exhibiting diminished rates, just like sGDH, all enzyme forms were able to oxidize a broad spectrum of aldose sugars, and their reduced forms could be oxidized with the usual artificial electron acceptor. On inspection of the plots for the reductive half-reaction, it appeared that the enzyme forms exhibited a negative cooperativity effect similar to that of sGDH itself under turnover conditions, supporting the view that simultaneous binding of substrate to the two subunits of sGDH causes the effect. Stopped-flow spectroscopy of the reductive half-reaction of Ba-E with glucose showed a fluorescing transient previously observed in the reaction of sGDH with glucose-1-d, whereas no intermediate was detected at all in the reactions of E-NPQ and Holo X. Using hydrazine as a probe, the fluorescing C5 adduct of PQQ and hydrazine was formed in sGDH, Ba-E, and Holo X, but E-NPQ did not react with hydrazine. When this is combined with other properties of E-NPQ and the behavior of enzyme forms containing a PQQ analogue, we concluded that the catalytic potential of the cofactor in the enzyme is not determined by its adduct-forming ability but by whether it is or can be activated with Ca(2+), activation being reflected by the large red shift of the absorption maximum induced by this metal ion when binding to the reduced cofactor in the enzyme. This conclusion, together with the observed deuterium kinetic isotope effect of 7.8 on transient formation in Ba-E, and that already known on transient decay, indicate that the sequential steps in the mechanism of sGDH must be (1) reversible substrate binding, (2) direct transfer of a hydride ion (reversible or irreversible) from the C1 position of the beta-anomer of glucose to the C5 of PQQ, (3) irreversible, rate-determining tautomerization of the fluorescing, C5-reduced PQQ to PQQH(2) and release (or earlier) of the product, D-glucono-delta-lactone, and (4) oxidation of PQQH(2) by an electron acceptor. The PQQ-activating Ca(2+) greatly facilitates the reactions occurring in step 2. His144 may also play a role in this by acting as a general base catalyst, initiating hydride transfer by abstracting a proton from the anomeric OH group of glucose. The validity of the proposed mechanism is discussed for other PQQ-containing dehydrogenases.


Subject(s)
Calcium/chemistry , Coenzymes/chemistry , Glucose Dehydrogenases/chemistry , Glucose/chemistry , Hydrogen/chemistry , PQQ Cofactor , Quinolines/chemistry , Acinetobacter calcoaceticus/enzymology , Barium/chemistry , Carbon , Catalysis , Deuterium , Dimerization , Electron Transport , Holoenzymes/chemistry , Isomerism , Kinetics , Oxidation-Reduction , Solubility , Spectrometry, Fluorescence , Substrate Specificity
12.
Biofactors ; 10(2-3): 201-6, 1999.
Article in English | MEDLINE | ID: mdl-10609883

ABSTRACT

Glutathione is not a universal coenzyme for formaldehyde oxidation. MySH (mycothiol, 1-O-(2'-[N-acetyl-L-cysteinyl]amido-2'-deoxy-alpha-D-glucopyranosyl)-D-m yo-inositol) is GSH's counterpart as coenzyme in formaldehyde dehydrogenase from certain gram-positive bacteria. However, formaldehyde dissimilation and detoxification not only proceed via thiol-dependent but also via thiol-independent dehydrogenases. The distinct structures and enzymatic properties of MySH-dependent and GSH-dependent formaldehyde dehydrogenases could provide clues for development of selective drugs against pathogenic Mycobacteria. It is to be expected that other new types of thiol-dependent formaldehyde dehydrogenases will be discovered in the future. Indications exist that the product of thiol-dependent formaldehyde oxidation, the thiol formate ester, is not only hydrolytically converted into thiol and formate but can also be oxidatively converted in some cases by a molybdoprotein aldehyde dehydrogenase into the corresponding carbonate ester, decomposing spontaneously into CO2 and the thiol.


Subject(s)
Disaccharides/metabolism , Formaldehyde/metabolism , Gram-Positive Bacteria/metabolism , Pyrazoles , Sulfhydryl Compounds/metabolism , Aldehyde Oxidoreductases/metabolism , Cysteine , Formaldehyde/pharmacokinetics , Glutathione/metabolism , Glycopeptides , Inactivation, Metabolic , Inositol , Mycobacterium/metabolism
13.
EMBO J ; 18(19): 5187-94, 1999 Oct 01.
Article in English | MEDLINE | ID: mdl-10508152

ABSTRACT

Soluble glucose dehydrogenase (s-GDH; EC 1.1.99.17) is a classical quinoprotein which requires the cofactor pyrroloquinoline quinone (PQQ) to oxidize glucose to gluconolactone. The reaction mechanism of PQQ-dependent enzymes has remained controversial due to the absence of comprehensive structural data. We have determined the X-ray structure of s-GDH with the cofactor at 2.2 A resolution, and of a complex with reduced PQQ and glucose at 1.9 A resolution. These structures reveal the active site of s-GDH, and show for the first time how a functionally bound substrate interacts with the cofactor in a PQQ-dependent enzyme. Twenty years after the discovery of PQQ, our results finally provide conclusive evidence for a reaction mechanism comprising general base-catalyzed hydride transfer, rather than the generally accepted covalent addition-elimination mechanism. Thus, PQQ-dependent enzymes use a mechanism similar to that of nicotinamide- and flavin-dependent oxidoreductases.


Subject(s)
Glucose Dehydrogenases/chemistry , Binding Sites , Calcium/metabolism , Crystallography, X-Ray , Glucose/metabolism , Glucose Dehydrogenases/metabolism , Models, Molecular , Oxidation-Reduction , Protein Conformation , Recombinant Proteins/chemistry , Recombinant Proteins/metabolism , Solubility
14.
J Mol Biol ; 289(2): 319-33, 1999 Jun 04.
Article in English | MEDLINE | ID: mdl-10366508

ABSTRACT

The crystal structure of a dimeric apo form of the soluble quinoprotein glucose dehydrogenase (s-GDH) from Acinetobacter calcoaceticus has been solved by multiple isomorphous replacement followed by density modification, and was subsequently refined at 1. 72 A resolution to a final crystallographic R-factor of 16.5% and free R-factor of 20.8% [corrected]. The s-GDH monomer has a beta-propeller fold consisting of six four-stranded anti-parallel beta-sheets aligned around a pseudo 6-fold symmetry axis. The enzyme binds three calcium ions per monomer, two of which are located in the dimer interface. The third is bound in the putative active site, where it may bind and functionalize the pyrroloquinoline quinone (PQQ) cofactor. A data base search unexpectedly showed that four uncharacterized protein sequences are homologous to s-GDH with many residues in the putative active site absolutely conserved. This indicates that these homologs may have a similar structure and that they may catalyze similar PQQ-dependent reactions.A structure-based sequence alignment of the six four-stranded beta-sheets in s-GDH's beta-propeller fold shows an internally conserved sequence repeat that gives rise to two distinct conserved structural motifs. The first structural motif is found at the corner of the short beta-turn between the inner two beta-strands of the beta-sheets, where an Asp side-chain points back into the beta-sheet to form a hydrogen-bond with the OH/NH of a Tyr/Trp side-chain in the same beta-sheet. The second motif involves an Arg/Lys side-chain in the C beta-strand of one beta-sheet, which forms a bidentate salt-bridge with an Asp/Glu in the CD loop of the next beta-sheet. These intra and inter-beta-sheet hydrogen-bonds are likely to contribute to the stability of the s-GDH beta-propeller fold.


Subject(s)
Acinetobacter calcoaceticus/enzymology , Glucose Dehydrogenases/chemistry , Amino Acid Sequence , Apoenzymes/chemistry , Binding Sites , Computer Graphics , Conserved Sequence , Crystallography, X-Ray , Models, Molecular , Molecular Sequence Data , Protein Structure, Secondary , Recombinant Proteins/chemistry , Repetitive Sequences, Amino Acid , Sequence Alignment , Sequence Homology, Amino Acid , Software
15.
J Biosci Bioeng ; 88(3): 231-6, 1999.
Article in English | MEDLINE | ID: mdl-16232604

ABSTRACT

About twenty years ago, the cofactor pyrroloquinoline quinone, PQQ, was discovered. Here the author gives his personal view on the reasons why this cofactor was so lately discovered and how the steps in its identification were made. The discovery not only led to subsequent studies on the physiological significance of PQQ but also initiated investigations on other enzymes where the presence of PQQ was expected, resulting in the discovery of three other quinone cofactors, TPQ, TTQ, and LTQ, which differ from PQQ as they are part of the protein chain of the enzyme to which they belong. Enzymes using quinone cofactors, the so-called quinoproteins, copper-quinoproteins, and quinohemoproteins, are mainly involved in the direct oxidation of alcohols, sugars, and amines. Some of the PQQ-containing ones participate in incomplete bacterial oxidation processes like the conversion of ethanol into vinegar and of D-glucose into (5-keto)gluconic acid. Soluble glucose dehydrogenase is the sensor in diagnostic test strips used for glucose determination in blood samples of diabetic patients. Quinohemoprotein alcohol dehydrogenases have an enantiospecificity suited for the kinetic resolution of racemic alcohols to their enantiomerically pure form, certain enantiomers being interesting candidates as building block for synthesis of high-value-added chemicals. Making up for balance after twenty years of quinoprotein research, the following conclusions can be drawn: since quinoproteins do not catalyze unique reactions, we know now that there are more enzymes which catalyze one and the same reaction than we did before, but do not understand the reason for this (compare e.g. NAD/NADP-dependent glucose dehydrogenases, flavoprotein glucose oxidase/dehydrogenase, and soluble/membrane-bound, PQQ-containing glucose dehydrogenases, enzymes all catalyzing the oxidation of beta-D-glucose to delta-gluconolactone but being quite different from each other); however, taking a pragmatic point of view, the foregoing can also be regarded as a positive development since as illustrated by the examples given above, the enlargement of the catalytic arsenal with quinoprotein enzymes provides in more possibilities for enzyme applications; the hopes that PQQ could be a new vitamin have diminished strongly after it has become clear that its occurrence is restricted to bacteria; the impact factor is broader than just the development of the field of quinoproteins, since together with that of enzymes containing a one-electron oxidized amino acid residue as cofactor, it has emphasized that cofactors not only derive from nucleotides (e.g. FAD, NAD) but also from amino acids. Finally, strong indications exist to assume that this is not the end of the story since other quinone cofactors seem awaiting their discovery.

16.
Biosci Biotechnol Biochem ; 63(1): 10-20, 1999.
Article in English | MEDLINE | ID: mdl-27392873

ABSTRACT

Resting cells of Acetobacter pasteurianus LMG 1635 (ATCC 12874) show appreciable enantioselectivity (E=16-18) in the oxidative kinetic resolution of racemic 2,3-epoxy-1-propanol, glycidol. Distinctly lower values (E=7-9) are observed for the ferricyanide-coupled oxidation of glycidol by the isolated quinohemoprotein alcohol dehydrogenase, QH-ADH, which is responsible for the enantiospecific oxidation step in whole cells. The accuracy of E-values from conversion experiments could be verified using complementary methods for the measurement of enantiomeric ratios. Effects of pH, detergent, the use of artificial electron acceptors, and the presence of intermediate aldehydes, could be accounted for. Measurements of E-values at successive stages of the purification showed that the drop in enantioselectivity correlates with the separation of QH-ADH from the cytoplasmic membrane. It is argued that the native arrangement of QH-ADH in the membrane-associated complex favors the higher E-values. The consequences of these findings for the use of whole cells versus purified enzymes in biocatalytic kinetic resolutions of chiral alcohols are discussed.

17.
Biochemistry ; 37(39): 13854-61, 1998 Sep 29.
Article in English | MEDLINE | ID: mdl-9753475

ABSTRACT

Kinetic and optical studies were performed on the reductive half-reaction of soluble, quinoprotein glucose dehydrogenase (sGDH), i.e., on the conversion of sGDHox plus aldose sugar into sGDHred plus corresponding aldonolactone. It appears that the nature and stereochemical configuration of the substituents at certain positions in the aldose molecule determine the substrate specificity pattern: absolute specificity exists with respect to the C1-position (only sugars being oxidized which have the same configuration of the H/OH substituents at this site as the beta-anomer of glucose, not those with the opposite one) and with respect to the overall conformation of the sugar molecule (sugars with a 4C1 chair conformation are substrates, those with a 1C4 one are not); the nature and configuration of the substituents at the 3-position are hardly relevant for activity, and an equatorial pyranose group at the 4-position exhibits only aspecific hindering of the binding of the aldose moiety of a disaccharide. The pH optimum determined for glucose oxidation appeared to be 7.0, implying that reoxidation of sGDHred is rate-limiting with those electron acceptors displaying a different value under steady-state conditions. The kinetic mechanism of sGDH consists of (a) step(s) in which a fluorescing intermediate is formed, and a subsequent, irreversible step, determining the overall rate of the reductive half-reaction. The consequences of this for the likeliness of chemical mechanisms where glucose is oxidized by covalent catalysis in which a C5-adduct of glucose and PQQ are involved, or by hydride transfer from glucose to PQQ, followed by tautomerization of C5-reduced PQQ to PQQH2, are discussed. The negative cooperative behavior of sGDH seems to be due to substrate-occupation-dependent subunit interaction in the dimeric enzyme molecule, leading to a large increase of the turnover rate under saturating conditions.


Subject(s)
Carbohydrates/chemistry , Glucose Dehydrogenases/chemistry , Aldehyde Reductase/chemistry , Glucose/chemistry , Holoenzymes/chemistry , Kinetics , Oxidation-Reduction , Solubility , Spectrophotometry , Substrate Specificity , Titrimetry
18.
Biochemistry ; 37(32): 11366-75, 1998 Aug 11.
Article in English | MEDLINE | ID: mdl-9698384

ABSTRACT

The EPR spectra of three different molybdoprotein aldehyde dehydrogenases, one purified from Comamonas testosteroni and two purified from Amycolatopsis methanolica, showed in their oxidized state a novel type of signal. These three enzymes contain two different [2Fe-2S] centers, one flavin and one molybdopterin cytosine dinucleotide, as cofactors all of which are expected to be EPR silent in the oxidized state. The new EPR signal is isotropic with g = 2.004 both at X-band and Q-band frequencies, consists of six partially resolved lines, and shows Curie temperature behavior suggesting that the signal is due to an organic radical with S = 1/2. The EPR spectra of Comamonas testosteroni aldehyde dehydrogenase obtained after cultivation in media containing 15NH4Cl and/or after substitution of H2O for D2O show the presence of both nitrogen and proton hyperfine interactions. Simulations of the spectra of the four possible isotope combinations yield a single set of hyperfine coupling constants. The electron spin shows hyperfine interaction with a single I = 1 (0.9 mT) ascribed to a N nucleus, with a single I = 1/2 (1.5 mT) ascribed to one nonexchangeable H nucleus, and with two, exchangeable, identical I = 1/2 spins (0.6 mT) ascribed to two identical exchangeable protons. Taken together, the observations and simulations rule out amino acid residues or flavin as the origin of the radical. The values of the various hyperfine coupling constants are consistent with the properties expected for a molybdenum(VI)-trihydropterin radical in which the N5 atom is engaged in two hydrogen-bonding interactions with the protein. The majority of the electron (spin) density of the radical is located at and around the N5 atom and at the proton bound to the C6 atom of the pterin ring. The EPR spectrum of the molybdopterin radical broadens above 65 K and is no longer detectable above 168 K, indicating that it is not magnetically isolated. The line broadening is ascribed to cross-relaxation with a nearby, rapidly relaxing, oxidized [2Fe-2S] center involving its magnetic S = 1 excited state in this process. The amount of radical was apparently not changed by addition of aldehydes or oxidants, but it disappeared upon reduction by sodium dithionite. Therefore, whether the molybdenum(VI) trihydropterin radical as detected here is a functional intermediate in catalysis remains to be investigated further.


Subject(s)
Aldehyde Dehydrogenase/metabolism , Bacterial Proteins/metabolism , Coenzymes , Metalloproteins/metabolism , Pteridines/metabolism , Actinobacteria/enzymology , Aldehyde Dehydrogenase/chemistry , Bacterial Proteins/chemistry , Catalysis , Electron Spin Resonance Spectroscopy , Electron Transport , Gram-Negative Aerobic Rods and Cocci/enzymology , Iron/metabolism , Metalloproteins/chemistry , Molybdenum Cofactors , Oxidation-Reduction , Pteridines/chemistry , Sulfur/metabolism
19.
Eur J Biochem ; 255(1): 255-61, 1998 Jul 01.
Article in English | MEDLINE | ID: mdl-9692926

ABSTRACT

Steady-state-kinetics investigations were carried out for the oxidation of aldose sugars by soluble quinoprotein glucose dehydrogenase (GDH) from Acinetobacter calcoaceticus using N-methylphenazonium methyl sulfate (PMS) as artificial electron acceptor. As is not uncommon for a dye-linked dehydrogenase, the enzyme showed ping-pong behaviour and double-substrate inhibition. However, under conditions that avoided its masking by sugar-substrate inhibition as much as possible, negative kinetic cooperativity with respect to sugar substrate oxidation by this enzyme was demonstrated. Arguments are presented that exclude trivial factors as a cause for the phenomenon observed. Experimental data could be fitted with an equation accounting for biphasic cooperativity containing two sets of apparent kinetic parameters, V1 and K1, and V1 and K2, representing the enzyme's Michaelis-Menten behaviour at low and high substrate concentrations, respectively. Assuming that subunit interaction causes the cooperativity effect, the sets express the performance of soluble GDH's two subunits in two states of mutual interaction. From fitting the experimental data for several sugars with this equation, it appeared that their V1 values were similar, although their K1 values varied considerably. This showed that the cooperativity effect dramatically changes the performance of soluble GDH, as reflected by the V2 and K2 values for glucose (in phosphate buffer) being about 10-fold and 100-fold higher than the V1 and K1 values, respectively. Substituting the Ca2+ involved in activation of pyrroloquinoline quinone (PQQ) in soluble GDH by Sr2+ affected the cooperativity effect (an increase of the K1 value) but not the two turnover rates of the hybrid enzyme for glucose. The data suggest that the two catalytic cycles of soluble GDH have different rate-limiting steps compared with that of PQQ-containing methanol dehydrogenase.


Subject(s)
Acinetobacter calcoaceticus/enzymology , Carbohydrate Metabolism , Glucose Dehydrogenases/metabolism , Allosteric Regulation , Glucose/metabolism , Kinetics , Lactose/metabolism , Methylphenazonium Methosulfate/metabolism , Models, Chemical , Monosaccharides/metabolism , Solubility
20.
Protein Eng ; 11(3): 185-98, 1998 Mar.
Article in English | MEDLINE | ID: mdl-9613842

ABSTRACT

A molecular model of QH-ADH, the quinohaemoprotein alcohol dehydrogenase from Comamonas testosteroni, has been built by homology modelling. Sequence similarity of N-terminal residues 1-570 with the alpha-subunit of quinoprotein methanol dehydrogenases (MDHs) from Methylophilus methylotrophus W3A1 and Methylobacterium extorquens provided a basis for the design of the PQQ-binding domain of QH-ADH. Minimal sequence similarity with cytochrome c551 from Ectothiorhodospira halophila and cytochrome c5 from Azotobacter vinelandii has been used to model the C-terminal haem c-binding domain, residues 571-677, absent in MDHs. Distance constraints inferred from 19F-NMR relaxation studies of trifluoromethylphenylhydrazine-derivatized PQQ bound to QH-ADH apoenzyme as well as theoretical relations for optimal electron transfer have been employed to position the haem- and PQQ-binding domains relative to each other. The homology model obtained shows overall topological similarity with the crystal structure of cd1-nitrite reductase from Thiosphera pantotropha. The proposed model accounts for the following: (i) the site that is sensitive to in vivo proteolytic attack; (ii) the substrate specificity in comparison with MDHs; (iii) changes of the spectral properties of the haem c upon reconstitution of apo-enzyme with PQQ; (iv) electronic interaction between haem and PQQ; and (v) enantioselectivity in the conversion of a chiral sec alcohol.


Subject(s)
Alcohol Dehydrogenase/chemistry , Gram-Negative Aerobic Rods and Cocci/chemistry , Gram-Negative Aerobic Rods and Cocci/enzymology , Hemeproteins/chemistry , Amino Acid Sequence , Binding Sites , Heme/chemistry , Models, Chemical , Models, Molecular , Molecular Sequence Data , Protein Conformation , Protein Structure, Secondary , Protein Structure, Tertiary , Quinolones , Reproducibility of Results , Sequence Homology, Amino Acid , Software
SELECTION OF CITATIONS
SEARCH DETAIL
...