hosted by
publicationslist.org
    

Jeffrey H LS

www.nedermaasventures.nl

Journal articles

2009
2008
2007
2006
Dick Janssen, Maja Majerić-Elenkov, G Hasnaoui, B Hauer, JHLS (2006)  Enantioselective formation and ring-opening of epoxides catalysed by halohydrin dehalogenases.   Biochem Soc Trans 34: Pt 2. 291-295 Apr  
Abstract: Halohydrin dehalogenases catalyse the conversion of vicinal halohydrins into their corresponding epoxides, while releasing halide ions. They can be found in several bacteria that use halogenated alcohols or compounds that are degraded via halohydrins as a carbon source for growth. Biochemical and structural studies have shown that halohydrin dehalogenases are evolutionarily and mechanistically related to enzymes of the SDR (short-chain dehydrogenase/reductase) superfamily. In the reverse reaction, which is epoxide-ring opening, different nucleophiles can be accepted, including azide, nitrite and cyanide. This remarkable catalytic promiscuity allows the enzymatic production of a broad range of beta-substituted alcohols from epoxides. In these oxirane-ring-opening reactions, the halohydrin dehalogenase from Agrobacterium radiobacter displays high enantioselectivity, making it possible to use the enzyme for the preparation of enantiopure building blocks for fine chemicals.
Notes:
2005
Marco W Fraaije, Jin Wu, Dominic P H M Heuts, Erik W van Hellemond, JHLS, Dick B Janssen (2005)  Discovery of a thermostable Baeyer-Villiger monooxygenase by genome mining.   Appl Microbiol Biotechnol 66: 4. 393-400 Jan  
Abstract: Baeyer-Villiger monooxygenases represent useful biocatalytic tools, as they can catalyze reactions which are difficult to achieve using chemical means. However, only a limited number of these atypical monooxygenases are available in recombinant form. Using a recently described protein sequence motif, a putative Baeyer-Villiger monooxygenase (BVMO) was identified in the genome of the thermophilic actinomycete Thermobifida fusca. Heterologous expression of the respective protein in Escherichia coli and subsequent enzyme characterization showed that it indeed represents a BVMO. The NADPH-dependent and FAD-containing monooxygenase is active with a wide range of aromatic ketones, while aliphatic substrates are also converted. The best substrate discovered so far is phenylacetone (k(cat) = 1.9 s(-1), K(M) = 59 microM). The enzyme exhibits moderate enantioselectivity with alpha-methylphenylacetone (enantiomeric ratio of 7). In addition to Baeyer-Villiger reactions, the enzyme is able to perform sulfur oxidations. Different from all known BVMOs, this newly identified biocatalyst is relatively thermostable, displaying an activity half-life of 1 day at 52 degrees C. This study demonstrates that, using effective annotation tools, genomes can efficiently be exploited as a source of novel BVMOs.
Notes:
2004
2003
Lixia Tang, JHLS, Marco W Fraaije, Dick B Janssen (2003)  Kinetic mechanism and enantioselectivity of halohydrin dehalogenase from Agrobacterium radiobacter.   Biochemistry 42: 18. 5378-5386 May  
Abstract: Halohydrin dehalogenase (HheC) from Agrobacterium radiobacter AD1 catalyzes the reversible intramolecular nucleophilic displacement of a halogen by a hydroxyl group in vicinal haloalcohols, producing the corresponding epoxides. The enzyme displays high enantioselectivity toward some aromatic halohydrins. To understand the kinetic mechanism and enantioselectivity of the enzyme, steady-state and pre-steady-state kinetic analysis was performed with p-nitro-2-bromo-1-phenylethanol (PNSHH) as a model substrate. Steady-state kinetic analyses indicated that the k(cat) of the enzyme with the (R)-enantiomer (22 s(-1)) is 3-fold higher than with the (S)-enantiomer and that the K(m) for the (R)-enantiomer (0.009 mM) is about 45-fold lower than that for the (S)-enantiomer, resulting in a high enantiopreference for the (R)-enantiomer. Product inhibition studies revealed that HheC follows an ordered Uni Bi mechanism for both enantiomers, with halide as the first product to be released. To identify the rate-limiting step in the catalytic cycle, pre-steady-state experiments were performed using stopped-flow and rapid-quench methods. The results revealed the existence of a pre-steady-state burst phase during conversion of (R)-PNSHH, whereas no such burst was observed with the (S)-enantiomer. This indicates that a product release step is rate-limiting for the (R)-enantiomer but not for the (S)-enantiomer. This was further examined by doing single-turnover experiments, which revealed that during conversion of the (R)-enantiomer the rate of bromide release is 21 s(-1). Furthermore, multiple turnover analyses showed that the binding of (R)-PNSHH is a rapid equilibrium step and that the rate of formation of product ternary complex is 380 s(-1). Taken together, these findings enabled the formulation of an ordered Uni Bi kinetic mechanism for the conversion of (R)-PNSHH by HheC in which all of the rate constants are obtained. The high enantiopreference for the (R)-enantiomer can be explained by weak substrate binding of the (S)-enantiomer and a lower rate of reaction at the active site.
Notes:
Lixia Tang, Annet E J van Merode, JHLS, Marco W Fraaije, Dick B Janssen (2003)  Steady-state kinetics and tryptophan fluorescence properties of halohydrin dehalogenase from Agrobacterium radiobacter. Roles of W139 and W249 in the active site and halide-induced conformational change.   Biochemistry 42: 47. 14057-14065 Dec  
Abstract: Halohydrin dehalogenase (HheC) from Agrobacterium radiobacter AD1 is a homotetrameric protein containing four tryptophan residues per subunit. The fluorescence properties of the enzyme are strongly influenced by halide binding. To examine the role of the tryptophans (W139, W192, W238, and W249) in halide binding and catalysis, they were individually mutated to a phenylalanine. All mutations, except for W238F, influenced the enzymatic properties. Mutating W192 to phenylalanine inactivated the enzyme and led to dissociation into dimers and monomers. In the structure of HheC, residue W139 and residue W249 from the opposite subunit are close to the active site of the enzyme. Substitution of W139 mainly affected K(m) values with all tested substrates and reduced the enantiopreference for p-nitro-2-bromo-1-phenylethanol. Replacing W249 increased both k(cat) and K(m) values with all tested substrates except for the (S)-enantiomer of p-nitro-2-bromo-1-phenylethanol, for which k(cat) was 3-fold decreased, resulting in a 6-fold increase of the enantioselectivity. Fluorescence measurements revealed that in the ligand-free state the intrinsic protein fluorescence of mutant W139F is higher than that of the wild-type enzyme, while the fluorescence intensity of mutants W238F and W249F was lower. The fluorescence intensities of the W238F and W249F enzymes were increased when they were unfolded or when bromide was added, whereas the fluorescence of mutant W139F was not increased by unfolding or addition of bromide. These results demonstrate that the fluorescence of residues W238 and W249 is partially quenched in the folded ligand-free state, and that W139 is completely quenched and acts as an energy acceptor for the other tryptophan residues as well. Changes of the maximum fluorescence emission wavelength of the HheC variants and the results of acrylamide quenching experiments confirmed that bromide binding induces a local conformational change around the active site, resulting in residue W139 and the quencher group being separated.
Notes:
2002
2001
JHLS, J E van van Vlieg, L Tang, D B Janssen, R M Kellogg (2001)  Highly enantioselective and regioselective biocatalytic azidolysis of aromatic epoxides.   Org Lett 3: 1. 41-43 Jan  
Abstract: [figure: see text] The halohydrin dehalogenase from Agrobacterium radiobacter AD1 catalyzed the highly enantioselective and beta-regioselective azidolysis of (substituted) styrene oxides. By means of kinetic resolutions the remaining epoxide and the formed azido alcohol could be obtained in very high ee. In a large scale conversion, the decrease in yield and selectivity due to the uncatalyzed chemical side reaction could be overcome by slow addition of azide.
Notes:
Johan van van Vlieg, Lixia Tang, JHLS, T Smilda, G J Poelarends, T Bosma, A E van Merode, M W Fraaije, D B Janssen (2001)  Halohydrin dehalogenases are structurally and mechanistically related to short-chain dehydrogenases/reductases.   J Bacteriol 183: 17. 5058-5066 Sep  
Abstract: Halohydrin dehalogenases, also known as haloalcohol dehalogenases or halohydrin hydrogen-halide lyases, catalyze the nucleophilic displacement of a halogen by a vicinal hydroxyl function in halohydrins to yield epoxides. Three novel bacterial genes encoding halohydrin dehalogenases were cloned and expressed in Escherichia coli, and the enzymes were shown to display remarkable differences in substrate specificity. The halohydrin dehalogenase of Agrobacterium radiobacter strain AD1, designated HheC, was purified to homogeneity. The k(cat) and K(m) values of this 28-kDa protein with 1,3-dichloro-2-propanol were 37 s(-1) and 0.010 mM, respectively. A sequence homology search as well as secondary and tertiary structure predictions indicated that the halohydrin dehalogenases are structurally similar to proteins belonging to the family of short-chain dehydrogenases/reductases (SDRs). Moreover, catalytically important serine and tyrosine residues that are highly conserved in the SDR family are also present in HheC and other halohydrin dehalogenases. The third essential catalytic residue in the SDR family, a lysine, is replaced by an arginine in halohydrin dehalogenases. A site-directed mutagenesis study, with HheC as a model enzyme, supports a mechanism for halohydrin dehalogenases in which the conserved Tyr145 acts as a catalytic base and Ser132 is involved in substrate binding. The primary role of Arg149 may be lowering of the pK(a) of Tyr145, which abstracts a proton from the substrate hydroxyl group to increase its nucleophilicity for displacement of the neighboring halide. The proposed mechanism is fundamentally different from that of the well-studied hydrolytic dehalogenases, since it does not involve a covalent enzyme-substrate intermediate.
Notes:
2000
Johan van van Vlieg, Hans Leemhuis, JHLS, Dick Janssen (2000)  Characterization of the gene cluster involved in isoprene metabolism in Rhodococcus sp. strain AD45.   J. Bacteriol. 182: 7. 1956-1963 Apr  
Abstract: The genes involved in isoprene (2-methyl-1,3-butadiene) utilization in Rhodococcus sp. strain AD45 were cloned and characterized. Sequence analysis of an 8.5-kb DNA fragment showed the presence of 10 genes of which 2 encoded enzymes which were previously found to be involved in isoprene degradation: a glutathione S-transferase with activity towards 1,2-epoxy-2-methyl-3-butene (isoI) and a 1-hydroxy-2-glutathionyl-2-methyl-3-butene dehydrogenase (isoH). Furthermore, a gene encoding a second glutathione S-transferase was identified (isoJ). The isoJ gene was overexpressed in Escherichia coli and was found to have activity with 1-chloro-2,4-dinitrobenzene and 3,4-dichloro-1-nitrobenzene but not with 1, 2-epoxy-2-methyl-3-butene. Downstream of isoJ, six genes (isoABCDEF) were found; these genes encoded a putative alkene monooxygenase that showed high similarity to components of the alkene monooxygenase from Xanthobacter sp. strain Py2 and other multicomponent monooxygenases. The deduced amino acid sequence encoded by an additional gene (isoG) showed significant similarity with that of alpha-methylacyl-coenzyme A racemase. The results are in agreement with a catabolic route for isoprene involving epoxidation by a monooxygenase, conjugation to glutathione, and oxidation of the hydroxyl group to a carboxylate. Metabolism may proceed by fatty acid oxidation after removal of glutathione by a still-unknown mechanism.
Notes:
Rick Rink, Jaap Kingma, JHLS, Dick Janssen (2000)  Tyrosine residues serve as proton donor in the catalytic mechanism of epoxide hydrolase from Agrobacterium radiobacter.   Biochemistry 39: 18. 5600-5613 May  
Abstract: Epoxide hydrolase from Agrobacterium radiobacter catalyzes the hydrolysis of epoxides to their diols via an alkyl-enzyme intermediate. The recently solved X-ray structure of the enzyme shows that two tyrosine residues (Tyr152 and Tyr215) are positioned close to the nucleophile Asp107 in such a way that they can serve as proton donor in the alkylation reaction step. The role of these tyrosines, which are conserved in other epoxide hydrolases, was studied by site-directed mutagenesis. Mutation of Tyr215 to Phe and Ala and mutation of Tyr152 to Phe resulted in mutant enzymes of which the k(cat) values were only 2-4-fold lower than for wild-type enzyme, whereas the K(m) values for the (R)-enantiomers of styrene oxide and p-nitrostyrene oxide were 3 orders of magnitude higher than the K(m) values of wild-type enzyme, showing that the alkylation half-reaction is severely affected by the mutations. Pre-steady-state analysis of the conversion of (R)-styrene oxide by the Y215F and Y215A mutants showed that the 1000-fold elevated K(m) values were mainly caused by a 15-40-fold increase in K(S) and a 20-fold reduction in the rate of alkylation. The rates of hydrolysis of the alkyl-enzyme intermediates were not significantly affected by the mutations. The double mutant Y152F+Y215F showed only a low residual activity for (R)-styrene oxide, with a k(cat)/K(m) value that was 6 orders of magnitude lower than with wild-type enzyme and 3 orders of magnitude lower than with the single tyrosine mutants. This indicates that the effects of the mutations were cumulative. The side chain of Gln134 is positioned in the active site of the X-ray structure of epoxide hydrolase. Mutation of Gln134 to Ala resulted in an active enzyme with slightly altered steady-state kinetic parameters compared to wild-type enzyme, indicating that Gln134 is not essential for catalysis and that the side chain of Gln134 mimics bound substrate. Based upon this observation, the inhibitory potential of various unsubstituted amides was tested, resulting in the identification of phenylacetamide as a competitive inhibitor with an inhibition constant of 30 microM.
Notes:
Antonin Tuynman, JHLS, I M Kooter, H E Schoemaker, R Wever (2000)  Enantioselective epoxidation and carbon-carbon bond cleavage catalyzed by Coprinus cinereus peroxidase and myeloperoxidase.   J. Biol. Chem. 275: 5. 3025-3030 Feb  
Abstract: We demonstrate that myeloperoxidase (MPO) and Coprinus cinereus peroxidase (CiP) catalyze the enantioselective epoxidation of styrene and a number of substituted derivatives with a reasonable enantiomeric excess (up to 80%) and in a moderate yield. Three major differences with respect to the chloroperoxidase from Caldariomyces fumago (CPO) are observed in the reactivity of MPO and CiP toward styrene derivatives. First, in contrast to CPO, MPO and CiP produced the (S)-isomers of the epoxides in enantiomeric excess. Second, for MPO and CiP the H(2)O(2) had to be added very slowly (10 eq in 16 h) to prevent accumulation of catalytically inactive enzyme intermediates. Under these conditions, CPO hardly showed any epoxidizing activity; only with a high influx of H(2)O(2) (300 eq in 1.6 h) was epoxidation observed. Third, both MPO and CiP formed significant amounts of (substituted) benzaldehydes as side products as a consequence of C-alpha-C-beta bond cleavage of the styrene derivatives, whereas for CPO and cytochrome c peroxidase this activity is not observed. C-alpha-C-beta cleavage was the most prominent reaction catalyzed by CiP, whereas with MPO the relative amount of epoxide formed was higher. This is the first report of peroxidases catalyzing both epoxidation reactions and carbon-carbon bond cleavage. The results are discussed in terms of mechanisms involving ferryl oxygen transfer and electron transfer, respectively.
Notes:
1999
1998
1995

Book chapters

2012

PhD theses

2003
JHLS (2003)  Enantioselective Biocatalytic Conversions of Epoxides.   University of Groningen, Prof. D.B. Janssen and Prof R.M. Kellogg  
Abstract: This thesis describes the biocatalyic scope and limitations of an epoxide hydrolase and a haloalcohol dehalogenase. These enzymes, obtained from Agrobucterium radiobacter AD1, were tested for their suitability to prepare optically pure epoxides and derivatives thereof. These compounds can be used for the synthesis of various biologically active agents such as pharmaceuticals. The gram-negative bacterium A. rudiobacter AD1 was isolated from polluted freshwater sediment. Its epoxide hydrolase (EchA) and haloalcohol dehalogenase (HheC) are involved in the degradation of epichlorohydrin and dichloropropanols, enabling the organism to grow on these compounds. The 34 kD epoxide hydrolase was purified and the corresponding gene was cloned by means of PCR. The recombinant EchA can be produced at up to 40% of the total protein content in Escherichia coli, enabling the isolation of 200 to 300 mg of pure protein from a one-liter culture. This makes the enzyme available in sufficient quantities for applications in (synthetic) organic chemistry. In 1997, it was the first microbial epoxide hydrolase gene to be cloned. Until then, all conversions with microbial epoxide hydrolases were perfomred with whole cell suspensions or small quantities of (partially) purified enzyme
Notes:

Patents

2007
2000
1998
1997
Powered by PublicationsList.org.