(Below N is a link to NCBI taxonomic web page and E link to ESTHER at designed phylum.) > cellular organisms: NE > Bacteria: NE > Proteobacteria: NE > Alphaproteobacteria: NE > Rhizobiales: NE > Xanthobacteraceae: NE > Xanthobacter: NE > Xanthobacter autotrophicus: NE
LegendThis sequence has been compared to family alignement (MSA) red => minority aminoacid blue => majority aminoacid color intensity => conservation rate title => sequence position(MSA position)aminoacid rate Catalytic site Catalytic site in the MSA MINAIRTPDQRFSNLDQYPFSPNYLDDLPGYPGLRAHYLDEGNSDAEDVF LCLHGEPTWSYLYRKMIPVFAESGARVIAPDFFGFGKSDKPVDEEDYTFE FHRNFLLALIERLDLRNITLVVQDWGGFLGLTLPMADPSRFKRLIIMNAC LMTDPVTQPAFSAFVTQPADGFTAWKYDLVTPSDLRLDQFMKRWAPTLTE AEASAYAAPFPDTSYQAGVRKFPKMVAQRDQACIDISTEAISFWQNDWNG QTFMAIGMKDKLLGPDVMYPMKALINGCPEPLEIADAGHFVQEFGEQVAR EALKHFAETE
References
20 moreTitle: One-step orientated immobilization of nanobodies and its application for immunoglobulin purification Fu J, Li J, Wang W, Wu H, Zhou P, Li Y, He Q, Tu Z Ref: Journal of Chromatography A, 1603:15, 2019 : PubMed
Affinity chromatography technologies play an important role in the purification of antibodies. To prepare affinity materials, prior isolation and purification of affinity ligands are required before coupling onto solid supports, which is quite expensive and laborious in large-scale applications. In this study, a one-step approach which circumvents the ligand purification procedures was developed to fabricate affinity gel for purifying immunoglobulin G (IgG). A self-labeling tag, haloalkane dehalogenase, was fused to the C-terminal of an anti-Fc variable domain of the heavy chain of the heavy-chain antibody (AFV) which was isolated in previous work. The AFV binds to various sources of IgG and is highly thermal stable. The fusion protein, namely HAFV, was expressed in Escherichia coli as a soluble protein. The binding affinity of HAFV to the Fc region of IgG was characterized and compared with the untagged anti-Fc nanobody. Next, the HAFV was immobilized directly from the crude cell lysate of isopropylthio-beta-D-galactoside (IPTG) induced E. coli. The effects of NaCl concentrations and pH on the capacity of the HAFV resin were investigated. In addition, the one-step coupled HAFV resin was compared with the AFV resin and commercial resins (Protein A and Protein G) by evaluating the static capacity and stability. Though the Protein A (8.34+/-0.37mg/ml) and Protein G (9.19+/-0.28mg/ml) showed higher static capacity, the static capacity of HAFV resin (8.21+/-0.30mg/ml) was better than that of the untagged AFV gel (6.48+/-0.56mg/ml). The recovery results calculated for the reusability and stability show that there is no significant difference between the results obtained for the HAFV gel with those of the untagged AFV gel and commercial Protein A and G. After stored at 37 for 7 days and recycled 10 times, the static capacity of HAFV gel remains above 78%. Our strategy is site-specific, cost-effective, reproducible, and has the potential to dramatically cut down the costs of affinity materials for IgG purification.
Rational enzyme design presents a major challenge that has not been overcome by computational approaches. One of the key challenges is the difficulty in assessing the magnitude of the maximum possible catalytic activity. In an attempt to overcome this challenge, we introduce a strategy that takes an active enzyme (assuming that its activity is close to the maximum possible activity), design mutations that reduce the catalytic activity, and then try to restore that catalysis by mutating other residues. Here we take as a test case the enzyme haloalkane dehalogenase (DhlA), with a 1,2-dichloroethane substrate. We start by demonstrating our ability to reproduce the results of single mutations. Next, we design mutations that reduce the enzyme activity and finally design double mutations that are aimed at restoring the activity. Using the computational predictions as a guide, we conduct an experimental study that confirms our prediction in one case and leads to inconclusive results in another case with 1,2-dichloroethane as substrate. Interestingly, one of our predicted double mutants catalyzes dehalogenation of 1,2-dibromoethane more efficiently than the wild-type enzyme.
        
Title: Bacillus subtilis Spore Surface Display of Haloalkane Dehalogenase DhaA Wang F, Song T, Jiang H, Pei C, Huang Q, Xi H Ref: Curr Microbiol, 76:1161, 2019 : PubMed
The haloalkane dehalogenase DhaA can degrade sulfur mustard (2,2'-dichlorethyl sulfide; also known by its military designation HD) in a rapid and environmentally safe manner. However, DhaA is sensitive to temperature and pH, which limits its applications in natural or harsh environments. Spore surface display technology using resistant spores as a carrier to ensure enzymatic activity can reduce production costs and extend the range of applications of DhaA. To this end, we cloned recombinant Bacillus subtilis spores pHY300PLK-cotg-dhaa-6his/DB104(FH01) for the delivery of DhaA from Rhodococcus rhodochrous NCIMB 13064. A dot blotting showed that the fusion protein CotG-linker-DhaA accounted for 0.41% +/- 0.03% (P < 0.01) of total spore coat proteins. Immunofluorescence analyses confirmed that DhaA was displayed on the spore surface. The hydrolyzing activity of DhaA displayed on spores towards the HD analog 2-chloroethyl ethylsulfide was 1.74 +/- 0.06 U/mL (P < 0.01), with a specific activity was 0.34 +/- 0.04 U/mg (P < 0.01). This is the first demonstration that DhaA displayed on the surface of B. subtilis spores retains enzymatic activity, which suggests that it can be used effectively in real-world applications including bioremediation of contaminated environments.
The application of native enzymes may not be economical owing to the stability factor. A smaller protein molecule may be less susceptible to external stresses. Haloalkane dehalogenases (HLDs) that act on toxic haloalkanes may be incorporated as bioreceptors to detect haloalkane contaminants. Therefore, this study aims to develop mini proteins of HLD as an alternative bioreceptor which was able to withstand extreme conditions. Initially, the mini proteins were designed through computer modeling. Based on the results, five designed mini proteins were deemed to be viable stable mini proteins. They were then validated through experimental study. The smallest mini protein (model 5) was chosen for subsequent analysis as it was expressed in soluble form. No dehalogenase activity was detected, thus the specific binding interaction of between 1,3-dibromopropane with mini protein was investigated using isothermal titration calorimetry. Higher binding affinity between 1,3-dibromopropane and mini protein was obtained than the native. Thermal stability study with circular dichroism had proven that the mini protein possessed two times higher Tm value at 83.73 degrees C than the native at 43.97 degrees C. In conclusion, a stable mini protein was successfully designed and may be used as bioreceptors in the haloalkane sensor that is suitable for industrial application.
        
Title: One-step orientated immobilization of nanobodies and its application for immunoglobulin purification Fu J, Li J, Wang W, Wu H, Zhou P, Li Y, He Q, Tu Z Ref: Journal of Chromatography A, 1603:15, 2019 : PubMed
Affinity chromatography technologies play an important role in the purification of antibodies. To prepare affinity materials, prior isolation and purification of affinity ligands are required before coupling onto solid supports, which is quite expensive and laborious in large-scale applications. In this study, a one-step approach which circumvents the ligand purification procedures was developed to fabricate affinity gel for purifying immunoglobulin G (IgG). A self-labeling tag, haloalkane dehalogenase, was fused to the C-terminal of an anti-Fc variable domain of the heavy chain of the heavy-chain antibody (AFV) which was isolated in previous work. The AFV binds to various sources of IgG and is highly thermal stable. The fusion protein, namely HAFV, was expressed in Escherichia coli as a soluble protein. The binding affinity of HAFV to the Fc region of IgG was characterized and compared with the untagged anti-Fc nanobody. Next, the HAFV was immobilized directly from the crude cell lysate of isopropylthio-beta-D-galactoside (IPTG) induced E. coli. The effects of NaCl concentrations and pH on the capacity of the HAFV resin were investigated. In addition, the one-step coupled HAFV resin was compared with the AFV resin and commercial resins (Protein A and Protein G) by evaluating the static capacity and stability. Though the Protein A (8.34+/-0.37mg/ml) and Protein G (9.19+/-0.28mg/ml) showed higher static capacity, the static capacity of HAFV resin (8.21+/-0.30mg/ml) was better than that of the untagged AFV gel (6.48+/-0.56mg/ml). The recovery results calculated for the reusability and stability show that there is no significant difference between the results obtained for the HAFV gel with those of the untagged AFV gel and commercial Protein A and G. After stored at 37 for 7 days and recycled 10 times, the static capacity of HAFV gel remains above 78%. Our strategy is site-specific, cost-effective, reproducible, and has the potential to dramatically cut down the costs of affinity materials for IgG purification.
Rational enzyme design presents a major challenge that has not been overcome by computational approaches. One of the key challenges is the difficulty in assessing the magnitude of the maximum possible catalytic activity. In an attempt to overcome this challenge, we introduce a strategy that takes an active enzyme (assuming that its activity is close to the maximum possible activity), design mutations that reduce the catalytic activity, and then try to restore that catalysis by mutating other residues. Here we take as a test case the enzyme haloalkane dehalogenase (DhlA), with a 1,2-dichloroethane substrate. We start by demonstrating our ability to reproduce the results of single mutations. Next, we design mutations that reduce the enzyme activity and finally design double mutations that are aimed at restoring the activity. Using the computational predictions as a guide, we conduct an experimental study that confirms our prediction in one case and leads to inconclusive results in another case with 1,2-dichloroethane as substrate. Interestingly, one of our predicted double mutants catalyzes dehalogenation of 1,2-dibromoethane more efficiently than the wild-type enzyme.
        
Title: Bacillus subtilis Spore Surface Display of Haloalkane Dehalogenase DhaA Wang F, Song T, Jiang H, Pei C, Huang Q, Xi H Ref: Curr Microbiol, 76:1161, 2019 : PubMed
The haloalkane dehalogenase DhaA can degrade sulfur mustard (2,2'-dichlorethyl sulfide; also known by its military designation HD) in a rapid and environmentally safe manner. However, DhaA is sensitive to temperature and pH, which limits its applications in natural or harsh environments. Spore surface display technology using resistant spores as a carrier to ensure enzymatic activity can reduce production costs and extend the range of applications of DhaA. To this end, we cloned recombinant Bacillus subtilis spores pHY300PLK-cotg-dhaa-6his/DB104(FH01) for the delivery of DhaA from Rhodococcus rhodochrous NCIMB 13064. A dot blotting showed that the fusion protein CotG-linker-DhaA accounted for 0.41% +/- 0.03% (P < 0.01) of total spore coat proteins. Immunofluorescence analyses confirmed that DhaA was displayed on the spore surface. The hydrolyzing activity of DhaA displayed on spores towards the HD analog 2-chloroethyl ethylsulfide was 1.74 +/- 0.06 U/mL (P < 0.01), with a specific activity was 0.34 +/- 0.04 U/mg (P < 0.01). This is the first demonstration that DhaA displayed on the surface of B. subtilis spores retains enzymatic activity, which suggests that it can be used effectively in real-world applications including bioremediation of contaminated environments.
        
Title: Conformational changes allow processing of bulky substrates by a haloalkane dehalogenase with a small and buried active site Kokkonen P, Bednar D, Dockalova V, Prokop Z, Damborsky J Ref: Journal of Biological Chemistry, 293:11505, 2018 : PubMed
Haloalkane dehalogenases catalyze the hydrolysis of halogen-carbon bonds in organic halogenated compounds and as such are of great utility as biocatalysts. The crystal structures of the haloalkane dehalogenase DhlA from the bacterium from Xanthobacter autotrophicus GJ10, specifically adapted for the conversion of the small 1,2-dichloroethane (DCE) molecule, display the smallest catalytic site (110 A(3)) within this enzyme family. However, during a substrate-specificity screening, we noted that DhlA can catalyze the conversion of far bulkier substrates, such as the 4-(bromomethyl)-6,7-dimethoxy-coumarin (220 A(3)). This large substrate cannot bind to DhlA without conformational alterations. These conformational changes have been previously inferred from kinetic analysis, but their structural basis has not been understood. Using molecular dynamic simulations, we demonstrate here the intrinsic flexibility of part of the cap domain that allows DhlA to accommodate bulky substrates. The simulations displayed two routes for transport of substrates to the active site, one of which requires the conformational change and is likely the route for bulky substrates. These results provide insights into the structure-dynamics function relationships in enzymes with deeply buried active sites. Moreover, understanding the structural basis for the molecular adaptation of DhlA to 1,2-dichloroethane introduced into the biosphere during the industrial revolution provides a valuable lesson in enzyme design by nature.
An enzyme's substrate specificity is one of its most important characteristics. The quantitative comparison of broad-specificity enzymes requires the selection of a homogenous set of substrates for experimental testing, determination of substrate-specificity data and analysis using multivariate statistics. We describe a systematic analysis of the substrate specificities of nine wild-type and four engineered haloalkane dehalogenases. The enzymes were characterized experimentally using a set of 30 substrates selected using statistical experimental design from a set of nearly 200 halogenated compounds. Analysis of the activity data showed that the most universally useful substrates in the assessment of haloalkane dehalogenase activity are 1-bromobutane, 1-iodopropane, 1-iodobutane, 1,2-dibromoethane and 4-bromobutanenitrile. Functional relationships among the enzymes were explored using principal component analysis. Analysis of the untransformed specific activity data revealed that the overall activity of wild-type haloalkane dehalogenases decreases in the following order: LinB~DbjA>DhlA~DhaA~DbeA~DmbA>DatA~DmbC~DrbA. After transforming the data, we were able to classify haloalkane dehalogenases into four SSGs (substrate-specificity groups). These functional groups are clearly distinct from the evolutionary subfamilies, suggesting that phylogenetic analysis cannot be used to predict the substrate specificity of individual haloalkane dehalogenases. Structural and functional comparisons of wild-type and mutant enzymes revealed that the architecture of the active site and the main access tunnel significantly influences the substrate specificity of these enzymes, but is not its only determinant. The identification of other structural determinants of the substrate specificity remains a challenge for further research on haloalkane dehalogenases.
        
Title: The effect of deuteration on protein structure: a high-resolution comparison of hydrogenous and perdeuterated haloalkane dehalogenase Liu X, Hanson BL, Langan P, Viola RE Ref: Acta Crystallographica D Biol Crystallogr, 63:1000, 2007 : PubMed
Haloalkane dehalogenase from Xanthobacter autotrophicus (XaDHL) was overexpressed under different isotopic conditions to produce fully hydrogenous (h-XaDHL) and perdeuterated (d-XaDHL) enzyme forms. Deuterium atoms at labile positions were allowed to back-exchange during purification and hydrogenous solutions were used for crystallization. Optimal crystals of h-XaDHL and d-XaDHL were obtained under different pH conditions (pH 6.0 and 4.6, respectively) but had similar P2(1)2(1)2 unit cells. X-ray diffraction data were refined to 1.53 A (h-XaDHL) and 1.55 A (d-XaDHL) with excellent overall statistics. The conformations of h-XaDHL and d-XaDHL are similar, with slightly altered surface regions because of different packing environments, and h-XaDHL is found to have a more hydrophobic core than d-XaDHL. The active site of h-XaDHL is similar to those of previously determined structures, but the active site of d-XaDHL unexpectedly has some crucial differences. Asp124, the primary nucleophile in the hydrolysis of haloalkane substrates, is displaced from its position in h-XaDHL and rotates to form a hydrogen bond with His289. As a consequence, the water molecule proposed to function as the nucleophile in the next catalytic step is excluded from the active site. This is the first observation of this unusual active-site configuration, which is obtained as a result of perdeuteration that decreases the hydrophobicity of the enzyme, therefore shifting the optimal pH of crystallization. This d-XaDHL structure is likely to represent the termination state of the catalytic reaction and provides an explanation for the acid inhibition of XaDHL. These results underline the importance of carefully verifying the assumption that isotopic substitution does not produce significant structural changes in protein structures.
        
Title: Exploring the binding sites of the haloalkane dehalogenase DhlA from Xanthobacter autotrophicus GJ10 Silberstein M, Damborsky J, Vajda S Ref: Biochemistry, 46:9239, 2007 : PubMed
The catalytic site of haloalkane dehalogenase DhlA is buried more than 10 A from the protein surface. While potential access channels to this site have been reported, the precise mechanism of substrate import and product export is still unconfirmed. We used computational methods to examine surface pockets and their putative roles in ligand access to and from the catalytic site. Computational solvent mapping moves small organic molecule as probes over the protein surface in order to identify energetically favorable sites, that is, regions that tend to bind a variety of molecules. The mapping of three DhlA structures identifies seven such regions, some of which have been previously suggested to be involved in the binding and the import/export of substrates or products. These sites are the active site, the putative entrance of the channel leading to the active site, two pockets that bind Br- ions, a pocket in the slot region, and two additional sites between the main domain and the cap of DhlA. We also performed mapping and free energy analysis of the DhlA structures using the substrate, 1,2-dichloroethane, and halide ions as probes. The findings were compared to crystallographic data and to results obtained by CAVER, a program developed for finding routes from protein clefts and cavities to the surface. Solvent mapping precisely reproduced all three Br- binding sites identified by protein crystallography and the openings to four channels found by CAVER. The analyses suggest that (i) the active site has the highest affinity for the substrate molecule, (ii) the substrate initially binds at the entrance of the main tunnel, (iii) the site Br2, close to the entrance, is likely to serve as an intermediate binding site in product export, (iv) the site Br3, induced in the structure at high concentrations of Br-, could be part of an auxiliary route for product release, and (v) three of the identified sites are likely to be entrances of water-access channels leading to the active site. For comparison, we also mapped haloalkane dehalogenases DhaA and LinB, both of which contain significantly larger and more solvent accessible binding sites than DhlA. The mapping of DhaA and LinB places the majority of probes in the active site, but most of the other six regions consistently identified in DhlA were not observed, suggesting that the more open active site eliminates the need for intermediate binding sites for the collision complex seen in DhlA.
        
Title: Comparative binding energy analysis of the substrate specificity of haloalkane dehalogenase from Xanthobacter autotrophicus GJ10 Kmunicek J, Luengo S, Gago F, Ortiz AR, Wade RC, Damborsky J Ref: Biochemistry, 40:8905, 2001 : PubMed
Comparative binding energy (COMBINE) analysis was conducted for 18 substrates of the haloalkane dehalogenase from Xanthobacter autotrophicus GJ10 (DhlA): 1-chlorobutane, 1-chlorohexane, dichloromethane, 1,2-dichloroethane, 1,2-dichloropropane, 2-chloroethanol, epichlorohydrine, 2-chloroacetonitrile, 2-chloroacetamide, and their brominated analogues. The purpose of the COMBINE analysis was to identify the amino acid residues determining the substrate specificity of the haloalkane dehalogenase. This knowledge is essential for the tailoring of this enzyme for biotechnological applications. Complexes of the enzyme with these substrates were modeled and then refined by molecular mechanics energy minimization. The intermolecular enzyme-substrate energy was decomposed into residue-wise van der Waals and electrostatic contributions and complemented by surface area dependent and electrostatic desolvation terms. Partial least-squares projection to latent structures analysis was then used to establish relationships between the energy contributions and the experimental apparent dissociation constants. A model containing van der Waals and electrostatic intermolecular interaction energy contributions calculated using the AMBER force field explained 91% (73% cross-validated) of the quantitative variance in the apparent dissociation constants. A model based on van der Waals intermolecular contributions from AMBER and electrostatic interactions derived from the Poisson-Boltzmann equation explained 93% (74% cross-validated) of the quantitative variance. COMBINE models predicted correctly the change in apparent dissociation constants upon single-point mutation of DhlA for six enzyme-substrate complexes. The amino acid residues contributing most significantly to the substrate specificity of DhlA were identified; they include Asp124, Trp125, Phe164, Phe172, Trp175, Phe222, Pro223, and Leu263. These residues are suitable targets for modification by site-directed mutagenesis.
Haloalkane dehalogenase (DhlA) converts haloalkanes to their corresponding alcohols and halide ions. The rate-limiting step in the reaction of DhlA is the release of the halide ion. The kinetics of halide release have been analyzed by measuring halide binding with stopped-flow fluorescence experiments. At high halide concentrations, halide import occurs predominantly via the rapid formation of a weak initial collision complex, followed by transport of the ion to the active site. To obtain more insight in this collision complex, we determined the X-ray structure of DhlA in the presence of bromide and investigated the kinetics of mutants that were constructed on the basis of this structure. The X-ray structure revealed one bromide ion firmly bound in the active site and two bromide ions weakly bound on the surface of the enzyme. One of the weakly bound ions is close to Thr197 and Phe294, near the entrance of the earlier proposed tunnel for substrate import. Kinetic analysis of bromide import by the Thr197Ala and Phe294Ala mutants of DhlA at high halide concentration showed that the rate constants for halide binding no longer displayed a wild-type-like parabolic increase with increasing bromide concentrations. This is in agreement with an elimination or a decrease in affinity of the surface-located halide-binding site. Likewise, chloride binding kinetics of the mutants indicated significant differences with wild-type enzyme. The results indicate that Thr197 and Phe294 are involved in the formation of an initial collision complex for halide import in DhlA and provide experimental evidence for the role of the tunnel in substrate and product transport.
        
Title: Haloalkane dehalogenase from Xanthobacter autotrophicus GJ10 refined at 1.15 A resolution Ridder IS, Rozeboom HJ, Dijkstra BW Ref: Acta Crystallographica D Biol Crystallogr, 55:1273, 1999 : PubMed
Crystals of the 35 kDa protein haloalkane dehalogenase from Xanthobacter autotrophicus GJ10 diffract to 1.15 A resolution at cryogenic temperature using synchrotron radiation. Blocked anisotropic least-squares refinement with SHELXL gave a final conventional R factor of 10.51% for all reflections in the 15-1.15 A resolution range. The estimated r.m.s. errors of the model are 0.026 and 0.038 A for protein atoms and all atoms, respectively. The structure comprises all 310 amino acids, with 28 side chains and two peptide bonds in multiple conformations, two covalently linked Pb atoms, 601 water molecules, seven glycerol molecules, one sulfate ion and two chloride ions. Water molecules accounting for alternative solvent structure are modelled with a fixed occupancy of 0.5. The structure is described in detail and compared with previously reported dehalogenase structures refined at 1.9-2.3 A resolution. An analysis of the protein's geometry and stereochemistry reveals eight mean values of bond lengths and angles which deviate significantly from the Engh & Huber parameters, a wide spread in the main-chain omega torsion angle around its ideal value of 180 (6) degrees and a role for C-HcO interactions in satisfying the hydrogen-bond acceptor capacity of main-chain carbonyl O atoms in the central beta-sheet.
Haloalkane dehalogenase (DhlA) catalyzes the hydrolysis of haloalkanes via an alkyl-enzyme intermediate. Trp175 forms a halogen/halide-binding site in the active-site cavity together with Trp125. To get more insight in the role of Trp175 in DhlA, we mutated residue 175 and explored the kinetics and X-ray structure of the Trp175Tyr enzyme. The mutagenesis study indicated that an aromatic residue at position 175 is important for the catalytic performance of DhlA. Pre-steady-state kinetic analysis of Trp175Tyr-DhlA showed that the observed 6-fold increase of the Km for 1,2-dibromoethane (DBE) results from reduced rates of both DBE binding and cleavage of the carbon-bromine bond. Furthermore, the enzyme isomerization preceding bromide release became 4-fold faster in the mutant enzyme. As a result, the rate of hydrolysis of the alkyl-enzyme intermediate became the main determinant of the kcat for DBE, which was 2-fold higher than the wild-type kcat. The X-ray structure of the mutant enzyme at pH 6 showed that the backbone structure of the enzyme remains intact and that the tyrosine side chain lies in the same plane as Trp175 in the wild-type enzyme. The Clalpha-stabilizing aromatic rings of Tyr175 and Trp125 are 0.7 A further apart and due to the smaller size of the mutated residue, the volume of the cavity has increased by one-fifth. X-ray structures of mutant and wild-type enzyme at pH 5 demonstrated that the Tyr175 side chain rotated away upon binding of an acetic acid molecule, leaving one of its oxygen atoms hydrogen bonded to the indole nitrogen of Trp125 only. These structural changes indicate a weakened interaction between residue 175 and the halogen atom or halide ion in the active site and help to explain the kinetic changes induced by the Trp175Tyr mutation.
Conversion of halogenated aliphatics by haloalkane dehalogenase proceeds via the formation of a covalent alkyl-enzyme intermediate which is subsequently hydrolyzed by water. In the wild type enzyme, the slowest step for both 1,2-dichloroethane and 1,2-dibromoethane conversion is a unimolecular enzyme isomerization preceding rapid halide dissociation. Phenylalanine 172 is located in a helix-loop-helix structure that covers the active site cavity of the enzyme, interacts with the C1 beta of 1,2-dichloroethane during catalysis, and could be involved in stabilization of this helix-loop-helix region of the cap domain of the enzyme. To obtain more information about the role of this residue in dehalogenase function, we performed a mutational analysis of position 172 and studied the kinetics and X-ray structure of the Phe172Trp enzyme. The Phe172Trp mutant had a 10-fold higher Kcat/Km for 1-chlorohexane and a 2-fold higher Kcat for 1,2-dibromoethane than the wild-type enzyme. The X-ray structure of the Phe172Trp enzyme showed a local conformational change in the helix-loop-helix region that covers the active site. This could explain the elevated activity for 1-chlorohexane of the Phe172Trp enzyme, since it allows this large substrate to bind more easily in the active site cavity. Pre-steady-state kinetic analysis showed that the increase in Kcat found for 1,2-dibromoethane conversion could be attributed to an increase in the rate of an enzyme isomerization step that preceeds halide release. The observed conformational difference between the helix-loop-helix structures of the wild-type enzyme and the faster mutant suggests that the isomerization required for halide release could be a conformational change that takes place in this region of the cap domain of the dehalogenase. It is proposed that Phe172 is involved in stabilization of the helix-loop-helix structure that covers the active site of the enzyme and creates a rigid hydrophobic cavity for small apolar halogenated alkanes.
        
Title: Sequence analysis of the upstream region of dhlB, the gene encoding haloalkanoic acid dehalogenase of Xanthobacter autotrophicus GJ10 van der Ploeg J, Janssen DB Ref: Biodegradation, 6:257, 1995 : PubMed
The DNA sequence upstream of the dhlB gene encoding the haloalkanoic acid dehalogenase of Xanthobacter autotrophicus GJ10 was determined and contained an open reading frame, designated dhlC, which encoded a protein with a significant similarity with the family of Na(+)-dependent symport proteins. The dhlC gene was subcloned under control of a T7 promoter, and found to encode a polypeptide of 45 kDa on SDS-PAGE. Upstream of dhlC, a -24/-12 promoter sequence was found. Further upstream, in the opposite direction of transcription, another open reading frame, designated dhlR, with homology with the family of sigma 54-dependent transcriptional activator proteins was detected. The dhlR gene was cloned and expressed under the control of a T7 promoter and encoded a polypeptide of 51 kDa on SDS-PAGE. The genetic organization of the dhlB region suggested that the expression of dhlC and dhlB was controlled by the product of dhlR and sigma 54 which may explain the observed overexpression of the haloalkanoic acid dehalogenase under starvation conditions.
Crystal structures of haloalkane dehalogenase were determined in the presence of the substrate 1,2-dichloroethane. At pH 5 and 4 degrees C, substrate is bound in the active site without being converted; warming to room temperature causes the substrate's carbon-chlorine bond to be broken, producing a chloride ion with concomitant alkylation of the active-site residue Asp124. At pH 6 and room temperature the alkylated enzyme is hydrolysed by a water molecule activated by the His289-Asp260 pair in the active site. These results show that catalysis by the dehalogenase proceeds by a two-step mechanism involving an ester intermediate covalently bound at Asp124.
        
Title: Crystallographic and fluorescence studies of the interaction of haloalkane dehalogenase with halide ions. Studies with halide compounds reveal a halide binding site in the active site Verschueren KH, Kingma J, Rozeboom HJ, Kalk KH, Janssen DB, Dijkstra BW Ref: Biochemistry, 32:9031, 1993 : PubMed
Haloalkane dehalogenase from Xanthobacter autotrophicus GJ10 catalyzes the conversion of 1,2-dichloroethane to 2-chloroethanol and chloride without use of oxygen or cofactors. The active site is situated in an internal cavity, which is accessible from the solvent, even in the crystal. Crystal structures of the dehalogenase enzyme complexed with iodoacetamide, chloroacetamide, iodide, and chloride at pH 6.2 and 8.2 revealed a halide binding site between the ring NH's of two tryptophan residues, Trp-125 and Trp-175, located in the active site. The halide ion lies on the intersection of the planes of the rings of the tryptophans. The binding of iodide and chloride to haloalkane dehalogenase caused a strong decrease in protein fluorescence. The decrease could be fitted to a modified form of the Stern-Volmer equation, indicating the presence of fluorophors of different accessibilities. Halide binding was much stronger at pH 6.0 than at pH 8.2. Assuming ligand binding to Trp-125 and Trp-175 as the sole cause of fluorescence quenching, dissociation constants at pH 6.0 with chloride and iodide were calculated to be 0.49 +/- 0.04 and 0.074 +/- 0.007 mM, respectively. Detailed structural investigation showed that the halide binding site probably stabilizes the halide product as well as the negatively charged transition state occurring during the formation of the covalent intermediate.
        
Title: Non-covalent binding of the heavy atom compound [Au(CN)2]- at the halide binding site of haloalkane dehalogenase from Xanthobacter autotrophicus GJ10 Verschueren KH, Franken SM, Rozeboom HJ, Kalk KH, Dijkstra BW Ref: FEBS Letters, 323:267, 1993 : PubMed
The Na[Au(CN)2] heavy atom derivative contributed considerably to the successful elucidation of the crystal structure of haloalkane dehalogenase isolated from Xanthobacter autotrophicus GJ10. The gold cyanide was located in an internal cavity of the enzyme, which also contains the catalytic residues. Refinement of the dehalogenase-gold cyanide complex at 0.25 nm to an R-factor of 16.7% demonstrates that the heavy atom molecule binds non-covalently between two tryptophan residues pointing into the active site cavity. At this same site also chloride ions can be bound. Therefore, inhibition of dehalogenase activity by the Au(CN)-2 presumably occurs by competition for the same binding site as substrates.
        
Title: Crystal structure of haloalkane dehalogenase: an enzyme to detoxify halogenated alkanes Franken SM, Rozeboom HJ, Kalk KH, Dijkstra BW Ref: EMBO Journal, 10:1297, 1991 : PubMed
Haloalkane dehalogenase from Xanthobacter autotrophicus GJ10 converts 1-haloalkanes to the corresponding alcohols and halide ions with water as the sole cosubstrate and without any need for oxygen or cofactors. The three-dimensional structure has been determined by multiple isomorphous replacement techniques using three heavy atom derivatives. The structure has been refined at 2.4 A resolution to an R-factor of 17.9%. The monomeric enzyme is a spherical molecule and is composed to two domains: domain I has an alpha/beta type structure with a central eight-stranded mainly parallel beta-sheet. Domain II lies like a cap on top of domain I and consists of alpha-helices connected by loops. Except for the cap domain the structure resembles that of the dienelactone hydrolase in spite of any significant sequence homology. The putative active site is completely buried in an internal hydrophobic cavity which is located between the two domains. From the analysis of the structure it is suggested that Asp124 is the nucleophilic residue essential for the catalysis. It interacts with His289 which is hydrogen-bonded to Asp260.
        
Title: Involvement of a large plasmid in the degradation of 1,2-dichloroethane by Xanthobacter autotrophicus Tardif G, Greer CW, Labbe D, Lau PC Ref: Applied Environmental Microbiology, 57:1853, 1991 : PubMed
Xanthobacter autotrophicus GJ10 is a bacterium that can degrade short-chain halogenated aliphatic compounds such as 1,2-dichloroethane. A 200-kb plasmid, pXAU1, was isolated from this strain and shown to contain the dhlA gene, which codes for haloalkane dehalogenase, the first enzyme in the degradation pathway of 1,2-dichloroethane by GJ10. Loss of pXAU1 resulted in loss of haloalkane dehalogenase activity, significantly decreased chloroacetaldehyde dehydrogenase activity, and loss of resistance to mercuric chloride but did not affect the activity level of haloalkanoate dehalogenase, the second dehalogenase in the degradation of 1,2-dichloroethane.
        
Title: Characterization of the haloacid dehalogenase from Xanthobacter autotrophicus GJ10 and sequencing of the dhlB gene van der Ploeg J, van Hall G, Janssen DB Ref: Journal of Bacteriology, 173:7925, 1991 : PubMed
The haloacid dehalogenase of the 1,2-dichloroethane-utilizing bacterium Xanthobacter autotrophicus GJ10 was purified from a mutant with an eightfold increase in expression of the enzyme. The mutant was obtained by selecting for enhanced resistance to monobromoacetate. The enzyme was purified through (NH4)2SO4 fractionation, DEAE-cellulose chromatography, and hydroxylapatite chromatography. The molecular mass of the protein was 28 kDa as determined with sodium dodecyl sulfate-polyacrylamide gel electrophoresis and 36 kDa as determined with gel filtration on Superose 12 fast protein liquid chromatography. The enzyme was active with 2-halogenated carboxylic acids and converted only the L-isomer of 2-chloropropionic acid with inversion of configuration to produce D-lactate. The activity of the enzyme was not readily influenced by thiol reagents. The gene encoding the haloacid dehalogenase (dhlB) was cloned and could be allocated to a 6.5-kb EcoRI-BglII fragment. Part of this fragment was sequenced, and the dhlB open reading frame was identified by comparison with the N-terminal amino acid sequence of the protein. The gene was found to encode a protein of 27,433 Da that showed considerable homology (60.5 and 61.0% similarity) with the two other haloacid dehalogenases sequenced to date but not with the haloalkane dehalogenase from X. autotrophicus GJ10.
A gene bank from the chlorinated hydrocarbon-degrading bacterium Xanthobacter autotrophicus GJ10 was prepared in the broad-host-range cosmid vector pLAFR1. By using mutants impaired in dichloroethane utilization and strains lacking dehalogenase activities, several genes involved in 1,2-dichloroethane metabolism were isolated. The haloalkane dehalogenase gene dhlA was subcloned, and it was efficiently expressed from its own constitutive promoter in strains of a Pseudomonas sp., Escherichia coli, and a Xanthobacter sp. at levels up to 30% of the total soluble cellular protein. A 3-kilobase-pair BamHI DNA fragment on which the dhlA gene is localized was sequenced. The haloalkane dehalogenase gene was identified by the known N-terminal amino acid sequence of its product and found to encode a 310-amino-acid protein of molecular weight 35,143. Upstream of the dehalogenase gene, a good ribosome-binding site and two consensus E. coli promoter sequences were present.
        
Title: Crystallization of haloalkane dehalogenase from Xanthobacter autotrophicus GJ10 Rozeboom HJ, Kingma J, Janssen DB, Dijkstra BW Ref: Journal of Molecular Biology, 200:611, 1988 : PubMed
Haloalkane dehalogenases are enzymes that release chloride or bromide from n-halogenated alkanes. X-ray quality crystals of haloalkane dehalogenase from the 1,2-dichloroethane-degrading bacterium Xanthobacter autotrophicus GJ10 have been grown at room temperature from 64% saturated ammonium sulfate solutions (pH 6.2 to 6.4). The crystals diffract in the X-ray beam to at least 2.4 A resolution (1 A = 0.1 nm). Their space group is P2(1)2(1)2, with cell dimensions a = 94.1 A, b = 72.8 A, c = 41.4 A and alpha = beta = gamma = 90 degrees. There is one monomer (molecular weight 36,000) per asymmetric unit.
        
Title: Purification and characterization of hydrolytic haloalkane dehalogenase from Xanthobacter autotrophicus GJ10 Keuning S, Janssen DB, Witholt B Ref: Journal of Bacteriology, 163:635, 1985 : PubMed
A new enzyme, haloalkane dehalogenase, was isolated from the 1,2-dichloroethane-utilizing bacterium Xanthobacter autotrophicus GJ10. The purified enzyme catalyzed the hydrolytic dehalogenation of n-halogenated C1 to C4 alkanes, including chlorinated, brominated, and iodinated compounds. The highest activity was found with 1,2-dichloroethane, 1,3-dichloropropane, and 1,2-dibromoethane. The enzyme followed Michaelis-Menten kinetics, and the Km for 1,2-dichloroethane was 1.1 mM. Maximum activity was found at pH 8.2 and 37 degrees C. Thiol reagents such as p-chloromercuribenzoate and iodoacetamide rapidly inhibited the enzyme. The protein consists of a single polypeptide chain of a molecular weight of 36,000, and its amino acid composition and N-terminal sequence are given.
and N-terminal sequence are given