首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Enzymatic processes are useful for industrially important sugar production, and in vitro two-step isomerization has proven to be an efficient process in utilizing readily available sugar sources. A hypothetical uncharacterized protein encoded by ydaE of Bacillus licheniformis was found to have broad substrate specificities and has shown high catalytic efficiency on d-lyxose, suggesting that the enzyme is d-lyxose isomerase. Escherichia coli BL21 expressing the recombinant protein, of 19.5 kDa, showed higher activity at 40 to 45°C and pH 7.5 to 8.0 in the presence of 1.0 mM Mn2+. The apparent Km values for d-lyxose and d-mannose were 30.4 ± 0.7 mM and 26 ± 0.8 mM, respectively. The catalytic efficiency (kcat/Km) for lyxose (3.2 ± 0.1 mM−1 s−1) was higher than that for d-mannose (1.6 mM−1 s−1). The purified protein was applied to the bioproduction of d-lyxose and d-glucose from d-xylose and d-mannose, respectively, along with the thermostable xylose isomerase of Thermus thermophilus HB08. From an initial concentration of 10 mM d-lyxose and d-mannose, 3.7 mM and 3.8 mM d-lyxose and d-glucose, respectively, were produced by two-step isomerization. This two-step isomerization is an easy method for in vitro catalysis and can be applied to industrial production.  相似文献   

2.
α-Keto-γ-methylthiobutyric acid (KMTB), a keto derivative of l-methionine, has great potential for use as an alternative to l-methionine in the poultry industry and as an anti-cancer drug. This study developed an environment friendly process for KMTB production from l-methionine by an Escherichia coli whole-cell biocatalyst expressing an engineered l-amino acid deaminase (l-AAD) from Proteus vulgaris. We first overexpressed the P. vulgaris l-AAD in E. coli BL21 (DE3) and further optimized the whole-cell transformation process. The maximal molar conversion ratio of l-methionine to KMTB was 71.2% (mol/mol) under the optimal conditions (70 g/L l-methionine, 20 g/L whole-cell biocatalyst, 5 mM CaCl2, 40°C, 50 mM Tris-HCl [pH 8.0]). Then, error-prone polymerase chain reaction was used to construct P. vulgaris l-AAD mutant libraries. Among approximately 104 mutants, two mutants bearing lysine 104 to arginine and alanine 337 to serine substitutions showed 82.2% and 80.8% molar conversion ratios, respectively. Furthermore, the combination of these mutations enhanced the catalytic activity and molar conversion ratio by 1.3-fold and up to 91.4% with a KMTB concentration of 63.6 g/L. Finally, the effect of immobilization on whole-cell transformation was examined, and the immobilized whole-cell biocatalyst with Ca2+ alginate increased reusability by 41.3% compared to that of free cell production. Compared with the traditional multi-step chemical synthesis, our one-step biocatalytic production of KMTB has an advantage in terms of environmental pollution and thus has great potential for industrial KMTB production.  相似文献   

3.
4.
β-Galactosidase-catalysed hydrolysis of β-d-galactopyranosyl azide   总被引:3,自引:3,他引:0  
1. β-d-Galactopyranosyl azide is hydrolysed by the β-galactosidase of Escherichia coli to galactose and azide ion at a mechanistically significant rate. 2. Methyl 1-thio-β-d-galactopyranoside is a competitive inhibitor of the hydrolysis of the azide and of o-nitrophenyl β-d-galactopyranoside with Ki 1.8mm. 3. β-Galactosidase can thus hydrolyse a range of substrates of general structure β-d-galactopyranosyl-X(Y), where the atom X has a lone pair of electrons on which the enzyme may act as a Lewis or Brønsted acid, but in which the length of the bond cleaved varies significantly, which is inconsistent with the orbital steering hypothesis.  相似文献   

5.
6.
Escherichia coli that is unable to metabolize d-glucose (with knockouts in ptsG, manZ, and glk) accumulates a small amount of d-glucose (yield of about 0.01 g/g) during growth on the pentoses d-xylose or l-arabinose as a sole carbon source. Additional knockouts in the zwf and pfkA genes, encoding, respectively, d-glucose-6-phosphate 1-dehydrogenase and 6-phosphofructokinase I (E. coli MEC143), increased accumulation to greater than 1 g/liter d-glucose and 100 mg/liter d-mannose from 5 g/liter d-xylose or l-arabinose. Knockouts of other genes associated with interconversions of d-glucose-phosphates demonstrate that d-glucose is formed primarily by the dephosphorylation of d-glucose-6-phosphate. Under controlled batch conditions with 20 g/liter d-xylose, MEC143 generated 4.4 g/liter d-glucose and 0.6 g/liter d-mannose. The results establish a direct link between pentoses and hexoses and provide a novel strategy to increase carbon backbone length from five to six carbons by directing flux through the pentose phosphate pathway.  相似文献   

7.
Nonreversible d-Glyceraldehyde 3-Phosphate Dehydrogenase of Plant Tissues   总被引:6,自引:6,他引:0  
Kelly GJ  Gibbs M 《Plant physiology》1973,52(2):111-118
Preparations of TPN-linked nonreversible d-glyceraldehyde 3-phosphate dehydrogenase (EC 1.2.1.9), free of TPN-linked reversible d-glyceraldehyde 3-phosphate dehydrogenase, have been obtained from green shoots, etiolated shoots, and cotyledons of pea (Pisum sativum), cotyledons of peanut (Arachis hypogea), and leaves of maize (Zea mays). The properties of the enzyme were similar from each of these sources: the Km values for d-glyceraldehyde 3-phosphate and TPN were about 20 μm and 3 μm, respectively. The enzyme activity was inhibited by l-glyceraldehyde 3-phosphate, d-erythrose 4-phosphate, and phosphohydroxypyruvate. Activity was found predominantly in photosynthetic and gluconeogenic tissues of higher plants. A light-induced, phytochrome-mediated increase of enzyme activity in a photosynthetic tissue (pea shoots) was demonstrated. Appearance of enzyme activity in a gluconeogenic tissue (endosperm of castor bean, Ricinus communis) coincided with the conversion of fat to carbohydrate during germination. In photosynthetic tissue, the enzyme is located outside the chloroplast, and at in vivo levels of triose-phosphates and pyridine nucleotides, the activity is probably greater than that of DPN-linked reversible d-glyceraldehyde 3-phosphate dehydrogenase. Several possible roles for the enzyme in plant carbohydrate metabolism are considered.  相似文献   

8.
Production of N-acetyl-d-neuraminic acid (Neu5Ac) via biocatalysis is traditionally conducted using isolated enzymes or whole cells. The use of isolated enzymes is restricted by the time-consuming purification process, whereas the application of whole cells is limited by the permeability barrier presented by the microbial cell membrane. In this study, a novel type of biocatalyst, Neu5Ac aldolase presented on the surface of Bacillus subtilis spores, was used for the production of Neu5Ac. Under optimal conditions, Neu5Ac at a high concentration (54.7 g liter−1) and a high yield (90.2%) was obtained under a 5-fold excess of pyruvate over N-acetyl-d-mannosamine. The novel biocatalyst system, which is able to express and immobilize the target enzyme simultaneously on the surface of B. subtilis spores, represents a suitable alternative for value-added chemical production.  相似文献   

9.
d-Alanyl:d-lactate (d-Ala:d-Lac) and d-alanyl:d-serine ligases are key enzymes in vancomycin resistance of Gram-positive cocci. They catalyze a critical step in the synthesis of modified peptidoglycan precursors that are low binding affinity targets for vancomycin. The structure of the d-Ala:d-Lac ligase VanA led to the understanding of the molecular basis for its specificity, but that of d-Ala:d-Ser ligases had not been determined. We have investigated the enzymatic kinetics of the d-Ala:d-Ser ligase VanG from Enterococcus faecalis and solved its crystal structure in complex with ADP. The overall structure of VanG is similar to that of VanA but has significant differences mainly in the N-terminal and central domains. Based on reported mutagenesis data and comparison of the VanG and VanA structures, we show that residues Asp-243, Phe-252, and Arg-324 are molecular determinants for d-Ser selectivity. These residues are conserved in both enzymes and explain why VanA also displays d-Ala:d-Ser ligase activity, albeit with low catalytic efficiency in comparison with VanG. These observations suggest that d-Ala:d-Lac and d-Ala:d-Ser enzymes have evolved from a common ancestral d-Ala:d-X ligase. The crystal structure of VanG showed an unusual interaction between two dimers involving residues of the omega loop that are deeply anchored in the active site. We constructed an octapeptide mimicking the omega loop and found that it selectively inhibits VanG and VanA but not Staphylococcus aureus d-Ala:d-Ala ligase. This study provides additional insight into the molecular evolution of d-Ala:d-X ligases and could contribute to the development of new structure-based inhibitors of vancomycin resistance enzymes.  相似文献   

10.
Previously, we successfully cloned a d-cycloserine (d-CS) biosynthetic gene cluster consisting of 10 open reading frames (designated dcsA to dcsJ) from d-CS-producing Streptomyces lavendulae ATCC 11924. In this study, we put four d-CS biosynthetic genes (dcsC, dcsD, dcsE, and dcsG) in tandem under the control of the T7 promoter in an Escherichia coli host. SDS-PAGE analysis demonstrated that the 4 gene products were simultaneously expressed in host cells. When l-serine and hydroxyurea (HU), the precursors of d-CS, were incubated together with the E. coli resting cell suspension, the cells produced significant amounts of d-CS (350 ± 20 μM). To increase the productivity of d-CS, the dcsJ gene, which might be responsible for the d-CS excretion, was connected downstream of the four genes. The E. coli resting cells harboring the five genes produced d-CS at 660 ± 31 μM. The dcsD gene product, DcsD, forms O-ureido-l-serine from O-acetyl-l-serine (OAS) and HU, which are intermediates in d-CS biosynthesis. DcsD also catalyzes the formation of l-cysteine from OAS and H2S. To repress the side catalytic activity of DcsD, the E. coli chromosomal cysJ and cysK genes, encoding the sulfite reductase α subunit and OAS sulfhydrylase, respectively, were disrupted. When resting cells of the double-knockout mutant harboring the four d-CS biosynthetic genes, together with dcsJ, were incubated with l-serine and HU, the d-CS production was 980 ± 57 μM, which is comparable to that of d-CS-producing S. lavendulae ATCC 11924 (930 ± 36 μM).  相似文献   

11.
The oxidative d-xylose catabolic pathway of Caulobacter crescentus, encoded by the xylXABCD operon, was expressed in the gram-negative bacterium Pseudomonas putida S12. This engineered transformant strain was able to grow on d-xylose as a sole carbon source with a biomass yield of 53% (based on g [dry weight] g d-xylose−1) and a maximum growth rate of 0.21 h−1. Remarkably, most of the genes of the xylXABCD operon appeared to be dispensable for growth on d-xylose. Only the xylD gene, encoding d-xylonate dehydratase, proved to be essential for establishing an oxidative d-xylose catabolic pathway in P. putida S12. The growth performance on d-xylose was, however, greatly improved by coexpression of xylXA, encoding 2-keto-3-deoxy-d-xylonate dehydratase and α-ketoglutaric semialdehyde dehydrogenase, respectively. The endogenous periplasmic glucose dehydrogenase (Gcd) of P. putida S12 was found to play a key role in efficient oxidative d-xylose utilization. Gcd activity not only contributes to d-xylose oxidation but also prevents the intracellular accumulation of toxic catabolic intermediates which delays or even eliminates growth on d-xylose.The requirement for renewable alternatives to replace oil-based chemicals and fuels necessitates development of novel technologies. Lignocellulose provides a promising alternative feedstock. However, since the pentose sugar fraction may account for up to 25% of lignocellulosic biomass (12), it is essential that this fraction is utilized efficiently to obtain cost-effective biochemical production. In a previous study, the solvent-tolerant bacterium Pseudomonas putida S12, known for its use as a platform host for the production of aromatic compounds (15, 16, 19, 22), was engineered to use d-xylose as a sole carbon source. This was achieved by introducing genes encoding the phosphorylative d-xylose metabolic pathway of Escherichia coli, followed by laboratory evolution (14). Prior to evolutionary improvement, extensive oxidation of d-xylose to d-xylonate occurred, resulting in a very low biomass-for-substrate yield as d-xylonate is a metabolic dead-end product in P. putida. The evolution approach resulted in elimination of the activity of periplasmic glucose dehydrogenase (Gcd), the enzyme responsible for d-xylose oxidation, which turned out to be a critical step in optimizing phosphorylative d-xylose utilization in P. putida S12.Instead of prevention of endogenous oxidation of d-xylose, this oxidation may be used to our advantage when it is combined with an oxidative d-xylose metabolic pathway, such as the pathways described for several Pseudomonas species, Caulobacter crescentus, and Haloarcula marismortui (7, 11, 18, 20). In these pathways, d-xylonate is dehydrated to 2-keto-3-deoxy-d-xylonate. This intermediate either can be cleaved into pyruvate and glycolaldehyde (7) or is further dehydrated to α-ketoglutaric semialdehyde (α-KGSA). In the final step of the latter pathway, α-KGSA is oxidized to the tricarboxylic acid (TCA) cycle intermediate α-ketoglutarate (18, 20).In addition to Gcd (PP1444), some of the enzymes required for oxidative d-xylose metabolism are expected to be endogenous in P. putida S12. Transport of d-xylonate into the cytoplasm likely occurs through the gluconate transporter (encoded by gntP [PP3417]). The enzyme catalyzing the final step of the pathway, α-KGSA dehydrogenase, is also likely to be present (presumably PP1256 and/or PP3602) because of the requirement for metabolism of 4-hydroxyproline (1), a compound that is efficiently utilized by P. putida S12. In view of these properties, the most obvious approach for constructing d-xylose-utilizing P. putida S12 is reconstruction of a complete oxidative d-xylose metabolic pathway by introducing the parts of such a pathway that complement the endogenous activities. Recently, the genetic information for one such oxidative d-xylose pathway has become available (18), enabling the approach used in the present study, i.e., expression of the oxidative d-xylose metabolic pathway of C. crescentus in P. putida S12 and investigation of the contribution of endogenous enzyme activities.  相似文献   

12.
The lipopolysaccharide of Pseudomonas aeruginosa PAO1 contains an unusual sugar, 2,3-diacetamido-2,3-dideoxy-d-mannuronic acid (d-ManNAc3NAcA). wbpB, wbpE, and wbpD are thought to encode oxidase, transaminase, and N-acetyltransferase enzymes. To characterize their functions, recombinant proteins were overexpressed and purified from heterologous hosts. Activities of His6-WbpB and His6-WbpE were detected only when both proteins were combined in the same reaction. Using a direct MALDI-TOF mass spectrometry approach, we identified ions that corresponded to the predicted products of WbpB (UDP-3-keto-d-GlcNAcA) and WbpE (UDP-d-GlcNAc3NA) in the coupled enzyme-substrate reaction. Additionally, in reactions involving WbpB, WbpE, and WbpD, an ion consistent with the expected product of WbpD (UDP-d-GlcNAc3NAcA) was identified. Preparative quantities of UDP-d-GlcNAc3NA and UDP-d-GlcNAc3NAcA were enzymatically synthesized. These compounds were purified by high-performance liquid chromatography, and their structures were elucidated by NMR spectroscopy. This is the first report of the functional characterization of these proteins, and the enzymatic synthesis of UDP-d-GlcNAc3NA and UDP-d-GlcNAc3NAcA.Gram-negative organisms such as Pseudomonas aeruginosa produce lipopolysaccharide (LPS)4 as an essential component of the outer leaflet of the outer membrane. LPS can be conceptually divided into three parts: lipid A, which anchors LPS into the membrane; core oligosaccharide, which contributes to membrane stability; and the O-antigen, which is a polysaccharide that extends away from the cell surface. In P. aeruginosa, two types of O-antigen are observed: A-band O-antigen, which is common to most strains, and B-band O-antigen, which is variable and therefore used as the basis of the International Antigenic Typing Scheme (1). P. aeruginosa serotypes O2, O5, O16, O18, and O20 collectively belong to serogroup O2, because they all share common backbone sugar structures in their O-antigen repeat units consisting of two di-N-acetylated uronic acids and one 2-acetamido-2,6-dideoxy-d-galactose (N-acetyl-d-fucosamine). The minor structural variations in the O-antigen repeat units that differentiate this serogroup into five serotypes are: the type of glycosidic linkage between O-units (alpha versus beta) that is formed by the O-antigen polymerase (Wzy), isomers present (d-mannuronic or l-guluronic acid), and acetyl group substituents (24). The B-band O-antigen of P. aeruginosa PAO1 (serotype O5) contains a repeating trisaccharide of 2-acetamido-3-acetamidino-2,3-dideoxy-d-mannuronic acid (d-ManNAc3NAmA), 2,3-diacetamido-2,3-dideoxy-d-mannuronic acid (d-ManNAc3NAcA), and 2-acetamido-2,6-dideoxy-d-galactose (3).The biosynthesis of the two mannuronic acid derivatives has yet to be fully understood and has been the subject of investigation by our group. To produce UDP-d-ManNAc3NAcA, a five-step pathway has been proposed (Fig. 1) that requires the products of five genes localized to the B-band O-antigen biosynthesis cluster (5). The O-antigen biosynthesis cluster was shown to be identical for all serotypes within serogroup O2, which further underscores the high similarity between these serotypes (5). The five genes, including wbpA, wbpB, wbpE, wbpD, and wbpI, have been shown to be essential for B-band LPS biosynthesis, because knockout mutants of each of these genes are deficient in B-band O-antigen (68). Homologs of all five of the proteins required for the UDP-d-ManNAc3NAcA biosynthesis pathway are conserved in other bacterial pathogens, including Bordetella pertussis, Bordetella parapertussis, and Bordetella bronchiseptica. Cross-complementation of P. aeruginosa knockout mutants lacking wbpA, wbpB, wbpE, wbpD, or wbpI with the homologues from B. pertussis could fully restore LPS production in the P. aeruginosa LPS mutants, suggesting that the genes from B. pertussis are functional homologs of the wbp genes (7). Homologs of these genes could be identified in diverse bacterial species, demonstrating the importance of UDP-d-ManNAc3NAcA biosynthesis beyond its role in P. aeruginosa (7).Open in a separate windowFIGURE 1.Proposed pathway for the biosynthesis of UDP-d-ManNAc3NAcA in P. aeruginosa PAO1. The full names of the sugars are as follows: GlcNAc, 2-acetamido-2-deoxy-d-glucose; GlcNAcA, 2-acetamido-2-deoxy-d-glucuronic acid; 3-keto-d-GlcNAcA, 2-acetamido-2-deoxy-d-ribo-hex-3-uluronic acid; GlcNAc3NA, 2-acetamido-3-amino-2,3-dideoxy-d-glucuronic acid; GlcNAc3NAcA, 2,3-diacetamido-2,3-dideoxy-d-glucuronic acid; ManNAc3NAcA, 2,3-diacetamido-2,3-dideoxy-d-mannuronic acid. Adapted from Ref. 8.The first enzyme of the UDP-d-ManNAc3NAcA biosynthesis pathway, WbpA, is a 6-dehydrogenase that converts UDP-2-acetamido-2-deoxy-d-glucose (N-acetyl-d-glucosamine; UDP-d-GlcNAc) to UDP-2-acetamido-2-deoxy-d-glucuronic acid (N-acetyl-d-glucosaminuronic acid, UDP-d-GlcNAcA) using NAD+ as a coenzyme (9) (Fig. 1). Following this, the second step in UDP-d-ManNAc3NAcA biosynthesis is proposed to be an oxidation reaction catalyzed by WbpB, forming UDP-2-acetamido-2-deoxy-d-ribo-hex-3-uluronic acid (3-keto-d-GlcNAcA), which in turn is used as the substrate for transamination by WbpE, creating UDP-2-acetamido-3-amino-2,3-dideoxy-d-glucuronic acid (d-GlcNAc3NA).This residue is thought to be the substrate for WbpD, a putative N-acetyltransferase of the hexapeptide acyltransferase superfamily (10) that requires acetyl-CoA as a co-substrate (8). WbpD has been proposed to synthesize UDP-2,3-diacetamido-2,3-dideoxy-d-glucuronic acid (UDP-d-GlcNAc-3NAcA), which is utilized in the B-band O-antigen of P. aeruginosa serotype O1. In P. aeruginosa serogroup O2, the UDP-d-GlcNAc3NAcA is then epimerized by WbpI to create the UDP-d-ManNAc3NAcA required for incorporation into B-band LPS (11). A derivative of UDP-d-ManNAc3NAcA is also used in the synthesis of B-band O-antigen of P. aeruginosa serogroup O2. UDP-d-ManNAc3NAmA is thought to be produced through additional modification of UDP-d-ManNAc3NAcA via the action of WbpG, an amidotransferase, which has also been demonstrated to be essential for the production of B-band O-antigen (12, 13).In the current study, our aim was to define the function of WbpB, WbpE, and WbpD, because only genetic evidence has previously been given for the involvement of wbpB and wbpE (7), and the reaction catalyzed by WbpD could not be demonstrated due to the unavailability of its presumed substrate, UDP-d-GlcNAc3NA (8). The functional characterization of these proteins is also important for understanding LPS biosynthesis in B. pertussis, because the genes in the LPS locus of this species, wlbA, wlbC, and wlbB, could cross-complement knockouts of wbpB, wbpE, and wbpD, respectively, when expressed in P. aeruginosa PAO1 (7). Furthermore, these three proteins form a cassette for the generation of C-3 N-acetylated hexoses and may be important for the biosynthesis of a variety of other sugars. Capillary electrophoresis and MALDI-TOF mass spectrometry were used to analyze reaction mixtures of WbpB and WbpE and showed that the expected products were produced only when both enzymes were present together. Achieving the enzymatic synthesis of the product of both enzymes, which was demonstrated to be UDP-d-GlcNAc3NA by 1H NMR spectroscopy, was a key breakthrough, because this rare sugar has never before been produced by any means. UDP-d-GlcNAc3NA was also essential for use as the substrate of WbpD, which not only allowed us to determine the enzymatic activity of this protein but also allowed the enzymatic synthesis of UDP-d-GlcNAc3NAcA to be achieved as well. Although this sugar had previously been produced through a 17-step chemical synthesis (11, 14), the 4-step concurrent enzymatic reaction demonstrates the advantage of linking chemistry with biology and represents a significant saving of both time and reagents as compared with chemical synthesis. Finally, our data also showed the success in reconstituting in vitro the 5-step pathway for the biosynthesis of UDP-d-ManNAc3NAcA in P. aeruginosa.  相似文献   

13.
l-Glutamine d-fructose 6-phosphate amidotransferase (EC 2.6.1.16) was extracted and purified 600-fold by acetone fractionation and diethylaminoethyl cellulose column chromatography from mung bean seeds (Phaseolus aureus). The partially purified enzyme was highly specific for l-glutamine as an amide nitrogen donor, and l-asparagine could not replace it. The enzyme showed a pH optimum in the range of 6.2 to 6.7 in phosphate buffer. Km values of 3.8 mm and 0.5 mm were obtained for d-fructose 6-phosphate and l-glutamine, respectively. The enzyme was competitively inhibited with respect to d-fructose 6-phosphate by uridine diphosphate-N-acetyl-d-glucosamine which had a Ki value of 13 μm. Upon removal of l-glutamine and its replacement by d-fructose 6-phosphate and storage over liquid nitrogen, the enzyme was completely desensitized to inhibition by uridine diphosphate-N-acetyl-d-glucosamine. This indicates that the inhibitor site is distinct from the catalytic site and that uridine diphosphate-N-acetyl-d-glucosamine acts as a feedback inhibitor of the enzyme.  相似文献   

14.
15.
Escherichia coli K-12 provided with glucose and a mixture of amino acids depletes l-serine more quickly than any other amino acid even in the presence of ammonium sulfate. A mutant without three 4Fe4S l-serine deaminases (SdaA, SdaB, and TdcG) of E. coli K-12 is unable to do this. The high level of l-serine that accumulates when such a mutant is exposed to amino acid mixtures starves the cells for C1 units and interferes with cell wall synthesis. We suggest that at high concentrations, l-serine decreases synthesis of UDP-N-acetylmuramate-l-alanine by the murC-encoded ligase, weakening the cell wall and producing misshapen cells and lysis. The inhibition by high l-serine is overcome in several ways: by a large concentration of l-alanine, by overproducing MurC together with a low concentration of l-alanine, and by overproducing FtsW, thus promoting septal assembly and also by overexpression of the glycine cleavage operon. S-Adenosylmethionine reduces lysis and allows an extensive increase in biomass without improving cell division. This suggests that E. coli has a metabolic trigger for cell division. Without that reaction, if no other inhibition occurs, other metabolic functions can continue and cells can elongate and replicate their DNA, reaching at least 180 times their usual length, but cannot divide.The Escherichia coli genome contains three genes, sdaA, sdaB, and tdcG, specifying three very similar 4Fe4S l-serine deaminases. These enzymes are very specific for l-serine for which they have unusually high Km values (3, 32). Expression of the three genes is regulated so that at least one of the gene products is synthesized under all common growth conditions (25). This suggests an important physiological role for the enzymes. However, why E. coli needs to deaminate l-serine has been a long-standing problem of E. coli physiology, the more so since it cannot use l-serine as the sole carbon source.We showed recently that an E. coli strain devoid of all three l-serine deaminases (l-SDs) loses control over its size, shape, and cell division when faced with complex amino acid mixtures containing l-serine (32). We attributed this to starvation for single-carbon (C1) units and/or S-adenosylmethionine (SAM). C1 units are usually made from serine via serine hydroxymethyl transferase (GlyA) or via glycine cleavage (GCV). The l-SD-deficient triple mutant strain is starved for C1 in the presence of amino acids, because externally provided glycine inhibits GlyA and a very high internal l-serine concentration along with several other amino acids inhibits glycine cleavage. While the parent cell can defend itself by reducing the l-serine level by deamination, this crucial reaction is missing in the ΔsdaA ΔsdaB ΔtdcG triple mutant. We therefore consider these to be “defensive” serine deaminases.The fact that an inability to deaminate l-serine leads to a high concentration of l-serine and inhibition of GlyA is not surprising. However, it is not obvious why a high level of l-serine inhibits cell division and causes swelling, lysis, and filamentation. Serine toxicity due to inhibition of biosynthesis of isoleucine (11) and aromatic amino acids (21) has been reported but is not relevant here, since these amino acids are provided in Casamino Acids.We show here that at high internal concentrations, l-serine also causes problems with peptidoglycan synthesis, thus weakening the cell wall. Peptidoglycan is a polymer of long glycan chains made up of alternating N-acetylglucosamine and N-acetylmuramic acid residues, cross-linked by l-alanyl-γ-d-glutamyl-meso-diaminopimelyl-d-alanine tetrapeptides (1, 28). The glucosamine and muramate residues and the pentapeptide (from which the tetrapeptide is derived) are all synthesized in the cytoplasm and then are exported to be polymerized into extracellular peptidoglycan (2).In this paper, we show that lysis is caused by l-serine interfering with the first step of synthesis of the cross-linking peptide, the addition of l-alanine to uridine diphosphate-N-acetylmuramate. This interference is probably due to a competition between serine and l-alanine for the ligase, MurC, which adds the first l-alanine to UDP-N-acetylmuramate (7, 10, 15). As described here, the weakening of the cell wall by l-serine can be overcome by a variety of methods that reduce the endogenous l-serine pool or counteract the effects of high levels of l-serine.  相似文献   

16.
Limited uptake is one of the bottlenecks for l-arabinose fermentation from lignocellulosic hydrolysates in engineered Saccharomyces cerevisiae. This study characterized two novel l-arabinose transporters, LAT-1 from Neurospora crassa and MtLAT-1 from Myceliophthora thermophila. Although the two proteins share high identity (about 83%), they display different substrate specificities. Sugar transport assays using the S. cerevisiae strain EBY.VW4000 indicated that LAT-1 accepts a broad substrate spectrum. In contrast, MtLAT-1 appeared much more specific for l-arabinose. Determination of the kinetic properties of both transporters revealed that the Km values of LAT-1 and MtLAT-1 for l-arabinose were 58.12 ± 4.06 mM and 29.39 ± 3.60 mM, respectively, with corresponding Vmax values of 116.7 ± 3.0 mmol/h/g dry cell weight (DCW) and 10.29 ± 0.35 mmol/h/g DCW, respectively. In addition, both transporters were found to use a proton-coupled symport mechanism and showed only partial inhibition by d-glucose during l-arabinose uptake. Moreover, LAT-1 and MtLAT-1 were expressed in the S. cerevisiae strain BSW2AP containing an l-arabinose metabolic pathway. Both recombinant strains exhibited much faster l-arabinose utilization, greater biomass accumulation, and higher ethanol production than the control strain. In conclusion, because of higher maximum velocities and reduced inhibition by d-glucose, the genes for the two characterized transporters are promising targets for improved l-arabinose utilization and fermentation in S. cerevisiae.  相似文献   

17.
We isolated thermotolerant Gluconobacter strains that are able to produce 5-keto-d-gluconic acid (5KGA) at 37°C, a temperature at which regular mesophilic 5KGA-producing strains showed much less growth and 5KGA production. The thermotolerant strains produced 2KGA as the major product at both 30 and 37°C. The amount of ketogluconates produced at 37°C was slightly less than the amount produced at 30°C. To improve the yield of 5KGA in these strains, we disrupted flavin adenine dinucleotide-gluconate dehydrogenase (FAD-GADH), which is responsible for 2KGA production. Genes for FAD-GADH were cloned by using inverse PCR and an in vitro cloning strategy. The sequences obtained for three thermotolerant strains were identical and showed high levels of identity to the FAD-GADH sequence reported for the genome of Gluconobacter oxydans 621 H. A kanamycin resistance gene cassette was used to disrupt the FAD-GADH genes in the thermotolerant strains. The mutant strains produced 5KGA exclusively, and the final yields were over 90% at 30°C and 50% at 37°C. We found that the activity of pyrroloquinoline quinone (PQQ)-dependent glycerol dehydrogenase, which is responsible for 5KGA production, increased in response to addition of PQQ and CaCl2 in vitro when cells were grown at 37°C. Addition of 5 mM CaCl2 to the culture medium of the mutant strains increased 5KGA production to the point where over 90% of the initial substrate was converted. The thermotolerant Gluconobacter strains that we isolated in this study provide a promising new option for industrial 5KGA production.Gluconobacter is a genus of acetic acid bacteria that are able to oxidize a broad range of sugars, sugar alcohols, and sugar acids, and large amounts of the corresponding oxidized products accumulate in the culture medium. Such “incomplete” oxidation is carried out by membrane-bound enzymes, whose catalytic sites face the periplasm. These enzymes catalyze the dehydrogenization of d-glucose, d-sorbitol, d-mannitol, glycerol, d-gluconate, and the keto-d-gluconates. All of these enzymes are firmly attached to the cytoplasmic membrane, and the electrons abstracted from the substrates are passed on to ubiquinone and then to terminal ubiquinol oxidases, forming simple respiratory chains which create the membrane potential necessary to produce biological energy for these microorganisms.The oxidation of d-glucose to ketogluconates is known to be catalyzed by a series of enzymes. Pyrroloquinoline quinone (PQQ)-dependent glucose dehydrogenase oxidizes d-glucose to glucono-δ-lactone, and then gluconolactonase converts the glucono-δ-lactone to d-gluconate. The formation of ketogluconates in Gluconobacter strains has been reported to be catalyzed by two types of membrane-bound gluconate dehydrogenases (GADH) (10). One type is flavin adenine dinucleotide (FAD)-GADH, an FAD-containing, 2-keto-d-gluconate (2KGA)-producing enzyme, and the other type is a PQQ-containing, 5-keto-d-gluconate (5KGA)-producing enzyme. The former enzyme has three subunits: an FAD-containing dehydrogenase, a c-type cytochrome subunit containing three hemes, and a small subunit of unknown function (17). The latter enzyme, which produces 5KGA, is identical to the PQQ-containing polyol dehydrogenase (9), which is known as d-arabitol dehydrogenase (1), d-sorbitol dehydrogenase (20), or PQQ-dependent glycerol dehydrogenase (PQQ-GLDH) (2). PQQ-GLDH has broad substrate specificity but high regio- and stereospecificity, and it catalyzes reactions as predicted by the Bertrand-Hudson rule. This enzyme can oxidize d-gluconate only at the C-5 position to produce 5KGA from d-gluconate; however, the affinity of the enzyme for d-gluconate is quite low. The gene encoding this enzyme was cloned from Gluconobacter suboxydans IFO 3255 (11), and two open reading frames (ORFs) were found. One of these ORFs is believed to encode a hydrophobic protein with five membrane-spanning regions, and the other encodes a dehydrogenase subunit similar to that found in several PQQ-dependent enzymes, particularly the PQQ domain of membrane-bound glucose dehydrogenase. In contrast, 2KGA reductase and 5KGA reductase, the NADPH-dependent enzymes located in the cytoplasm, are thought to be involved in gluconate metabolism in the assimilation of 2KGA and 5KGA.5KGA is a useful raw material for the production of tartaric acid and xylaric acid and is used as a precursor for the synthesis of a number of flavor compounds, including 4-hydroxy-5- methyl-2,3-dihydrofuranone-3 (15). Moreover, it has been reported that 5KGA can be used to produce vitamin C by Gray''s method (6, 7), which is different from Reichstein''s method, which is now commonly used in industry. Reichstein''s method requires the use of high temperatures and an organic solvent in processing; however, Gray''s method does not.Most Gluconobacter strains produce both 2KGA and 5KGA from d-gluconate. Thus, production of 5KGA by Gluconobacter species generates 2KGA as a major by-product, and production of the two ketogluconates is competitive in vivo. Recently, an FAD-GADH-defective mutant strain of Gluconobacter oxydans 621 H which produced almost exclusively 5KGA from d-glucose was discovered (5). However, the optimum temperature for production of 5KGA in this mesophilic strain was around 20°C (19). For cost-effective industrial synthesis of 5KGA, we sought to develop a Gluconobacter strain which is able to produce 5KGA at higher temperatures, such as 37°C, in order to reduce the cost of cooling during fermentation.We successfully isolated thermotolerant Gluconobacter strains that are able to produce 5KGA at 37°C. We cloned the FAD-GADH gene and constructed FAD-GADH-defective mutants that produced almost exclusively 5KGA from d-gluconate at both ambient temperatures and higher temperatures up to 37°C. We believe that the thermotolerant strains reported in this study should be useful for industrial 5KGA production.  相似文献   

18.
Vancomycin-resistant enterococci acquire high-level resistance to glycopeptide antibiotics through the synthesis of peptidoglycan terminating in d-alanyl-d-lactate. A key enzyme in this process is a d-alanyl-d-alanine ligase homologue, VanA or VanB, which preferentially catalyzes the synthesis of the depsipeptide d-alanyl-d-lactate. We report the overexpression, purification, and enzymatic characterization of DdlN, a VanA and VanB homologue encoded by a gene of the vancomycin-producing organism Amycolatopsis orientalis C329.2. Evaluation of kinetic parameters for the synthesis of peptides and depsipeptides revealed a close relationship between VanA and DdlN in that depsipeptide formation was kinetically preferred at physiologic pH; however, the DdlN enzyme demonstrated a narrower substrate specificity and commensurately increased affinity for d-lactate in the C-terminal position over VanA. The results of these functional experiments also reinforce the results of previous studies that demonstrated that glycopeptide resistance enzymes from glycopeptide-producing bacteria are potential sources of resistance enzymes in clinically relevant bacteria.The origin of antibiotic resistance determinants is of significant interest for several reasons, including the prediction of the emergence and spread of resistance patterns, the design of new antimicrobial agents, and the identification of potential reservoirs for resistance elements. Antibiotic resistance can occur either through spontaneous mutation in the target or by the acquisition of external genetic elements such as plasmids or transposons which carry resistance genes (7). The origins of these acquired genes are varied, but it has long been recognized that potential reservoirs are antibiotic-producing organisms which naturally harbor antibiotic resistance genes to protect themselves from the actions of toxic compounds (6).High-level resistance to glycopeptide antibiotics such as vancomycin and teicoplanin in vancomycin-resistant enterococci (VRE) is conferred by the presence of three genes, vanH, vanA (or vanB), and vanX, which, along with auxiliary genes necessary for inducible gene expression, are found on transposons integrated into plasmids or the bacterial genome (1, 20). These three genes are essential to resistance and serve to change the C-terminal peptide portion of the peptidoglycan layer from d-alanyl-d-alanine (d-Ala-d-Ala) to d-alanyl-d-lactate (d-Ala-d-Lac). This change results in the loss of a critical hydrogen bond between vancomycin and the d-Ala-d-Ala terminus and in a 1,000-fold decrease in binding affinity between the antibiotic and the peptidoglycan layer, which is the basis for the bactericidal action of this class of compounds (5). The vanH gene encodes a d-lactate dehydrogenase which provides the requisite d-Lac (3, 5), while the vanX gene encodes a highly specific dd-peptidase which cleaves only d-Ala-d-Ala produced endogenously while leaving d-Ala-d-Lac intact (19, 21). The final gene, vanA or vanB, encodes an ATP-dependent d-Ala-d-Lac ligase (4, 8, 10). This enzyme has sequence homology with the chromosomal d-Ala-d-Ala ligases, which are essential for peptidoglycan synthesis but which generally lack the ability to synthesize d-Ala-d-Lac (9).We have recently cloned vanH, vanA, and vanX homologues from two glycopeptide antibiotic-synthesizing organisms: Amycolatopsis orientalis C329.2, which produces vancomycin, and Streptomyces toyocaensis NRRL 15009, which produces A47934 (14). In addition, the vanH-vanA-vanX gene cluster was identified in several other glycopeptide producers. We have also demonstrated that the VanA homologue from S. toyocaensis NRRL 15009 can synthesize d-Ala-d-Lac in vitro and in the glycopeptide-sensitive host Streptomyces lividans (15, 16). We now report the expression of the A. orientalis C329.2 VanA homologue DdlN in Escherichia coli, its purification, and its enzymatic characterization. These data reinforce the striking similarity between vancomycin resistance elements in VRE and glycopeptide-producing organisms and support the possibility of a common origin for these enzymes.

Expression, purification, and specificity of DdlN.

DdlN was overexpressed in E. coli under the control of the bacteriophage T7 promoter. The construct gave good yields of highly purified enzyme following a four-step purification procedure (Table (Table1;1; Fig. Fig.1).1). Like other dd-ligases, DdlN behaved like a dimer in solution (not shown).

TABLE 1

Purification of DdlN from E. coli BL21 (DE3)/pETDdlN
SampleProtein (mg)Activity (nmol/min)Sp act (nmol/ min/mg)Recovery (%)Purification (fold)
Lysate1248436.82100
Ammonium sulfate (20–50% saturation)67.678011.5921.7
Sephacryl S20011.682571.49811
Q Sepharose2.87422658839
Phenyl Superose0.429974835110
Open in a separate windowOpen in a separate windowFIG. 1Purification of DdlN from E. coli BL21 (DE3)/pETDdlN. Proteins were separated on an SDS–11% polyacrylamide gel and stained with Coomassie blue. Lane 1, molecular mass markers (masses are noted at the left in kilodaltons); lane 2, whole-cell lysate; lane 3, ammonium sulfate fraction (20 to 50% saturation); lane 4, Sephacryl S200; lane 5, Q Sepharose; lane 6, phenyl Superose.The amino acid substrate specificity of DdlN was assessed by incubation of 14C-d-Ala with all 20 common amino acids in the d configuration. Purified DdlN catalyzed the synthesis of d-Ala-d-Ala in addition to that of several other mixed dipeptides, including d-Ala-d-Met and d-Ala-d-Phe (Fig. (Fig.2).2). Thus, DdlN exhibits a substrate specificity which is similar to that of VanA (4), with the capacity to synthesize not only d-Ala-d-Ala but also mixed dipeptides with bulky side chains in the C-terminal position.Open in a separate windowFIG. 2Substrate specificity of DdlN. Autoradiogram from thin-layer chromatography analysis of DdlN substrate specificity. All reaction mixtures contained 2.5 mM d-Ala and 1 mM ATP, and the radiolabel was 14C-d-Ala, except where noted. Lane 1, d-Ala; lane 2, d-Lac with 14C-d-Lac label; lane 3, d,l-methionine; lane 4, dl-phenylalanine; lane 5, d-Hbut; lane 6, d-hydroxyvalerate. Letters indicate the following: A, d-Ala-d-Lac; B, d-Lac; C, d-Ala-d-Met; D, d-Ala-d-Phe; E, d-Ala-d-Hbut; F, d-Ala-d-hydroxyvalerate.Importantly, DdlN is a depsipeptide synthase with the ability to synthesize d-Ala-d-Lac, d-Ala-d-hydroxybutyrate (Hbut), and d-Ala-d-hydroxyvalerate (Fig. (Fig.2).2). However, unlike VanA (5), d-hydroxycaproate and d-phenyllactate are not substrates (not shown). Thus, DdlN is a broad-spectrum d-Ala-d-X ligase with depsipeptide synthase activity.

Characterization of d-Ala-d-X ligase activity.

Following the initial assessment of the specificity of the enzyme, several substrates were selected for quantitative analysis by evaluation of their steady-state kinetic parameters (Table (Table2).2). DdlN has two amino acid (or hydroxy acid) Km values. Steady-state kinetic plots indicated that, like other dd-ligases, the N-terminal Km (Km1) was significantly lower (higher specificity) than the C-terminal Km (Km2). Since the former value is expected to be independent of the C-terminal substrate, only Km2 values were determined and are reported here.

TABLE 2

Characterization of steady-state parameters of DdlN and VanA
LigaseSubstrateKm2 (mM)kcat (min−1)kcat/Km2 (M−1 s−1)
DdlNd-Ala21 ± 2229 ± 71.8 × 102
d-Lac0.4 ± 0.0555 ± 12.3 × 103
d-Hbut2.5 ± 0.332 ± 22.1 × 102
ATPa1.2 ± 0.271 ± 50.98 × 102
DdlMbd-Ala166 ± 27
d-Lac1.08 ± 0.10
VanAcd-Ala382951.3 × 102
d-Lac7.1942.2 × 102
d-Hbut0.601083.0 × 103
Open in a separate windowa Determined in the presence of 10 mM d-Lac. b Data from reference 16c Data from reference 5. DdlN showed good d-Ala-d-Ala ligase activity but with a very high and physiologically questionable Km2 (21 mM). On the other hand, d-Ala-d-Lac synthesis was excellent, with a 4-fold decrease in kcat, compared to d-Ala-d-Ala synthesis, which was offset by a 52-fold drop in Km that resulted in a >12-fold increase in specificity (kcat/Km2). d-Hbut was also a good substrate, with a kcat/Km2 comparable to that of d-Ala.Steady-state kinetic parameters for d-Ala-d-X formation showed trends similar to those found with both VanA and DdlN. For example, the kcat values between VanA and DdlN were virtually the same for most substrates. There were significant differences, however. For instance, while the Km2 values for d-Ala were very high for all three enzymes, DdlN does have greater affinity for d-Ala, with a 1.8- and 7.9-fold lower Km2 than those of VanA and DdlM, respectively. Additionally, the Km2 for d-Lac was 17.8- and 2.7-fold lower than those for VanA and DdlM. Thus, DdlN has a more restrictive specificity for the C-terminal residue than VanA, which is compensated for by a higher affinity for the critical substrate d-Lac.

pH dependence of peptide versus that of depsipeptide synthesis activity.

The partitioning of the syntheses of d-Ala-d-Ala and d-Ala-d-Hbut in VanA and other depsipeptide-competent dd-ligases has been shown to be pH dependent (17). Determination of the pH dependence of DdlN in synthesizing peptide versus depsipeptide (Fig. (Fig.3)3) directly paralleled the results obtained with VanA in similar experiments. At lower pHs (<7), d-Ala-d-Hbut synthesis predominates and is exclusive at a pH of <6 (Fig. (Fig.3).3). At pH 7.5, levels of synthesis of d-Ala-d-Hbut and d-Ala-d-Ala are relatively equal, while at a pH greater than 8, the capacity to synthesize peptide overtakes the capacity to synthesize depsipeptide, although the latter is never abolished. Open in a separate windowFIG. 3pH dependence of partitioning of the syntheses of peptide and depsipeptide by DdlN. (A) Autoradiogram of a thin-layer chromatography separation of the products of reaction mixtures containing 14C-D-Ala, unlabeled D-Ala, and d-Hbut. (B) Quantification of reaction products following phosphorimage analysis. Filled circles, D-Ala-d-Hbut; open circles, D-Ala-D-Ala.The partitioning of the formation of peptide versus depsipeptide as a function of pH by DdlM is comparable to that by VanA and depsipeptide-competent mutants of DdlB (17), which show essentially exclusively depsipeptide formation at lower pHs and increasing peptide formation as the pH increases. This implies a potential role for the protonated ammonium group of d-Ala2 in second-substrate recognition and suggests a mechanism for the discrimination between d-Ala and d-Lac at physiologic pH. The structural basis for this distinction remains obscure for DdlB and VanA or DdlN.

Concluding remarks.

Resistance to vancomycin and other glycopeptides is mediated through the synthesis of a peptidoglycan which does not terminate with the canonical d-Ala-d-Ala dipeptide. Thus, enterococci which exhibit the VanC phenotype, which consists of low-level, noninducible resistance to vancomycin only, have peptidoglycan terminating in d-Ala-d-Ser (19). On the other hand, bacteria which are constitutively resistant to high concentrations of glycopeptides, such as lactic acid bacteria and VRE exhibiting the VanA or VanB phenotype (high-level inducible resistance to vancomycin), incorporate the depsipeptide d-Ala-d-Lac into their cell walls (2, 12, 13). The enzymes responsible for the intracellular synthesis of d-Ala-d-Lac not surprisingly have significant amino acid sequence similarity with d-Ala-d-Ala ligases, which are responsible for d-Ala-d-Ala synthesis in all bacteria with a cell wall (9).The d-Ala-d-Lac synthases can be subdivided into two groups based on sequence homology: those found in the constitutively resistant lactic acid bacteria and those found in glycopeptide-producing organisms and VanA or VanB VRE (9, 14). The former have more similarity with exclusive d-Ala-d-Ala ligases. Indeed, single point mutations in d-Ala-d-Ala ligases which yield sequences more similar to those of lactic acid bacterium d-Ala-d-Lac ligases are sufficient to induce significant depsipeptide synthase activity in these enzymes (17). Similarly, mutational studies of the d-Ala-d-Lac ligase from Leuconostoc mesenteroides have demonstrated that the converse also holds (18). On the other hand, the molecular basis for depsipeptide synthesis by the VanA or VanB ligases is unknown, in large part due to the lack of protein structural information on which to base mutational studies, unlike the situation with d-Ala-d-Ala ligases, where the E. coli DdlB structure serves as a template for mechanistic research (11).Significantly, a major difference in the VanA or VanB ligases and other dd-ligases lies in the amino acid sequence of the ω-loop region, which closes off the active site of DdlB (11) and has been shown to contribute amino acid residues with the capacity to control the syntheses of d-Ala-d-Ala and d-Ala-d-Lac, notably, Tyr216 (17, 18). Until recently, the VanA and VanB ligases were exceptional in amino acid structure and had no known homologues. The sequencing of resistance genes from glycopeptide-producing bacteria has uncovered enzymes with >60% homology to VanA or VanB and which are virtually superimposable in the critical ω-loop region (14, 15). One of these, DdlM from S. toyocaensis NRRL 15009, has been shown to have d-Ala-d-Lac ligase ability (15, 16), although no rigorous analysis of this activity has been performed. The results presented here demonstrate that DdlN from the vancomycin producer A. orientalis C329.2 not only is a d-Ala-d-Lac ligase but also has significant functional homology with VanA. It is not known at present if, like S. toyocaensis NRRL 15009 (16), A. orientalis C329.2 also possess a d-Ala-d-Ala-exclusive ligase, though the presence of a vanX gene (14) suggests that it may.These studies demonstrate that DdlN cloned from a vancomycin-producing bacterium is a d-Ala-d-Lac ligase which has not only amino acid sequence homology with the dd-ligases from VRE but also functional homology. Thus, VanA, VanB, DdlN, and DdlM have likely evolved from similar origins. The fact that a vanH-vanA-vanX gene cluster can be found in other glycopeptide producers as well (14) suggests that the genes now found in VRE may have originated in glycopeptide-producing bacteria. Our finding that overexpressed, purified, DdlN shows many enzymatic characteristics similar (though not identical) to those of VanA suggests that the genes from glycopeptide-producing bacteria can be important in elucidating biochemical and protein structural aspects of the VRE proteins.  相似文献   

19.
20.
Bacteria which were β-d-galactosidase and β-d-glucuronidase positive or expressed only one of these enzymes were isolated from environmental water samples. The enzymatic activity of these bacteria was measured in 25-min assays by using the fluorogenic substrates 4-methylumbelliferyl-β-d-galactoside and 4-methylumbelliferyl-β-d-glucuronide. The enzyme activity, enzyme induction, and enzyme temperature characteristics of target and nontarget bacteria in assays aimed at detecting coliform bacteria and Escherichia coli were investigated. The potential interference of false-positive bacteria was evaluated. Several of the β-d-galactosidase-positive nontarget bacteria but none of the β-d-glucuronidase-positive nontarget bacteria contained unstable enzyme at 44.5°C. The activity of target bacteria was highly inducible. Nontarget bacteria were induced much less or were not induced by the inducers used. The results revealed large variations in the enzyme levels of different β-d-galactosidase- and β-d-glucuronidase-positive bacteria. The induced and noninduced β-d-glucuronidase activities of Bacillus spp. and Aerococcus viridans were approximately the same as the activities of induced E. coli. Except for some isolates identified as Aeromonas spp., all of the induced and noninduced β-d-galactosidase-positive, noncoliform isolates exhibited at least 2 log units less mean β-d-galactosidase activity than induced E. coli. The noncoliform bacteria must be present in correspondingly higher concentrations than those of target bacteria to interfere in the rapid assay for detection of coliform bacteria.Indicators of pollution (e.g., coliforms, fecal coliforms, and Escherichia coli) are traditionally used for monitoring the microbiological safety of water supplies and recreational water. Several techniques for detection of coliforms and E. coli are based on enzymatic hydrolysis of fluorogenic or chromogenic substrates for β-d-galactosidase and β-d-glucuronidase (9, 20). Current methods of recovery are usually culture based, and the analysis time is 18 to 24 h. In addition to enzymatic activity, these techniques use growth at appropriate temperatures in the presence of inhibitors, combined with demonstration of enzymatic activity, to selectively detect target bacteria.Rapid methods which require less than 6 h and are based on chromogenic, fluorogenic, or chemiluminogenic substrates for detection of coliforms, fecal coliforms, or E. coli have been described (13, 10, 27, 28). These rapid assays are based on the assumption that β-d-galactosidase and β-d-glucuronidase are markers for coliforms and E. coli, respectively. However, when the incubation time is 1 h or less, growth is not a selective step, and all β-d-galactosidase-positive or β-d-glucuronidase-positive microorganisms in a water sample contribute to the activity measured. At low initial concentrations of target bacteria (i.e., E. coli and total coliforms), increasing the preincubation time to 5 to 6 h did not result in a predominance of target bacteria compared to nontarget bacteria (28).The β-d-galactosidase or β-d-glucuronidase activity calculated per cultivable coliform or fecal coliform bacterium in environmental samples can be 1 to 2 log units higher than the activity per induced E. coli cell in pure culture (11, 26). The presence of active, noncultivable bacteria can be one reason for this. Studies of survival (7, 24, 25) and disinfection (26) of E. coli have shown that loss of cultivability does not necessarily result in a loss of β-d-galactosidase activity. The presence of false-positive bacteria can be another reason.β-d-Galactosidase has been found in numerous microorganisms, including gram-negative bacteria (e.g., strains belonging to the Enterobacteriaceae, Vibrionaceae, Pseudomonadaceae, and Neisseriaceae), several gram-positive bacteria, yeasts, protozoa, and fungi (17, 29). β-d-Glucuronidase is produced by most E. coli strains and also by other members of the Enterobacteriaceae, including some Shigella and Salmonella strains and a few Yersinia, Citrobacter, Edwardia, and Hafnia strains. Production of β-d-glucuronidase by Flavobacterium spp., Bacteroides spp., Staphylococcus spp., Streptococcus spp., anaerobic corynebacteria, and Clostridium has also been reported (12).High numbers of false-positive bacteria in sewage and contaminated water have been revealed by enumeration of β-d-galactosidase- and β-d-glucuronidase-positive CFU on nonselective agar supplemented with fluorogenic or chromogenic substrates (11, 28). Whether the activity from nontarget organisms can be neglected in a rapid assay depends on the number of nontarget organisms compared with the number of target bacteria and also on the level of their enzyme activity. Plant and algal biomass must be present at high concentrations to interfere in rapid bacterial β-d-galactosidase and β-d-glucuronidase assays (8).The main objective of this study was to investigate the enzyme characteristics of β-d-galactosidase- and β-d-glucuronidase-positive bacteria isolated from environmental water samples and to evaluate the potential influence of false-positive bacteria in rapid assays for coliform bacteria or E. coli in water. The effect of temperature on enzyme activity and on the interference of nontarget bacteria in the rapid assays was investigated as an important factor.(Some of the results were presented at the 97th General Meeting of the American Society for Microbiology 1997, Miami Beach, Fla., 4 to 8 May 1997.)  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号