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 46 毫秒
1.
Among mutants of Pseudomonas aeruginosa isolated from fluoroacetamide medium were some which synthesized amidase at about 5% of the rate of the parent constitutive strain, PAC101. Seven fluoroacetamide-resistant mutants with low amidase activity gave rise to secondary mutant strains on succinate+butyramide plates. One appeared to be an 'up-promotor' mutant and synthesized amidase at a high rate. This mutant, PAC433, was not stimulated by cyclic-AMP and was much less sensitive to catabolite repression by succinate. The mutation conferring resistance to catabolite repression was cotransduced at a frequency of 96% (26/27) with the amidase genes amiR, amiE. Five other revertants had catabolite repression-resistance mutations which were linked to the amidase genes and these also were probably promotor mutants. One strain had a mutation conferring resistance to catabolite repression which was unlinked to the amidase genes.  相似文献   

2.
A family of mutant amidases has been derived by experimental evolution of the aliphatic amidase of Pseudomonas aeruginosa strain PAC1. Mutation amiE16, in the structural gene for the enzyme, results in the production of the mutant B amidase by strain B6. This strain, unlike the wild-type, can utilize butyramide for growth. Strain B6 gave rise by a single mutational event to strain V9, utilizing valeramide, and strain PhB3, utilizing phenylacetamide. Strain V9 was not itself able to utilize phenylacetamide but gave rise by mutation to the phenylacetamide-utilizing mutant PhV1. Peptide 108 was isolated from chymotryptic digests of mutant amidases from strains B6, PhB3 and PhV1, but could not be detected in chymotryptic digests of the wild-type amidase. The sequence of peptide 108 was established as Met-Arg-His-Gly-Asp-Ile-Phe. Thermolytic digests of mutant amidases from strains B6, PhB3, PhV1 and V9 were compared with digests of the wild-type amidase. A peptide of the composition Met, Arg, His, Gly2, Asp3, Ile, Ser3, Thr, Val was found in the digest of the wild-type amidase and was replaced in the digests of the mutant amidases by a peptide of the composition Met, Arg, His, Gly2, Asp3, Ile, Ser3, Thr, Val, Phe. Mutation amiE16 is common to the four mutant enzymes and can be accounted for by the mutation Ser leads to Phe. The sequence of the chymotryptic peptide corresponds with the N-terminal sequence of the amidase protein, and can also be related to the thermolysin peptides. It is concluded that mutation amiE16 is a Ser leads to Phe change at position 7 from the N-terminus and the effect of this on the enzyme conformation is discussed.  相似文献   

3.
4.
5.
6.
Mutants of Pseudomonas aeruginosa were isolated that were acetamide-negative in growth phenotype at 41 degrees C and constitutive for amidase synthesis at 28 degrees C. Two mutants were derived from the magno-constitutive amidase mutant PAC111 (C11), and a third from a mutant that had enhanced inducibility by formamide, PAC153 (F6). The three temperature-sensitive mutants produced amidases with the same thermal stabilities as the wild-type enzyme. Cultures growing exponentially at 28 degrees C, synthesizing amidase constitutively, ceased amidase synthesis almost immediately on transfer to 41 degrees C. Cultures growing at 41 degrees C were transferred to 28 degrees C and had a lag of about 0.5 of a generation before amidase synthesis became detectable. Pulse-heating for 10 min at 45 degrees C of a culture growing exponentially at 28 degrees C resulted in a lag of about 0.5 of a generation before amidase synthesis recommenced after returning to 28 degrees C. Acetamide-negative mutants that were unable to synthesize amidase at any growth temperature were isolated from an inducible strain producing the mutant B amidase PAC398 (IB10). Two mutants were examined that gave revertants producing B amidase but with novel regulatory phenotypes. It is suggested that amidase synthesis is regulated by positive control exerted by gene amiR.  相似文献   

7.
Pseudomonas aeruginosa Ph1 is a mutant strain derived from strain AI3. The strain AI3 is able to use acetanilide as a carbon source through a mutation (T103I) in the amiE gene that encodes an aliphatic amidase (EC 3.5.1.4). The mutations in the amiE gene have been identified (Thr103Ile and Trp138Gly) by direct sequencing of PCR-amplified mutant gene from strain Ph1 and confirmed by sequencing the cloned PCR-amplified gene. Site-directed mutagenesis was used to alter the wild-type amidase gene at position 138 for Gly. The wild-type and mutant amidase genes (W138G, T103I-W138G, and T103I) were cloned into an expression vector and these enzymes were purified by affinity chromatography on epoxy-activated Sepharose 6B-acetamide/phenylacetamide followed by gel filtration chromatography. Altered amidases revealed several differences in kinetic properties, namely, in substrate specificity, sensitivity to urea, optimum pH, and enzyme stability, compared with the wild-type enzyme. The W138G enzyme acted on acetamide, acrylamide, phenylacetamide, and p-nitrophenylacetamide, whereas the double mutant (W138G and T103I) amidase acted only on p-nitrophenylacetamide and phenylacetamide. On the other hand, the T103I enzyme acted on p-nitroacetanilide and acetamide. The heat stability of altered enzymes revealed that they were less thermostable than the wild-type enzyme, as the mutant (W138G and W138G-T103I) enzymes exhibited t 1/2 values of 7.0 and 1.5 min at 55°C, respectively. The double substitution T103I and W138G on the amidase molecule was responsible for increased instabiliby due to a conformational change in the enzyme molecule as detected by monoclonal antibodies. This conformational change in altered amidase did not alter its M r value and monoclonal antibodies reacted differently with the active and inactive T103I-W138G amidase.  相似文献   

8.
A Bell  M Bains    R E Hancock 《Journal of bacteriology》1991,173(21):6657-6664
Overexpression of major outer membrane protein OprH of Pseudomonas aeruginosa as a result of mutation (in strain H181) or adaptation to low Mg2+ concentrations (in parent strain H103) is accompanied by increased resistance to polymyxin B, gentamicin, and EDTA. A 2.8-kb EcoRI fragment containing the oprH gene was subcloned into several different expression plasmids in Escherichia coli. These experiments showed that significant levels of OprH could be produced from a promoter on the EcoRI fragment; that the cloned oprH gene was not regulated by Mg2+ deficiency; that there were no differences in the expression of OprH in any construction, regardless of whether the gene from strain H103 or its OprH-overexpressing, polymyxin B-resistant derivative, strain H181, was used; and that overexpression of OprH in E. coli to the level observed in P. aeruginosa H181 did not result in a resistance phenotype. These results favored the conclusion that the mutation in strain H181 was a regulatory rather than a promoter mutation. The oprH gene was cloned behind the benzoate-inducible pm promoter in plasmid pGB25 and transferred to P. aeruginosa H103. Overexpression of OprH from the cloned gene in H103/pGB25 resulted in EDTA resistance but not polymyxin B resistance. This result suggested that another factor, possibly lipopolysaccharide, was affected by the mutation in strain H181. Consistent with this suggestion was the demonstration that mutants of strain H181 with alterations in lipopolysaccharide had reverted to wild-type polymyxin B susceptibility but had unaltered gentamicin and EDTA resistance. These data were consistent with the hypothesis that OprH replaces outer membrane-stabilizing divalent cations.  相似文献   

9.
The time-dependent inhibition of amidase from Pseudomonas aeruginosa strain AI 3 by urea, hydroxyurea and cyanate displayed saturation kinetics fitting a model for the reaction sequence in which formation of a complex in a reversible step was followed by an irreversible step. Altered amidases from mutant strains AIU 1N and OUCH 4, selected for their resistance to inhibition of growth by urea and hydroxyurea respectively, had altered kinetic constants for inhibition indicating reduced binding capacity for the inhibitors. The substrate acetamide protected AI 3 amidase against inhibition by urea,.and altered Ki values for inhibition of the mutant amidases were paralleled by alterations in Km values for acetamide indicating that urea acted at the active site. Inhibition of AI 3 amidase involved the binding of one molecule of urea per molecule of enzyme. Urea inhibited amidase slowly regained activity at pH 7.2 through release of urea.  相似文献   

10.
The invasion of Pseudomonas aeruginosa and Salmonella enterica serovar Typhi into epithelial cells depends on the cystic fibrosis transmembrane conductance regulator (CFTR) protein as an epithelial receptor. In the case of P. aeruginosa , the bacterial ligand for CFTR is the outer core oligosaccharide portion of the lipopolysaccharide (LPS). To determine whether serovar Typhi LPS is also a bacterial ligand mediating internalization, we used both P. aeruginosa and serovar Typhi LPS as a competitive inhibitor of serovar Typhi invasion into the epithelial cell line T84. P. aeruginosa LPS containing a complete core efficiently inhibited serovar Typhi invasion. However, neither killed wild-type Typhi cells nor purified LPS were effective inhibitors. LPS from mutant Typhi strains defective in O side-chain synthesis, but with an apparently normal core, was capable of inhibiting invasion, but LPS obtained from a deeper rough mutant strain with alterations in fast-migrating core oligosaccharide failed to inhibit invasion. Lastly, exposure of wild-type serovar Typhi to T84 cultures before heat killing resulted in a structural alteration in its LPS that allowed the heat-killed cells to inhibit invasion of wild-type serovar Typhi. These data indicate that the serovar Typhi LPS core, like the P. aeruginosa LPS core, is a ligand mediating internalization of bacteria by epithelial cells, and that exposure of this ligand on wild-type Typhi is induced by the bacteria's interaction with host cells.  相似文献   

11.
A novel assay method was investigated for wild-type and recombinant mutant amidases (EC 3.5.1.4) from Pseudomonas aeruginosa by ammonium ion-selective electrode (ISE). The initial velocity is proportional to the enzyme concentration by using the wild-type enzyme. The specific activities of the purified amidase were found to be 88.2 and 104.2 U mg protein(-1) for the linked assay and ISE methods, respectively. The kinetic constants--Vmax, Km, and Kcat--determined by Michaelis-Menten plot were 101.13 U mg protein(-1), 1.12x10(-2) M, and 64.04 s(-1), respectively, for acrylamide as the substrate. On the other hand, the lower limit of detection and range of linearity of enzyme concentration were found to be 10.8 and 10.8 to 500 ng, respectively, for the linked assay method and 15.0 and 15.0 to 15,000 ng, respectively, for the ISE method. Hydroxylamine was found to act as an uncompetitive activator of hydrolysis reaction catalyzed by amidase given that there is an increase in Vmax and Km when acetamide was used as the substrate. However, the effect of hydroxylamine on the hydrolysis reaction was dependent on the type of amidase and substrate involved in the reaction mixture. The degrees of activation (epsilon(a)) of the wild-type and mutant (T103I and C91A) enzymes were found to be 2.54, 12.63, and 4.33, respectively, for acetamide as the substrate. However, hydroxylamine did not activate the reaction catalyzed by wild-type and altered (C91A and W138G) amidases by using acrylamide and acetamide, respectively, as the substrate. The activating effect of hydroxylamine on the hydrolysis of acetamide, acrylamide, and p-nitrophenylacetamide can be explained by the fact that additional formation of ammonium ions occurred due to the transferase activity of amidases. However, the activating effect of hydroxylamine on the hydrolysis of p-nitroacetanilide may be due to a change in conformation of enzyme molecule. Therefore, the use of ISE permitted the study of the kinetic properties of wild-type and mutant amidases because it was possible to measure initial velocity of the enzyme-catalyzed reaction in real time.  相似文献   

12.
Lipopolysaccharides (LPS) were isolated from rough-type mutant strains of Pseudomonas aeruginosa (Delta algC) derived from wild-type strains PAO1 (serogroup O5) and PAC1R (serogroup O3). Structural studies of the LPS core region with a special focus on the phosphorylation pattern were performed by 2D NMR spectroscopy, including a 1H,(31)P HMQC-TOCSY experiment, MALDI-TOF MS, and Fourier-transform ion cyclotron resonance ESIMS using the capillary skimmer dissociation technique. Both LPS were found to contain two residues each of 3-deoxy-D-manno-oct-2-ulosonic acid (Kdo) and L-glycero-D-manno-heptose (Hep), one residue of N-(L-alanyl)-D-galactosamine and one O-carbamoyl group (Cm) on the distal Hep residue. The following structures of a tetrasaccharide trisphosphate from strain PAC1R Delta algC and that carrying an additional ethanolamine phosphate group (PEtN) from strain PAO1 Delta algC were elucidated: [carbohydrate structre: see text] where R=P in PAC1R Delta algC and PPEtN in PAO1 Delta algC. To our knowledge, in this work the presence of ethanolamine diphosphate is unambiguously confirmed and its position established for the first time in the LPS core of a rough-type strain of P. aeruginosa. In addition, the structure of the complete LPS core of wild-type strain P. aeruginosa PAO1 was reinvestigated and the position of the phosphorylation sites was revised.  相似文献   

13.
14.
This study demonstrates the use of Fourier transform infrared (FTIR) spectroscopy for monitoring both synthesis and hydrolysis reactions catalyzed by a recombinant amidase (EC 3.5.1.4) from Pseudomonas aeruginosa. The kinetics of hydrolysis of acetamide, propionamide, butyramide, acrylamide, benzamide, phenylalaninamide, alaninamide, glycinamide, and leucinamide were determined. This revealed that very short-chain substrates displayed higher amidase activity than did branched side-chain or aromatic substrates. In addition, on reducing the polarity and increasing the substrates' bulkiness, a reduction of the amidase affinity for the substrates took place. Using FTIR spectroscopy it was possible to monitor and quantify the synthesis of several hydroxamic acid derivatives and ester hydrolysis products. These products may occur simultaneously in a reaction catalyzed by the amidase. The substrates used for the study of such reactions were ethyl acetate and glycine ethyl ester. Hydroxylamine was the nucleophile substrate used for the synthesis of acetohydroxamate compounds. Results presented in this article demonstrate the usefulness of FTIR spectroscopy as an important tool for understanding the enzyme structure-activity relationship because it provides a simple and rapid real-time assay for the detection and quantification of amidase hydrolysis and synthesis reactions in situ.  相似文献   

15.
A mutant strain, KLAM59, of Pseudomonas aeruginosa has been isolated that synthesizes a catalytically inactive amidase. The mutation in the amidase gene has been identified (Glu59Val) by direct sequencing of PCR-amplified mutant gene and confirmed by sequencing the cloned PCR-amplified gene. The wild-type and altered amidase genes were cloned into an expression vector and both enzymes were purified by affinity chromatography on epoxy-activated Sepharose 6B-acetamide followed by gel filtration chromatography. The mutant enzyme was catalytically inactive, and it was detected in column fractions by monoclonal antibodies previously raised against the wild-type enzyme using an ELISA sandwich method. The recombinant wild-type and mutant enzymes were purified with a final recovery of enzyme in the range of 70–80%. The wild-type and mutant enzymes behaved differently on the affinity column as shown by their elution profiles. The molecular weights of the purified wild-type and mutant amidases were found to be 210,000 and 78,000 Dalton, respectively, by gel filtration chromatography. On the other hand, the mutant enzyme ran as a single protein band on SDS-PAGE and native PAGE with a M r of 38,000 and 78,000 Dalton, respectively. These data suggest that the substitution Glu59Val was responsible for the dimeric structure of the mutant enzyme as opposed to the hexameric form of the wild-type enzyme. Therefore, the Glu59 seems to be a critical residue in the maintenance of the native quaternary structure of amidase.  相似文献   

16.
Pseudomonas putida PP3 carrying dehalogenases I and II and Pseudomonas aeruginosa PAU3 carrying dehalogenase I coded for by plasmid pUU2 were able to grow on 2-monochloropropionic acid (2MCPA). Neither strain utilized 2-chloropropionamide (2CPA) as a carbon or nitrogen source for growth. Mutations in both strains to 2Cpa+ phenotypes (designated P. putida PPW3 and P. aeruginosa PAU5, respectively) involved the expression of an acquired 2CPA-amidase activity. The amidase followed by dehalogenase reactions in these strains constituted a novel metabolic pathway for growth on 2CPA. P. putida PPW3 synthesized a constitutive amidase of molecular mass 59 kDa consisting of two identical subunits of 29 kDa. For those amides tested this acquired enzyme was most active against chlorinated aliphatic amides, although substrate affinities (Km) and maximum rates of activity (Vmax) were poor. P. aeruginosa PAU5 acquired a 2Cpa+ phenotype by overproducing the A-amidase normally used by this species to hydrolyse aliphatic amides. The A-amidase had only slight activity towards 2CPA. However, with constitutive synthesis the mutant grew on the chlorinated substrates. Chloroacetamide (CAA) was a toxic substrate analogue for these Pseudomonas strains. A strain resistant to CAA was isolated from P. aeruginosa PAU5 when exposed to 1-10 mM-CAA. This mutant, P. aeruginosa PAU6, synthesized an inducible A-amidase. CAA-resistance depended upon the simultaneous expression of CAA-inducible amidase and dehalogenase activities.  相似文献   

17.
A constitutively expressed aliphatic amidase from a Rhodococcus sp. catalyzing acrylamide deamination was purified to electrophoretic homogeneity. The molecular weight of the native enzyme was estimated to be 360,000. Upon sodium dodecyl sulfate-polyacrylamide gel electrophoresis, the purified preparation yielded a homogeneous protein band having an apparent molecular weight of about 44,500. The amidase had pH and temperature optima of 8.5 and 40 degrees C, respectively, and its isoelectric point was pH 4.0. The amidase had apparent K(m) values of 1.2, 2.6, 3.0, 2.7, and 5.0 mM for acrylamide, acetamide, butyramide, propionamide, and isobutyramide, respectively. Inductively coupled plasma-atomic emission spectometry analysis indicated that the enzyme contains 8 mol of iron per mol of the native enzyme. No labile sulfide was detected. The amidase activity was enhanced by, but not dependent on Fe(2+), Ba(2+), and Cr(2+). However, the enzyme activity was partially inhibited by Mg(2+) and totally inhibited in the presence of Ni(2+), Hg(2+), Cu(2+), Co(2+), specific iron chelators, and thiol blocking reagents. The NH2-terminal sequence of the first 18 amino acids displayed 88% homology to the aliphatic amidase of Brevibacterium sp. strain R312.  相似文献   

18.
Zhao G  Song H  Chen ZW  Mathews FS  Jorns MS 《Biochemistry》2002,41(31):9751-9764
Conservative mutation of His269 (to Asn, Ala, or Gln) does not-significantly affect the expression of monomeric sarcosine oxidase (MSOX), covalent flavinylation, the physicochemical properties of bound FAD, or the overall protein structure. Turnover with sarcosine and the limiting rate of the reductive half-reaction with L-proline at pH 8.0 are, however, nearly 2 orders of magnitude slower than that with with wild-type MSOX. The crystal structure of the His269Asn complex with pyrrole-2-carboxylate shows that the pyrrole ring of the inhibitor is displaced as compared with wild-type MSOX. The His269 mutants all form charge-transfer complexes with pyrrole-2-carboxylate or methylthioacetate, but the charge-transfer bands are shifted to shorter wavelengths (higher energy) as compared with wild-type MSOX. Both wild-type MSOX and the His269Asn mutant bind the zwitterionic form of L-proline. The E(ox).L-proline complex formed with the His269Asn mutant or wild-type MSOX contains an ionizable group (pK(a) = 8.0) that is required for conversion of the zwitterionic L-proline to the reactive anionic form, indicating that His269 is not the active-site base. We propose that the change in ligand orientation observed upon mutation of His269 results in a less than optimal overlap of the highest occupied orbital of the ligand with the lowest unoccupied orbital of the flavin. The postulated effect on orbital overlap may account for the increased energy of charge-transfer bands and the slower rates of electron transfer observed for mutant enzyme complexes with charge-transfer ligands and substrates, respectively.  相似文献   

19.
Late infantile neuronal ceroid lipofuscinosis (LINCL) is caused by the deficiency of the lysosomal tripeptidyl peptidase-I encoded by CLN2. We previously detected in two LINCL patients a homozygous missense mutation, p.Asn286Ser, that affects a potential N-glycosylation site. We introduced the p.Asn286Ser mutation into the wild-type CLN2 cDNA and performed transient expression analysis to determine the effect on the catalytic activity, intracellular targeting, and glycosylation of the CLN2 protein. Expression of mutant p.Asn286Ser CLN2 in HEK293 cells revealed that the mutant was enzymatically inactive. Western blot analysis demonstrated that at steady state the amounts of expressed p.Asn286Ser CLN2 were reduced compared with wild-type expressing cells. The rate of synthesis and the sorting of the newly synthesized p.Asn286Ser CLN2 in the Golgi was not affected compared with wild-type CLN2 protein. The electrophoretic mobility of the immunoprecipitated mutant p.Asn286Ser CLN2 was increased by approximately 2 kDa compared with the wild-type CLN2 protein, whereas deglycosylation led to the generation of polypeptides of the same apparent size. The data suggest that mutant p.Asn286Ser CLN2 lacks one oligosaccharide chain resulting in enzymatic inactivation.  相似文献   

20.
Miguel V  Monti MR  Argaraña CE 《DNA Repair》2008,7(11):1799-1808
The Escherichia coli DNA Mismatch Repair (MMR) protein MutS exist as dimers and tetramers in solution, and the identification of its functional oligomeric state has been matter of extensive study. In the present work, we have analyzed the oligomerization state of MutS from Pseudomonas aeruginosa a bacterial species devoid of Dam methylation and MutH homologue. By analyzing native MutS and different mutated versions of the protein, we determined that P. aeruginosa MutS is mainly tetrameric in solution and that its oligomerization capacity is conducted as in E. coli, by the C-terminal region of the protein. The analysis of mismatch oligonucleotide binding activity showed that wild-type MutS binds to DNA as tetramer. The DNA binding activity decreased when the C-terminal region was deleted (MutSDelta798) or when a full-length MutS with tetramerization defects (MutSR842E) was tested. The ATPase activity of MutSDelta798 was similar to MutSR842E and diminished respect to the wild-type protein. Experiments carried out on a P. aeruginosa mutS strain to test the proficiency of different oligomeric versions of MutS to function in vivo showed that MutSDelta798 is not functional and that full-length dimeric version MutSR842E, is not capable of completely restoring the MMR activity of the mutant strain. Additional experiments carried out in conditions of high mutation rate induced by the base analogue 2-AP confirm that the dimeric version of MutS is not as efficient as the tetrameric wild-type protein to prevent mutations. Therefore, it is concluded that although dimeric MutS is sufficient for MMR activity, optimal activity is obtained with the tetrameric version of the protein and therefore it should be considered as the active form of MutS in P. aeruginosa.  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