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 15 毫秒
1.
Translation of the isoleucine codon AUA in most prokaryotes requires a modified C (lysidine or agmatidine) at the wobble position of tRNA2Ile to base pair specifically with the A of the AUA codon but not with the G of AUG. Recently, a Bacillus subtilis strain was isolated in which the essential gene encoding tRNAIle-lysidine synthetase was deleted for the first time. In such a strain, C34 at the wobble position of tRNA2Ile is expected to remain unmodified and cells depend on a mutant suppressor tRNA derived from tRNA1Ile, in which G34 has been changed to U34. An important question, therefore, is how U34 base pairs with A without also base pairing with G. Here, we show (i) that unlike U34 at the wobble position of all B. subtilis tRNAs of known sequence, U34 in the mutant tRNA is not modified, and (ii) that the mutant tRNA binds strongly to the AUA codon on B. subtilis ribosomes but only weakly to AUG. These in vitro data explain why the suppressor strain displays only a low level of misreading AUG codons in vivo and, as shown here, grows at a rate comparable to that of the wild-type strain.  相似文献   

2.
Formate dehydrogenases (FDHs) are frequently used for the regeneration of cofactors in biotransformations employing NAD(P)H-dependent oxidoreductases. Major drawbacks of most native FDHs are their strong preference for NAD+ and their low operational stability in the presence of reactive organic compounds such as α-haloketones. In this study, the FDH from Mycobacterium vaccae N10 (MycFDH) was engineered in order to obtain an enzyme that is not only capable of regenerating NADPH but also stable toward the α-haloketone ethyl 4-chloroacetoacetate (ECAA). To change the cofactor specificity, amino acids in the conserved NAD+ binding motif were mutated. Among these mutants, MycFDH A198G/D221Q had the highest catalytic efficiency (k cat/K m) with NADP+. The additional replacement of two cysteines (C145S/C255V) not only conferred a high resistance to ECAA but also enhanced the catalytic efficiency 6-fold. The resulting quadruple mutant MycFDH C145S/A198G/D221Q/C255V had a specific activity of 4.00?±?0.13 U?mg?1 and a K m, NADP + of 0.147?±?0.020 mM at 30 °C, pH 7. The A198G replacement had a major impact on the kinetic constants of the enzyme. The corresponding triple mutant, MycFDH C145S/D221Q/C255V, showed the highest specific activity reported to date for a NADP+-accepting FDH (v max, 10.25?±?1.63 U?mg?1). However, the half-saturation constant for NADP+ (K m, NADP + , 0.92?±?0.10 mM) was about one order of magnitude higher than the one of the quadruple mutant. Depending on the reaction setup, both novel MycFDH variants could be useful for the production of the chiral synthon ethyl (S)-4-chloro-3-hydroxybutyrate [(S)-ECHB] by asymmetric reduction of ECAA with NADPH-dependent ketoreductases.  相似文献   

3.
The plasmid pGKE75-catA138T, which comprises pUC18 and the catA138T gene encoding thermostable chloramphenicol acetyltransferase with an A138T amino acid replacement (CATA138T), serves as an Escherichia coli-Geobacillus kaustophilus shuttle plasmid that confers moderate chloramphenicol resistance on G. kaustophilus HTA426. The present study examined the thermoadaptation-directed mutagenesis of pGKE75-catA138T in an error-prone thermophile, generating the mutant plasmid pGKE75αβ-catA138T responsible for substantial chloramphenicol resistance at 65°C. pGKE75αβ-catA138T contained no mutation in the catA138T gene but had two mutations in the pUC replicon, even though the replicon has no apparent role in G. kaustophilus. Biochemical characterization suggested that the efficient chloramphenicol resistance conferred by pGKE75αβ-catA138T is attributable to increases in intracellular CATA138T and acetyl-coenzyme A following a decrease in incomplete forms of pGKE75αβ-catA138T. The decrease in incomplete plasmids may be due to optimization of plasmid replication by RNA species transcribed from the mutant pUC replicon, which were actually produced in G. kaustophilus. It is noteworthy that G. kaustophilus was transformed with pGKE75αβ-catA138T using chloramphenicol selection at 60°C. In addition, a pUC18 derivative with the two mutations propagated in E. coli at a high copy number independently of the culture temperature and high plasmid stability. Since these properties have not been observed in known plasmids, the outcomes extend the genetic toolboxes for G. kaustophilus and E. coli.  相似文献   

4.
Substrate inhibition hinders chitinolytic β-N-acetyl-d-hexosaminidases in producing N-acetyl-d-glucosamine (GlcNAc), the valuable chemical widely applied in medical and food industries. Here we focused on a promising chitinolytic enzyme, OfHex1 from the insect, Ostrinia furnacalis. By structural analysis of OfHex1, five residues nearby the active pocket including V327, E328, Y471, V484 and W490 were chosen and nine mutants including V327G, E328Q, E328A, Y471V, V484R, W490A, W490H, V327G/V484R/W490A and V327G/Y471V/W490H were constructed and recombinantly expressed in Pichia pastoris. The best-performing mutant, W490A, obtained by a higher yield of 5 mg/L, did not show substrate inhibition even when 5 mM of the substrates, (GlcNAc)2–4, were applied. The kcat/Km values for (GlcNAc)2–4 are 239.8, 111.3 and 79.8 s?1 mM?1, respectively. Besides, the pH stability of the mutant ranges from pH 4 to 11 and the thermal stability is up to 50 °C. This work suggests the W490A mutant might be an ideal biocatalyst for GlcNAc production from chitin.  相似文献   

5.
6.
Nup159p/Rat7p is an essential FG repeat–containing nucleoporin localized at the cytoplasmic face of the nuclear pore complex (NPC) and involved in poly(A)+ RNA export and NPC distribution. A detailed structural–functional analysis of this nucleoporin previously demonstrated that Nup159p is anchored within the NPC through its essential carboxyl-terminal domain. In this study, we demonstrate that Nup159p specifically interacts through this domain with both Nsp1p and Nup82p. Further analysis of the interactions within the Nup159p/Nsp1p/Nup82p subcomplex using the nup82Δ108 mutant strain revealed that a deletion within the carboxyl-terminal domain of Nup82p prevents its interaction with Nsp1p but does not affect the interaction between Nup159p and Nsp1p. Moreover, immunofluorescence analysis demonstrated that Nup159p is delocalized from the NPC in nup82Δ108 cells grown at 37°C, a temperature at which the Nup82Δ108p mutant protein becomes degraded. This suggests that Nup82p may act as a docking site for a core complex composed of the repeat-containing nucleoporins Nup159p and Nsp1p. In vivo transport assays further revealed that nup82Δ108 and nup159-1/rat7-1 mutant strains have little if any defect in nuclear protein import and protein export. Together our data suggest that the poly(A)+ RNA export defect previously observed in nup82 mutant cells might be due to the loss from the NPCs of the repeat-containing nucleoporin Nup159p.  相似文献   

7.
Jasmonate (JA) regulates various plant defense and developmental processes. The F-box protein CORONATINE INSENSITIVE 1 (COI1) perceives JA signals to mediate diverse plant responses including male fertility, root growth, anthocyanin accumulation, and defense against abiotic and biotic stresses. In this study, we carried out genetic, physiological and biochemical analysis on a series of coi1 mutant alleles, and found that different amino acid mutations in COI1 distinctively affect JA-regulated male fertility in Arabidopsis. All the JA responses are disrupted by the COI1 mutations W467* in coi1-1, Q343* (coi1-6), G369E (coi1-4), G98D (coi1-5), G155E (coi1-7), D452A (coi1-9) and L490A (coi1-10), though the coi1-5 mutant (COI1G98D) contains adequate COI1 protein (~60% of wild-type). Interestingly, the low basal level of COI1E543K in the coi1-8 mutant (~10% of wild-type COI1 level) is sufficient for maintaining male fertility (~50% of wild-type fertility); the coi1-2 mutant with low level of COI1L245F (~10% of wild-type) is male sterile under normal growth condition (22°C) but male fertile (~80% of wild-type fertility) at low temperature (16°C); however, both coi1-2 and coi1-8 are defective in the other JA responses (root growth, anthocyanin accumulation, and plant response to the pathogen Pst DC3000 infection).  相似文献   

8.
Two mutant strains of Aspergillus parasiticus, both deficient in aflatoxin production, were used to elucidate the biosynthetic pathway of this mycotoxin. One of the mutants, A. parasiticus ATCC 24551, was capable of accumulating large amounts of averufin, and the other, A. parasiticus 1-11-105 wh-1, accumulated versicolorin A. The averufin producing mutant efficiently converted 14C-labeled versiconal acetate, versicolorin A, and sterigmatocystin into aflatoxin B1 and G1, indicating that averufin preceded these compounds in the aflatoxin biosynthetic pathway. In the presence of dichlorvos (dimethyl 2,2-dichlorovinyl phosphate), a known inhibitor of aflatoxin biosynthesis, the conversion of versicolorin A and sterigmatocystin was unaffected, but the conversion of versiconal acetate was markedly inhibited. The mutant accumulating versicolorin A incorporated 14C-labeled acetate, averufin, and versiconal acetate into versicolorin A. In the presence of dichlorvos, however, the major conversion product was versiconal acetate. This strongly suggested that dichlorvos inhibited the conversion step of versiconal acetate into versicolorin A. This mutant resumed production of aflatoxin B1 if sterigmatocystin was added to the resting cell cultures, indicating that the mutant was blocked at the enzymatic step catalyzing the conversion of versicolorin A into sterigmatocystin, and as a result was incapable of aflatoxin production. The experimental evidence is thus provided for the involvement and interrelationship of three anthraquinones (averufin, versiconal acetate, and versicolorin A) and a xanthone (sterigmatocystin) in aflatoxin biosynthesis. A pathway for the biosynthesis of aflatoxin B1 is proposed to be: acetate →→→ averufin → versiconal acetate → versicolorin A → sterigmatocystin → aflatoxin B1.  相似文献   

9.
A chimeric enzyme (GST121) of the human α-glutathione S-transferases GST1-1 and GST2-2, which has improved catalytic efficiency and thermostability from its wild-type parent proteins, has been crystallized in a space group that is isomorphous with that reported for crystals of GST1-1. However, a single-site (G82R) mutant of GST121, which exhibits a significant reduction both in vitro and in vivo in protein thermostability, forms crystals that are not isomorphous with GST1-1. The mutant protein crystallizes in space group P212121, with cell dimensions a = 49.5, b = 92.9, c = 115.9 Å, and one dimer per asymmetric unit. Preliminary crystallographic results show that a mutation of the surface residue Gly 82 from a neutral to a charged residue causes new salt bridges to be formed among the GST dimers, suggesting that the G82R mutant might aggregate more readily than does GST121 in solution resulting in a change of its solution properties. © 1994 Wiley-Liss, Inc.  相似文献   

10.
Human glucose 6-phosphate dehydrogenase (G6PD) has both the “catalytic” NADP+ site and a “structural” NADP+ site where a number of severe G6PD deficiency mutations are located. Two pairs of G6PD clinical mutants, G6PDWisconsin (R393G) and G6PDNashville (R393H), and G6PDFukaya (G488S) and G6PDCampinas (G488V), in which the mutations are in the vicinity of the “structural” NADP+ site, showed elevated Kd values of the “structural” NADP+, ranging from 53 nM to 500 nM compared with 37 nM for the wild-type enzyme. These recombinant enzymes were denatured by Gdn-HCl and refolded by rapid dilution in the presence of l-Arg, NADP+ and DTT at 25 °C. The refolding yields of the mutants exhibited strong NADP+-dependence and ranged from 1.5% to 59.4% with 1000 μM NADP+, in all cases lower than the figure of 72% for the wild-type enzyme. These mutant enzymes also displayed decreased thermostability and high susceptibility to chymotrypsin digestion, in good agreement with their corresponding melting temperatures in CD experiments. Taken together, the results support the view that impaired binding of “structural” NADP+ can hinder folding as well as cause instability of these clinical mutant enzymes in the fully folded state.  相似文献   

11.
Octaketide synthase (OKS) from Aloe arborescens is a plant-specific type III polyketide synthase (PKS) that catalyzes iterative condensations of eight molecules of malonyl-CoA to produce the C16 aromatic octaketides SEK4 and SEK4b. On the basis of the crystal structures of OKS, the F66L/N222G double mutant was constructed and shown to produce an unnatural dodecaketide TW95a by sequential condensations of 12 molecules of malonyl-CoA. The C24 naphthophenone TW95a is a product of the minimal type II PKS (whiE from Streptomyces coelicolor), and is structurally related to the C20 decaketide benzophenone SEK15, the product of the OKS N222G point mutant. The C24 dodecaketide naphthophenone TW95a is the first and the longest polyketide scaffold generated by a structurally simple type III PKS. A homology model predicted that the active-site cavity volume of the F66L/N222G mutant is increased to 748 Å3, from 652 Å3 of the wild-type OKS. The structure-based engineering thus greatly expanded the catalytic repertoire of the simple type III PKS to further produce larger and more complex polyketide molecules.  相似文献   

12.
Under stress conditions that bring about excessive absorption of light energy in the chloroplasts, the formation of singlet oxygen (1O2) can be strongly enhanced, triggering programmed cell death. However, the 1O2 signaling pathway can also lead to acclimation to photooxidative stress, when 1O2 is produced in relatively low amounts. This acclimatory response is associated with a strong downregulation of the jasmonate biosynthesis pathway and the maintenance of low jasmonate levels, even under high light stress conditions that normally induce jasmonate synthesis. These findings suggest a central role for this phytohormone in the orientation of the 1O2 signaling pathway toward cell death or acclimation. This conclusion is confirmed here in an Arabidopsis double mutant obtained by crossing the 1O2-overproducing mutant ch1 and the jasmonate-deficient mutant dde2. This double mutant was found to be constitutively resistant to 1O2 stress and to display a strongly stimulated growth rate compared with the single ch1 mutant. However, the involvement of other phytohormones, such as ethylene, cannot be excluded.  相似文献   

13.
An infectious chimeric feline immunodeficiency virus (FIV)/HIV strain carrying six HIV-like protease (PR) mutations (I37V/N55M/V59I/I98S/Q99V/P100N) was subjected to selection in culture against the PR inhibitor lopinavir (LPV), darunavir (DRV), or TL-3. LPV selection resulted in the sequential emergence of V99A (strain S-1X), I59V (strain S-2X), and I108V (strain S-3X) mutations, followed by V37I (strain S-4X). Mutant PRs were analyzed in vitro, and an isogenic virus producing each mutant PR was analyzed in culture for LPV sensitivity, yielding results consistent with the original selection. The 50% inhibitory concentrations (IC50s) for S-1X, S-2X, S-3X, and S-4X were 95, 643, 627, and 1,543 nM, respectively. The primary resistance mutations, V9982A, I5950V, and V3732I, are consistent with the resistance pattern developed by HIV-1 under similar selection conditions. While resistance to LPV emerged readily, similar PR mutations causing resistance to either DRV or TL-3 failed to emerge after passage for more than a year. However, a G37D mutation in the nucleocapsid (NC) was observed in both selections and an isogenic G37D mutant replicated in the presence of 100 nM DRV or TL-3, whereas parental chimeric FIV could not. An additional mutation, L92V, near the PR active site in the folded structure recently emerged during TL-3 selection. The L92V mutant PR exhibited an IC50 of 50 nM, compared to 35 nM for 6s-98S PR, and processed the NC-p2 junction more efficiently, consistent with increased viral fitness. These findings emphasize the role of mutations outside the active site of PR in increasing viral resistance to active-site inhibitors and suggest additional targets for inhibitor development.  相似文献   

14.
Kulsam Ali  Peter Heathcote  Saul Purton 《BBA》2006,1757(12):1623-1633
A conserved tryptophan residue located between the A1B and FX redox centres on the PsaB side of the Photosystem I reaction centre has been mutated to a glycine in Chlamydomonas reinhardtii, thereby matching the conserved residue found in the equivalent position on the PsaA side. This mutant (PsaB:W669G) was studied using EPR spectroscopy with a view to understanding the molecular basis of the reported kinetic differences in forward electron transfer from the A1A and the A1B phyllo(semi)quinones. The kinetics of A1 reoxidation due to forward electron transfer or charge recombination were measured by electron spin echo spectroscopy at 265 K and 100 K, respectively. At 265 K, the reoxidation kinetics are considerably lengthened in the mutant in comparison to the wild-type. Under conditions in which FX is initially oxidised the kinetics of charge recombination at 100 K are found to be biphasic in the mutant while they are substantially monophasic in the wild-type. Pre-reduction of FX leads to biphasic kinetics in the wild-type, but does not alter the already biphasic kinetic properties of the PsaB:W669G mutant. Reduction of the [4Fe-4S] clusters FA and FB by illumination at 15 K is suppressed in the mutant. The results provide further support for the bi-directional model of electron transfer in Photosystem I of C. reinhardtii, and indicate that the replacement of the tryptophan residue with glycine mainly affects the redox properties of the PsaB bound phylloquinone A1B.  相似文献   

15.
A kinetic analysis was made of l-valine uptake in protoplast-derived cells (mesophyll protoplasts cultured for 6 days) and in suspension-cultured cells of tobacco (Nicotiana tabacum L., cv Xanthi). Cells from wild-type and Valr-2 mutant plants were compared. A low-Km component was found in protoplast-derived cells (Km = 45 ± 5 micromolar) as well as in suspension-cultured cells (Km = 84 ± 21 micromolar). In the mutant cells the Vmax of this component was 12- to 14-fold less than in wild-type cells. A second component (Km = 2.4 ± 0.7 millimolar) was found in suspension-cultured cells but not in protoplast-derived cells; its Vmax was the same in wild-type and mutant cells. A third component was apparently unsaturable (linear component). It was present in protoplast-derived cells but not in suspension-cultured cells, and had the same magnitude in wild-type and mutant cells. The results are discussed with reference to the uptake of l-valine in leaf tissue, in which the three kinetic components have been found simultaneously. The reduced Vmax of the low-Km component in the Valr-2 mutant, and the differential expression of the other two components in suspension-cultured cells and protoplast-derived cells indicate that the kinetically distinguishable components represent physically distinct transport systems.  相似文献   

16.
The melibiose permease of Salmonella enterica serovar Typhimurium (MelBSt) catalyzes symport of melibiose with Na+, Li+, or H+. Bioinformatics and mutational analyses indicate that a conserved Gly117 (helix IV) is a component of the Na+-binding site. In this study, Gly117 was mutated to Ser, Asn, or Cys. All three mutations increase the maximum rate (Vmax) for melibiose transport in Escherichia coli DW2 and greatly decrease Na+ affinity, indicating that intracellular release of Na+ is facilitated. Rapid melibiose transport, particularly by the G117N mutant, triggers osmotic lysis in the lag phase of growth. The findings support the previous conclusion that Gly117 plays an important role in cation binding and translocation. Furthermore, a spontaneous second-site mutation (P148L between loop4-5 and helix V) in the G117C mutant prevents cell lysis. This mutation significantly decreases Vmax with little effect on cosubstrate binding in G117C, G117S, and G117N mutants. Thus, the P148L mutation specifically inhibits transport velocity and thereby blocks the lethal effect of elevated melibiose transport in the Gly117 mutants.  相似文献   

17.
An Escherichia coli mutant lacking the modified nucleotide m1G in rRNA has previously been isolated (G. R. Björk and L. A. Isaksson, J. Mol. Biol. 51:83–100, 1970). In this study, we localize the position of the m1G to nucleotide 745 in 23S rRNA and characterize a mutant deficient in this modification. This mutant shows a 40% decreased growth rate in rich media, a drastic reduction in loosely coupled ribosomes, a 20% decreased polypeptide chain elongation rate, and increased resistance to the ribosome binding antibiotic viomycin. The rrmA gene encoding 23S rRNA m1G745 methyltransferase was mapped to bp 1904000 on the E. coli chromosome and identified to be identical to the previously sequenced gene yebH.  相似文献   

18.
The base-pair changes induced by the highly carcinogenic agent, 4-nitroquinoline-1-oxide, have been determined from the reversion rates of defined tester strains and from the amino acid replacements of revertant iso-1-cytochromes c. The mutant codons and the base-pair changes of reverse mutations of 14 cyc1 mutants were previously determined from alterations of iso-1-cytochromes c in intragenic revertants. These 14 cyc1 mutants, which were used as tester strains, included nine mutants with altered AUG initiation codons, an ochre (UAA) mutant, an amber (UAG) mutant and three frameshift mutants (Stewart et al., 1971,1972; Stewart &; Sherman, 1972,1974; Sherman &; Stewart, 1973). NQO2 induced a high rate of reversion in the initiation mutant cyc1-131, the only mutant in the group which reverts to normal iso-1-cytochrome c by a G · C → A · T transition. In addition, NQO produces a significant rate of reversion of all cyc1 mutants which revert by G · C transversions, e.g. the amber (UAG) mutant and the initiation mutants containing AGG, and probably CUG mutant codons. It did not revert the ochre mutant which contains no G · C base pairs. Ten NQO-induced revertants of the amber mutant cyc1-179 contained the expected replacements of residues of tyrosine, and ten NQO-induced revertants of each of the cyc1-131 and cyc1-133 initiation mutants all contained the expected normal iso-1-cytochrome c. The structures of these iso-1-cytochromes c and the pattern of reversion of the tester strains indicate that base-pair substitutions arise at G · C base pairs which are the site of NQO attack. Thus NQO induces G · C → A · T transitions, G · C → T · A transversions and possibly G · C → C · G transversions. Because of its mode of action, NQO may be useful in less-defined systems for identifying G · C base pairs in mutant codons.  相似文献   

19.
A clinically‐relevant, drug‐resistant mutant of HIV‐1 protease (PR), termed Flap+(I54V) and containing L10I, G48V, I54V and V82A mutations, is known to produce significant changes in the entropy and enthalpy balance of drug‐PR interactions, compared to wild‐type PR. A similar mutant, Flap+(I54A), which evolves from Flap+(I54V) and contains the single change at residue 54 relative to Flap+(I54V), does not. Yet, how Flap+(I54A) behaves in solution is not known. To understand the molecular basis of V54A evolution, we compared nuclear magnetic resonance (NMR) spectroscopy, fluorescence spectroscopy, isothermal titration calorimetry, and enzymatic assay data from four PR proteins: PR (pWT), Flap+(I54V), Flap+(I54A), and Flap+(I54), a control mutant that contains only L10I, G48V and V82A mutations. Our data consistently show that selection to the smaller side chain at residue 54, not only decreases inhibitor affinity, but also restores the catalytic activity.  相似文献   

20.
Most archaea and bacteria use a modified C in the anticodon wobble position of isoleucine tRNA to base pair with A but not with G of the mRNA. This allows the tRNA to read the isoleucine codon AUA without also reading the methionine codon AUG. To understand why a modified C, and not U or modified U, is used to base pair with A, we mutated the C34 in the anticodon of Haloarcula marismortui isoleucine tRNA (tRNA2Ile) to U, expressed the mutant tRNA in Haloferax volcanii, and purified and analyzed the tRNA. Ribosome binding experiments show that although the wild-type tRNA2Ile binds exclusively to the isoleucine codon AUA, the mutant tRNA binds not only to AUA but also to AUU, another isoleucine codon, and to AUG, a methionine codon. The G34 to U mutant in the anticodon of another H. marismortui isoleucine tRNA species showed similar codon binding properties. Binding of the mutant tRNA to AUG could lead to misreading of the AUG codon and insertion of isoleucine in place of methionine. This result would explain why most archaea and bacteria do not normally use U or a modified U in the anticodon wobble position of isoleucine tRNA for reading the codon AUA. Biochemical and mass spectrometric analyses of the mutant tRNAs have led to the discovery of a new modified nucleoside, 5-cyanomethyl U in the anticodon wobble position of the mutant tRNAs. 5-Cyanomethyl U is present in total tRNAs from euryarchaea but not in crenarchaea, eubacteria, or eukaryotes.  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