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.
The M1 strain, able to grow on beta-myrcene as the sole carbon and energy source, was isolated by an enrichment culture and identified as a Pseudomonas sp. One beta-myrcene-negative mutant, called N22, obtained by transposon mutagenesis, accumulated (E)-2-methyl-6-methylen-2,7-octadien-1-ol (or myrcen-8-ol) as a unique beta-myrcene biotransformation product. This compound was identified by gas chromatography-mass spectrometry. We cloned and sequenced the DNA regions flanking the transposon and used these fragments to identify the M1 genomic library clones containing the wild-type copy of the interrupted gene. One of the selected cosmids, containing a 22-kb genomic insert, was able to complement the N22 mutant for growth on beta-myrcene. A 5,370-bp-long sequence spanning the region interrupted by the transposon in the mutant was determined. We identified four open reading frames, named myrA, myrB, myrC, and myrD, which can potentially code for an aldehyde dehydrogenase, an alcohol dehydrogenase, an acyl-coenzyme A (CoA) synthetase, and an enoyl-CoA hydratase, respectively. myrA, myrB, and myrC are likely organized in an operon, since they are separated by only 19 and 36 nucleotides (nt), respectively, and no promoter-like sequences have been found in these regions. The myrD gene starts 224 nt upstream of myrA and is divergently transcribed. The myrB sequence was found to be completely identical to the one flanking the transposon in the mutant. Therefore, we could ascertain that the transposon had been inserted inside the myrB gene, in complete agreement with the accumulation of (E)-2-methyl-6-methylen-2,7-octadien-1-ol by the mutant. Based on sequence and biotransformation data, we propose a pathway for beta-myrcene catabolism in Pseudomonas sp. strain M1.  相似文献   

2.
The biotransformation of lithocholic acid by Pseudomonas sp. strain NCIB 10590 under anaerobic conditions was studied. The major products were identified as androsta-1,4-diene-3,17-dione and 3-oxochol-4-ene-24-oic acid. The minor products included 17β-hydroxyandrost-4-ene-3-one, 17β-hydroxyandrosta-1,4-diene-3-one, 3-oxo-5β-cholan-24-oic acid, 3-oxochola-1,4-diene-24-oic acid, 3-oxopregn-4-ene-20-carboxylic acid, and 3-oxopregna-1,4-diene-20-carboxylic acid. Anaerobiosis increases the number of metabolites produced by Pseudomonas sp. NCIB 10590 from lithocholic acid.  相似文献   

3.
When Chlorella sorokiniana was cultured in the presence of 1 mg/1 triparanol succinate, there was a 42% reduction in total sterol concentration. Algal biomass was reduced by approximately the same amount. In addition to the cycloartenol, cyclolaudenol, 24-methyl-pollinastanol, ergosta-5, 7-dien-3β-ol, and ergosterol that occur in control culture, pollinastanol, 14α-methyl-5α-ergost-8-en-3β-ol, 5α-ergosta-8, 14, 22-trien-3β-ol, 5α-ergosta-8(14), 22-dien-3β-ol, 5α-ergosta-8(9), 22-dien-3β-ol, 5α-ergosta-8, 14-dien-3β-ol, 5α-ergost-8(9)-3n-3β-ol, 5α-ergost-8(14)-en-3β-ol, 5α-ergosta-7, 22-dien-3β-ol, and 5α-ergost-7-en-3β-ol were isolated and identified from triparanol succinate-treated cells. A biosynthetic pathway for sterol biosynthesis in this organism is postulated based on all the sterols that were isolated and identified in triparanol-treated cultures of C. sorokiniana. Cyclolaudenol appears to be the product of the first alkylation at C-24 in this organism rather than the more common 24-methylene cycloartanol. Since 24-methylene sterols are needed for the second alkylation reaction, this would explain the absence of C-29 sterols in C. sorokiniana. Four of the sterols identified in C. sorokiniana are reported for the first time in a living organism. They are: 24-methyl pollinastanol, 5α-ergosta-8, 14, 22-trien-3β-ol, 5α-ergosta-8(14), 22-dien-3β-ol and 5α-ergost-8(14)-en-3β-ol.  相似文献   

4.
Monoterpenes with an unsaturated hydrocarbon structure are mineralized anaerobically by the denitrifying β-proteobacterium Alcaligenes defragrans. Organic acids occurring in cells of A. defragrans and culture medium were characterized to identify potential products of the monoterpene activation reaction. Geranic acid (E,E-3,7-dimethyl-2,6-octadienoic acid) accumulated to 0.5 mM in cells grown on α-phellandrene under nitrate limitation. Cell suspensions of A. defragrans 65Phen synthesized geranic acid in the presence of β-myrcene, α-phellandrene, limonene, or α-pinene. Myrcene yielded the highest transformation rates. The alicyclic acid was consumed by cell suspensions during carbon limitation. Heat-labile substances present in cytosolic extracts catalyzed the formation of geranic acid from myrcene. These results indicated that a novel monoterpene degradation pathway must be present in A. defragrans.  相似文献   

5.
The gene loci fcs, encoding feruloyl coenzyme A (feruloyl-CoA) synthetase, ech, encoding enoyl-CoA hydratase/aldolase, and aat, encoding β-ketothiolase, which are involved in the catabolism of ferulic acid and eugenol in Pseudomonas sp. strain HR199 (DSM7063), were localized on a DNA region covered by two EcoRI fragments (E230 and E94), which were recently cloned from a Pseudomonas sp. strain HR199 genomic library in the cosmid pVK100. The nucleotide sequences of parts of fragments E230 and E94 were determined, revealing the arrangement of the aforementioned genes. To confirm the function of the structural genes fcs and ech, they were cloned and expressed in Escherichia coli. Recombinant strains harboring both genes were able to transform ferulic acid to vanillin. The feruloyl-CoA synthetase and enoyl-CoA hydratase/aldolase activities of the fcs and ech gene products, respectively, were confirmed by photometric assays and by high-pressure liquid chromatography analysis. To prove the essential involvement of the fcs, ech, and aat genes in the catabolism of ferulic acid and eugenol in Pseudomonas sp. strain HR199, these genes were inactivated separately by the insertion of omega elements. The corresponding mutants Pseudomonas sp. strain HRfcsΩGm and Pseudomonas sp. strain HRechΩKm were not able to grow on ferulic acid or on eugenol, whereas the mutant Pseudomonas sp. strain HRaatΩKm exhibited a ferulic acid- and eugenol-positive phenotype like the wild type. In conclusion, the degradation pathway of eugenol via ferulic acid and the necessity of the activation of ferulic acid to the corresponding CoA ester was confirmed. The aat gene product was shown not to be involved in this catabolism, thus excluding a β-oxidation analogous degradation pathway for ferulic acid. Moreover, the function of the ech gene product as an enoyl-CoA hydratase/aldolase suggests that ferulic acid degradation in Pseudomonas sp. strain HR199 proceeds via a similar pathway to that recently described for Pseudomonas fluorescens AN103.  相似文献   

6.
The Kluyveromyces lactis heterotrimeric G protein is a canonical Gαβγ complex; however, in contrast to Saccharomyces cerevisiae, where the Gγ subunit is essential for mating, disruption of the KlGγ gene yielded cells with almost intact mating capacity. Expression of a nonfarnesylated Gγ, which behaves as a dominant-negative in S. cerevisiae, did not affect mating in wild-type and ΔGγ cells of K. lactis. In contrast to the moderate sterility shown by the single ΔKlGα, the double ΔKlGα ΔKlGγ mutant displayed full sterility. A partial sterile phenotype of the ΔKlGγ mutant was obtained in conditions where the KlGβ subunit interacted defectively with the Gα subunit. The addition of a CCAAX motif to the C-end of KlGβ, partially suppressed the lack of both KlGα and KlGγ subunits. In cells lacking KlGγ, the KlGβ subunit cofractionated with KlGα in the plasma membrane, but in the ΔKlGα ΔKlGγ strain was located in the cytosol. When the KlGβ-KlGα interaction was affected in the ΔKlGγ mutant, most KlGβ fractionated to the cytosol. In contrast to the generic model of G-protein function, the Gβ subunit of K. lactis has the capacity to attach to the membrane and to activate mating effectors in absence of the Gγ subunit.  相似文献   

7.
In the bacterial degradation of steroid compounds, the enzymes initiating the breakdown of the steroid rings are well known, while the reactions for degrading steroid side chains attached to C-17 are largely unknown. A recent in vitro analysis with Pseudomonas sp. strain Chol1 has shown that the degradation of the C5 acyl side chain of the C24 steroid compound cholate involves the C22 intermediate 7α,12α-dihydroxy-3-oxopregna-1,4-diene-20S-carbaldehyde (DHOPDCA) with a terminal aldehyde group. In the present study, candidate genes with plausible functions in the formation and degradation of this aldehyde were identified. All deletion mutants were defective in growth with cholate but could transform it into dead-end metabolites. A mutant with a deletion of the shy gene, encoding a putative enoyl coenzyme A (CoA) hydratase, accumulated the C24 steroid (22E)-7α,12α-dihydroxy-3-oxochola-1,4,22-triene-24-oate (DHOCTO). Deletion of the sal gene, formerly annotated as the steroid ketothiolase gene skt, resulted in the accumulation of 7α,12α,22-trihydroxy-3-oxochola-1,4-diene-24-oate (THOCDO). In cell extracts of strain Chol1, THOCDO was converted into DHOPDCA in a coenzyme A- and ATP-dependent reaction. A sad deletion mutant accumulated DHOPDCA, and expression in Escherichia coli revealed that sad encodes an aldehyde dehydrogenase for oxidizing DHOPDCA to the corresponding acid 7α,12α-dihydroxy-3-oxopregna-1,4-diene-20-carboxylate (DHOPDC) with NAD+ as the electron acceptor. These results clearly show that the degradation of the acyl side chain of cholate proceeds via an aldolytic cleavage of an acetyl residue; they exclude a thiolytic cleavage for this reaction step. Based on these results and on sequence alignments with predicted aldolases from other bacteria, we conclude that the enzyme encoded by sal catalyzes this aldolytic cleavage.  相似文献   

8.
CWH41, a gene involved in the assembly of cell wall β-1,6-glucan, has recently been shown to be the structural gene for Saccharomyces cerevisiae glucosidase I that is responsible for initiating the trimming of terminal α-1,2-glucose residue in the N-glycan processing pathway. To distinguish between a direct or indirect role of Cwh41p in the biosynthesis of β-1,6-glucan, we constructed a double mutant, alg5Δ (lacking dolichol-P-glucose synthase) cwh41Δ, and found that it has the same phenotype as the alg5Δ single mutant. It contains wild-type levels of cell wall β-1,6-glucan, shows moderate underglycosylation of N-linked glycoproteins, and grows at concentrations of Calcofluor White (which interferes with cell wall assembly) that are lethal to cwh41Δ single mutant. The strong genetic interactions of CWH41 with KRE6 and KRE1, two other genes involved in the β-1,6-glucan biosynthetic pathway, disappear in the absence of dolichol-P-glucose synthase (alg5Δ). The triple mutant alg5Δcwh41Δkre6Δ is viable, whereas the double mutant cwh41Δkre6Δ in the same genetic background is not. The severe slow growth phenotype and 75% reduction in cell wall β-1,6-glucan, characteristic of the cwh41Δkre1Δ double mutant, are not observed in the triple mutant alg5Δcwh41Δkre1Δ. Kre6p, a putative Golgi glucan synthase, is unstable in cwh41Δ strains, and its overexpression renders these cells Calcofluor White resistant. These results demonstrate that the role of glucosidase I (Cwh41p) in the biosynthesis of cell wall β-1,6-glucan is indirect and that dolichol-P-glucose is not an intermediate in this pathway.  相似文献   

9.
Several aspects of fescue foot in cattle suggest that this disease is caused by fungi growing on fescue grass. Certain fungi isolated from winter pasture yield toxins when grown on synthetic medium. Most of these toxin producers belong to the genus Fusarium. All but 1 of the 21 toxic and 7 questionably toxic Fusarium isolates produce either 4-acetamido-4-hydroxy-2-butenoic acid γ-lactone, or 4β, 15-diacetoxy-8α-(3-methylbutyryloxy)-12, 13-epoxytrichothec-9-en-3α-ol, or both.  相似文献   

10.
Selected monoterpenes inhibited methane oxidation by methanotrophs (Methylosinus trichosporium OB3b, Methylobacter luteus), denitrification by environmental isolates, and aerobic metabolism by several heterotrophic pure cultures. Inhibition occurred to various extents and was transient. Complete inhibition of methane oxidation by Methylosinus trichosporium OB3b with 1.1 mM (−)-α-pinene lasted for more than 2 days with a culture of optical density of 0.05 before activity resumed. Inhibition was greater under conditions under which particulate methane monooxygenase was expressed. No apparent consumption or conversion of monoterpenes by methanotrophs was detected by gas chromatography, and the reason that transient inhibition occurs is not clear. Aerobic metabolism by several heterotrophs was much less sensitive than methanotrophy was; Escherichia coli (optical density, 0.01), for example, was not affected by up to 7.3 mM (−)-α-pinene. The degree of inhibition was monoterpene and species dependent. Denitrification by isolates from a polluted sediment was not inhibited by 3.7 mM (−)-α-pinene, γ-terpinene, or β-myrcene, whereas 50 to 100% inhibition was observed for isolates from a temperate swamp soil. The inhibitory effect of monoterpenes on methane oxidation was greatest with unsaturated, cyclic hydrocarbon forms [e.g., (−)-α-pinene, (S)-(−)-limonene, (R)-(+)-limonene, and γ-terpinene]. Lower levels of inhibition occurred with oxide and alcohol derivatives [(R)-(+)-limonene oxide, α-pinene oxide, linalool, α-terpineol] and a noncyclic hydrocarbon (β-myrcene). Isomers of pinene inhibited activity to different extents. Given their natural sources, monoterpenes may be significant factors affecting bacterial activities in nature.  相似文献   

11.
12.
In the present study, glutaryl-7-amino cephalosporanic acid acylase from Pseudomonas sp. strain 130 (CA130) was mutated to improve its enzymatic activity and stability. Based on the crystal structure of CA130, two series of amino acid residues, one from those directly involved in catalytic function and another from those putatively involved in surface charge, were selected as targets for site-directed mutagenesis. In the first series of experiments, several key residues in the substrate-binding pocket were substituted, and the genes were expressed in Escherichia coli for activity screening. Two of the mutants constructed, Y151αF and Q50βN, showed two- to threefold-increased catalytic efficiency (kcat/Km) compared to wild-type CA130. Their Km values were decreased by ca. 50%, and the kcat values increased to 14.4 and 16.9 s−1, respectively. The ability of these mutants to hydrolyze adipoyl 6-amino penicillinic acid was also improved. In the second series of mutagenesis, several mutants with enhanced stabilities were identified. Among them, R121βA and K198βA had a 30 to 58% longer half-life than wild-type CA130, and K198βA and D286βA showed an alkaline shift of optimal pH by about 1.0 to 2.0 pH units. To construct an engineered enzyme with the properties of both increased activity and stability, the double mutant Q50βN/K198βA was expressed. This enzyme was purified and immobilized for catalytic analysis. The immobilized mutant enzyme showed a 34.2% increase in specific activity compared to the immobilized wild-type CA130.  相似文献   

13.
Several factors regulate plant organ growth polarity. tortifolia2 (tor2), a right-handed helical growth mutant, has a conservative replacement of Arg-2 with Lys in the α-tubulin 4 protein. Based on a published high-resolution (2.89 Å) tubulin structure, we predict that Arg-2 of α-tubulin forms hydrogen bonds with the GTPase domain of β-tubulin, and structural modeling suggests that these contacts are interrupted in tor2. Consistent with this, we found that microtubule dynamicity is reduced in the tor2 background. We investigated the developmental origin of the helical growth phenotype using tor2. One hypothesis predicts that cell division patterns cause helical organ growth in Arabidopsis thaliana mutants. However, cell division patterns of tor2 root tips appear normal. Experimental uncoupling of cell division and expansion suggests that helical organ growth is based on cell elongation defects only. Another hypothesis is that twisting is due to inequalities in expansion of epidermal and cortical tissues. However, freely growing leaf trichomes of tor2 mutants show right-handed twisting and cortical microtubules form left-handed helices as early as the unbranched stage of trichome development. Trichome twisting is inverted in double mutants with tor3, a left-handed mutant. Single tor2 suspension cells also exhibit handed twisting. Thus, twisting of tor2 mutant organs appears to be a higher-order expression of the helical expansion of individual cells.  相似文献   

14.
1. Bile salts of Petromyzon marinus L. ammocoetes appeared to consist solely or chiefly of a crystalline substance, whose chromatographic and i.r.-spectral characteristics suggested that it was a monosulphate ester of a bile alcohol having the 3α,7α,12α-trihydroxy pattern of substitution in a 5α-steroid nucleus. 2. This substance on cleavage with dioxan–trichloroacetic acid gave petromyzonol, n.m.r. and mass-spectral examination of which suggested the structure 5α-cholane-3α,7α,12α,24-tetrol. 3. 3α,7α,12α-Trihydroxy-5α-cholanoic acid (allocholic acid) from the lizards Anolis lineatopus lineatopus Gray and Cyclura carinata Harlan (family Iguanidae) was esterified with propan-1-ol and reduced by lithium aluminium hydride to 5α-cholane-3α,7α,12α,24-tetrol, identical with petromyzonol. 4. Chromic acid oxidation of petromyzonol sulphate from lamprey bile, followed by acid hydrolysis, gave 24-hydroxy-5α-cholane-3,7,12-trione; hence the sulphate ester group is at C-24. 5. Petromyzonol sulphate is both primitive and unique: a study of its biogenesis might improve our understanding of evolution at the molecular level.  相似文献   

15.
The CelA β-glucosidase of Azospirillum irakense, belonging to glycosyl hydrolase family 3 (GHF3), preferentially hydrolyzes cellobiose and releases glucose units from the C3, C4, and C5 oligosaccharides. The growth of a ΔcelA mutant on these cellobiosides was affected. In A. irakense, the GHF3 β-glucosidases appear to be functional alternatives for the GHF1 β-glucosidases in the assimilation of β-glucosides by other bacteria.  相似文献   

16.
Serum antibodies and mannose-binding lectin (MBL) are important host defense factors for host adaptive and innate immunity, respectively. Antibodies and MBL also initiate the classical and lectin complement pathways, respectively, leading to opsonophagocytosis. We have shown previously that Staphylococcus aureus wall teichoic acid (WTA), a cell wall glycopolymer consisting of ribitol phosphate substituted with α- or β-O-N-acetyl-d-glucosamine (GlcNAc) and d-alanine, is recognized by MBL and serum anti-WTA IgG. However, the exact antigenic determinants to which anti-WTA antibodies or MBL bind have not been determined. To answer this question, several S. aureus mutants, such as α-GlcNAc glycosyltransferase-deficient S. aureus ΔtarM, β-GlcNAc glycosyltransferase-deficient ΔtarS, and ΔtarMS double mutant cells, were prepared from a laboratory and a community-associated methicillin-resistant S. aureus strain. Here, we describe the unexpected finding that β-GlcNAc WTA-deficient ΔtarS mutant cells (which have intact α-GlcNAc) escape from anti-WTA antibody-mediated opsonophagocytosis, whereas α-GlcNAc WTA-deficient ΔtarM mutant cells (which have intact β-GlcNAc) are efficiently engulfed by human leukocytes via anti-WTA IgG. Likewise, MBL binding in S. aureus cells was lost in the ΔtarMS double mutant but not in either single mutant. When we determined the serum concentrations of the anti-α- or anti-β-GlcNAc-specific WTA IgGs, anti-β-GlcNAc WTA-IgG was dominant in pooled human IgG fractions and in the intact sera of healthy adults and infants. These data demonstrate the importance of the WTA sugar conformation for human innate and adaptive immunity against S. aureus infection.  相似文献   

17.
The epithelial Na+ channel (ENaC) is comprised of three homologous subunits (α, β, and γ) that have a similar topology with two transmembrane domains, a large extracellular region, and cytoplasmic N and C termini. Although ENaC activity is regulated by a number of factors, palmitoylation of its cytoplasmic Cys residues has not been previously described. Fatty acid-exchange chemistry was used to determine whether channel subunits were Cys-palmitoylated. We observed that only the β and γ subunits were modified by Cys palmitoylation. Analyses of ENaCs with mutant β subunits revealed that Cys-43 and Cys-557 were palmitoylated. Xenopus oocytes expressing ENaC with a β C43A,C557A mutant had significantly reduced amiloride-sensitive whole cell currents, enhanced Na+ self-inhibition, and reduced single channel Po when compared with wild-type ENaC, while membrane trafficking and levels of surface expression were unchanged. Computer modeling of cytoplasmic domains indicated that β Cys-43 is in proximity to the first transmembrane α helix, whereas β Cys-557 is within an amphipathic α-helix contiguous with the second transmembrane domain. We propose that β subunit palmitoylation modulates channel gating by facilitating interactions between cytoplasmic domains and the plasma membrane.  相似文献   

18.
In Saccharomyces cerevisiae, meiosis and spore formation as well as mating are controlled by mating-type genes. Diploids heterozygous for mating type (aα) can sporulate but cannot mate; homozygous aa and αα diploids can mate, but cannot sporulate. From an αα diploid parental strain, we have isolated mutants which have gained the ability to sporulate. Those mutants which continue to mate as αα cells have been designated CSP (control of sporulation). Upon sporulation, CSP mutants yield asci containing 4α spores. The mutant gene which allows αα cells to sporulate is unlinked to the mating-type locus and also acts to permit sporulation in aa diploid cells. Segregation data from crosses between mutant αα and wild-type aa diploids and vice versa indicate (for all but one mutant) that the mutation which allows constitutive sporulation (CSP) is dominant over the wild-type allele. Some of the CSP mutants are temperature-sensitive, sporulating at 32°, but not at 23°. In addition to CSP mutants, our mutagenesis and screening procedure led to the isolation of mutants which sporulate by virtue of a change in the mating-type locus itself, resulting in loss of ability to mate.  相似文献   

19.
Acinetobacter sp. strain YAA has five genes (atdA1 to atdA5) involved in aniline oxidation as a part of the aniline degradation gene cluster. From sequence analysis, the five genes were expected to encode a glutamine synthetase (GS)-like protein (AtdA1), a glutamine amidotransferase-like protein (AtdA2), and an aromatic compound dioxygenase (AtdA3, AtdA4, and AtdA5) (M. Takeo, T. Fujii, and Y. Maeda, J. Ferment. Bioeng. 85:17-24, 1998). A recombinant Pseudomonas strain harboring these five genes quantitatively converted aniline into catechol, demonstrating that catechol is the major oxidation product from aniline. To elucidate the function of the GS-like protein AtdA1 in aniline oxidation, we purified it from recombinant Escherichia coli harboring atdA1. The purified AtdA1 protein produced gamma-glutamylanilide (γ-GA) quantitatively from aniline and l-glutamate in the presence of ATP and MgCl2. This reaction was identical to glutamine synthesis by GS, except for the use of aniline instead of ammonia as the substrate. Recombinant Pseudomonas strains harboring the dioxygenase genes (atdA3 to atdA5) were unable to degrade aniline but converted γ-GA into catechol, indicating that γ-GA is an intermediate to catechol and a direct substrate for the dioxygenase. Unexpectedly, a recombinant Pseudomonas strain harboring only atdA2 hydrolyzed γ-GA into aniline, reversing the γ-GA formation by AtdA1. Deletion of atdA2 from atdA1 to atdA5 caused γ-GA accumulation from aniline in recombinant Pseudomonas cells and inhibited the growth of a recombinant Acinetobacter strain on aniline, suggesting that AtdA2 prevents γ-GA accumulation that is harmful to the host cell.  相似文献   

20.
Methionine residues in α/β-type small, acid-soluble spore proteins (SASP) of Bacillus species were readily oxidized to methionine sulfoxide in vitro by t-butyl hydroperoxide (tBHP) or hydrogen peroxide (H2O2). These oxidized α/β-type SASP no longer bound to DNA effectively, but DNA binding protected α/β-type SASP against methionine oxidation by peroxides in vitro. Incubation of an oxidized α/β-type SASP with peptidyl methionine sulfoxide reductase (MsrA), which can reduce methionine sulfoxide residues back to methionine, restored the α/β-type SASP’s ability to bind to DNA. Both tBHP and H2O2 caused some oxidation of the two methionine residues of an α/β-type SASP (SspC) in spores of Bacillus subtilis, although one methionine which is highly conserved in α/β-type SASP was only oxidized to a small degree. However, much more methionine sulfoxide was generated by peroxide treatment of spores carrying a mutant form of SspC which has a lower affinity for DNA. MsrA activity was present in wild-type B. subtilis spores. However, msrA mutant spores were no more sensitive to H2O2 than were wild-type spores. The major mechanism operating for dealing with oxidative damage to α/β-type SASP in spores is DNA binding, which protects the protein’s methionine residues from oxidation both in vitro and in vivo. This may be important in vivo since α/β-type SASP containing oxidized methionine residues no longer bind DNA well and α/β-type SASP-DNA binding is essential for long-term spore survival.  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