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 42 毫秒
1.
Homoisocitrate dehydrogenase is involved in the alpha-aminoadipate pathway of biosynthesis of l-lysine in fungi, yeast, some prokaryotic bacteria, and archaea. This enzyme catalyzes the oxidative decarboxylation of (2R,3S)-homoisocitrate into 2-oxoadipate using NAD(+) as a coenzyme. Substrate specificity of two homoisocitrate dehydrogenases derived from Deinococcus radiodurans and Saccharomyces cerevisiae was analyzed using a series of synthetic substrate analogs, which indicated a relatively broad substrate specificity of these enzymes. Based on the substrate specificity, 3-hydroxyalkylidene- and 3-carboxyalkylidenemalate derivatives were designed as a specific inhibitor for homoisocitrate dehydrogenase. The synthetic inhibitors showed a moderate competitive inhibitory activity and (R,Z)-3-carboxypropylidenemalate was the most inhibitory among the synthesized inhibitors. Therefore, homoisocitrate dehydrogenase appeared to recognize preferentially an extended conformation of homoisocitrate.  相似文献   

2.
Although the presence of an enzyme that catalyzes beta-decarboxylating dehydrogenation of homoisocitrate to synthesize 2-oxoadipate has been postulated in the lysine biosynthesis pathway through alpha-aminoadipate (AAA), the enzyme has not yet been analyzed at all, because no gene encoding the enzyme has been identified until recently. A gene encoding a protein with a significant amino acid sequence identity to both isocitrate dehydrogenase and 3-isopropylmalate dehydrogenase was cloned from Thermus thermophilus HB27. The gene product produced in recombinant Escherichia coli cells demonstrated homoisocitrate dehydrogenase (HICDH) activity. A knockout mutant of the gene showed an AAA-auxotrophic phenotype, indicating that the gene product is involved in lysine biosynthesis through AAA. We therefore named this gene hicdh. HICDH, the gene product, did not catalyze the conversion of 3-isopropylmalate to 2-oxoisocaproate, a leucine biosynthetic reaction, but it did recognize isocitrate, a related compound in the tricarboxylic acid cycle, as well as homoisocitrate as a substrate. It is of interest that HICDH catalyzes the reaction with isocitrate about 20 times more efficiently than the reaction with the putative native substrate, homoisocitrate. The broad specificity and possible dual function suggest that this enzyme represents a key link in the evolution of the pathways utilizing citrate derivatives. Site-directed mutagenesis study reveals that replacement of Arg(85) with Val in HICDH causes complete loss of activity with isocitrate but significant activity with 3-isopropylmalate and retains activity with homoisocitrate. These results indicate that Arg(85) is a key residue for both substrate specificity and evolution of beta-decarboxylating dehydrogenases.  相似文献   

3.
The alpha-aminoadipate pathway for the biosynthesis of lysine is present only in fungi and euglena. Until now, this unique metabolic pathway has never been investigated in the opportunistic fungal pathogens Candida albicans, Cryptococcus neoformans, and Aspergillus fumigatus. Five of the eight enzymes (homocitrate synthase, homoisocitrate dehydrogenase, alpha-aminoadipate reductase, saccharopine reductase, and saccharopine dehydrogenase) of the alpha-aminoadipate pathway and glucose-6-phosphate dehydrogenase, a glycolytic enzyme used as a control, were demonstrated in wild-type cells of these organisms. All enzymes were present in Saccharomyces cerevisiae and the pathogenic organisms except C. neoformans 32608 serotype C, which exhibited no saccharopine reductase activity. The levels of enzyme activity varied considerably from strain to strain. Variation among organisms was also observed for the control enzyme. Among the pathogens, C. albicans exhibited much higher homocitrate synthase, homoisocitrate dehydrogenase, and alpha-aminoadipate reductase activities. Seven lysine auxotrophs of C. albicans and one of Candida tropicalis were characterized biochemically to determine the biochemical blocks and gene-enzyme relationships. Growth responses to alpha-aminoadipate- and lysine-supplemented media, accumulation of alpha-aminoadipate semialdehyde, and the lack of enzyme activity revealed that five of the mutants (WA104, WA153, WC7-1-3, WD1-31-2, and A5155) were blocked at the alpha-aminoadipate reductase step, two (STN57 and WD1-3-6) were blocked at the saccharopine dehydrogenase step, and the C. tropicalis mutant (X-16) was blocked at the saccharopine reductase step. The cloned LYS1 gene of C. albicans in the recombinant plasmid YpB1078 complemented saccharopine dehydrogenase (lys1) mutants of S. cerevisiae and C. albicans. The Lys1+ transformed strains exhibited significant saccharopine dehydrogenase activity in comparison with untransformed mutants. The cloned LYS1 gene has been localized on a 1.8-kb HindIII DNA insert of the recombinant plasmid YpB1041RG1. These results established the gene-enzyme relationship in the second half of the alpha-aminoadipate pathway. The presence of this unique pathway in the pathogenic fungi could be useful for their rapid detection and control.  相似文献   

4.
G W Plaut  R L Beach  T Aogaichi 《Biochemistry》1975,14(12):2581-2588
D-Garcinia acid (D-threo-1,2-dihydroxy-1,2,3-propanetricarboxylate), like D-isocitrate, has an alpha-DS-hydroxyl group and a beta-LS configuration of the second carboxyl group. The maximal velocity of pyridine nucleotide reduction with D-garcinia acid is 8 and 21% of D-threo-isocitrate with the DPN-linked and TPN-linked isocitrate dehydrogenase from bovine heart, respectively. The other stereoisomers of hydroxycitrate [L-garcinia acid, D- and L-hibiscus acid (D- and L-erythro-1,2-dihydroxy-1,2,3-propanetricarboxylate)] are inactive. DL-threo-Homoisocitrate (DL-threo-1-hydroxy-1,2,4-butanetricarboxylate) supports DPN+ reduction at 10-15% of the rate observed for isocitrate with the DPN-specific enzyme, but is not a substrate for TPN-linked isocitrate dehydrogenase. The values of apparent S0.5 for total isocitrate and total garcinia acid are similar with both enzymes; the apparent S0.5 of total homoisocitrate is two- to threefold higher than that of total isocitrate with the DPN-linked enzyme. Enzymatic oxidative decarboxylation of garcinia acid and homoisocitrate leads to formation of alpha-keto-beta-hydroxyglutarate and alpha-ketoadipate, respectively. DL-Methylmalate (DL-1-hydroxy-2-methylsuccinate) is inactive as a substrate for either dehydrogenase as are the newly synthesized compounds: DL-threo-gamma-isocitrate amide (DL-threo-1-hydroxy-3-carbamy01,2-propanedicarboxylate), beta-methyl-DL-isocitrate (DL-1-hydroxy-2-methyl-1,2,3-propanetricarboxylate), beta-methyl-DL-garcinia acid (DL-threo-1-hydroxyl-2-methoxy-1,2,3-propanetricarboxylate), DL-1-hydroxyl-1,2,2-ethanetricarboxylate, and DL-1,4-dihydroxy-1,2-butanedicarboxylate.  相似文献   

5.
Homoaconitase enzymes catalyze hydrolyase reactions in the alpha-aminoadipate pathway for lysine biosynthesis or the 2-oxosuberate pathway for methanogenic coenzyme B biosynthesis. Despite the homology of this iron-sulfur protein to aconitase, previously studied homoaconitases catalyze only the hydration of cis-homoaconitate to form homoisocitrate rather than the complete isomerization of homocitrate to homoisocitrate. The MJ1003 and MJ1271 proteins from the methanogen Methanocaldococcus jannaschii formed the first homoaconitase shown to catalyze both the dehydration of (R)-homocitrate to form cis-homoaconitate, and its hydration is shown to produce homoisocitrate. This heterotetrameric enzyme also used the analogous longer chain substrates cis-(homo)(2)aconitate, cis-(homo)(3)aconitate, and cis-(homo)(4)aconitate, all with similar specificities. A combination of the homoaconitase with the M. jannaschii homoisocitrate dehydrogenase catalyzed all of the isomerization and oxidative decarboxylation reactions required to form 2-oxoadipate, 2-oxopimelate, and 2-oxosuberate, completing three iterations of the 2-oxoacid elongation pathway. Methanogenic archaeal homoaconitases and fungal homoaconitases evolved in parallel in the aconitase superfamily. The archaeal homoaconitases share a common ancestor with isopropylmalate isomerases, and both enzymes catalyzed the hydration of the minimal substrate maleate to form d-malate. The variation in substrate specificity among these enzymes correlated with the amino acid sequences of a flexible loop in the small subunits.  相似文献   

6.
Homoisocitrate dehydrogenase (HIcDH, 3-carboxy-2-hydroxyadipate dehydrogenase) catalyzes the fourth reaction of the alpha-aminoadipate pathway for lysine biosynthesis, the conversion of homoisocitrate to alpha-ketoadipate using NAD as an oxidizing agent. A chemical mechanism for HIcDH is proposed on the basis of the pH dependence of kinetic parameters, dissociation constants for competitive inhibitors, and isotope effects. According to the pH-rate profiles, two enzyme groups act as acid-base catalysts in the reaction. A group with a p K a of approximately 6.5-7 acts as a general base accepting a proton as the beta-hydroxy acid is oxidized to the beta-keto acid, and this residue participates in all three of the chemical steps, acting to shuttle a proton between the C2 hydroxyl and itself. The second group acts as a general acid with a p K a of 9.5 and likely catalyzes the tautomerization step by donating a proton to the enol to give the final product. The general acid is observed in only the V pH-rate profile with homoisocitrate as a substrate, but not with isocitrate as a substrate, because the oxidative decarboxylation portion of the isocitrate reaction is limiting overall. With isocitrate as the substrate, the observed primary deuterium and (13)C isotope effects indicate that hydride transfer and decarboxylation steps contribute to rate limitation, and that the decarboxylation step is the more rate-limiting of the two. The multiple-substrate deuterium/ (13)C isotope effects suggest a stepwise mechanism with hydride transfer preceding decarboxylation. With homoisocitrate as the substrate, no primary deuterium isotope effect was observed, and a small (13)C kinetic isotope effect (1.0057) indicates that the decarboxylation step contributes only slightly to rate limitation. Thus, the chemical steps do not contribute significantly to rate limitation with the native substrate. On the basis of data from solvent deuterium kinetic isotope effects, viscosity effects, and multiple-solvent deuterium/ (13)C kinetic isotope effects, the proton transfer step(s) is slow and likely reflects a conformational change prior to catalysis.  相似文献   

7.
Fungi produce α‐aminoadipate, a precursor for penicillin and lysine via the α‐aminoadipate pathway. Despite the biotechnological importance of this pathway, the essential isomerization of homocitrate via homoaconitate to homoisocitrate has hardly been studied. Therefore, we analysed the role of homoaconitases and aconitases in this isomerization. Although we confirmed an essential contribution of homoaconitases from Saccharomyces cerevisiae and Aspergillus fumigatus, these enzymes only catalysed the interconversion between homoaconitate and homoisocitrate. In contrast, aconitases from fungi and the thermophilic bacterium Thermus thermophilus converted homocitrate to homoaconitate. Additionally, a single aconitase appears essential for energy metabolism, glutamate and lysine biosynthesis in respirating filamentous fungi, but not in the fermenting yeast S. cerevisiae that possesses two contributing aconitases. While yeast Aco1p is essential for the citric acid cycle and, thus, for glutamate synthesis, Aco2p specifically and exclusively contributes to lysine biosynthesis. In contrast, Aco2p homologues present in filamentous fungi were transcribed, but enzymatically inactive, revealed no altered phenotype when deleted and did not complement yeast aconitase mutants. From these results we conclude that the essential requirement of filamentous fungi for respiration versus the preference of yeasts for fermentation may have directed the evolution of aconitases contributing to energy metabolism and lysine biosynthesis.  相似文献   

8.
Sulfolobus acidocaldarius, a hyperthermoacidophilic archaeon, possesses two β-decarboxylating dehydrogenase genes, saci_0600 and saci_2375, in its genome, which suggests that it uses these enzymes for three similar reactions in lysine biosynthesis through 2-aminoadipate, leucine biosynthesis, and the tricarboxylic acid cycle. To elucidate their roles, these two genes were expressed in Escherichia coli in the present study and their gene products were characterized. Saci_0600 recognized 3-isopropylmalate as a substrate, but exhibited slight and no activity for homoisocitrate and isocitrate, respectively. Saci_2375 exhibited distinct and similar activities for isocitrate and homoisocitrate, but no detectable activity for 3-isopropylmalate. These results suggest that Saci_0600 is a 3-isopropylmalate dehydrogenase for leucine biosynthesis and Saci_2375 is a dual function enzyme serving as isocitrate-homoisocitrate dehydrogenase. The crystal structure of Saci_0600 was determined as a closed-form complex that binds 3-isopropylmalate and Mg2+, thereby revealing the structural basis for the extreme thermostability and novel-type recognition of the 3-isopropyl moiety of the substrate.  相似文献   

9.
Beta-decarboxylating dehydrogenases comprise 3-isopropylmalate dehydrogenase, isocitrate dehydrogenase, and homoisocitrate dehydrogenase. They share a high degree of amino acid sequence identity and occupy equivalent positions in the amino acid biosynthetic pathways for leucine, glutamate, and lysine, respectively. Therefore, not only the enzymes but also the whole pathways should have evolved from a common ancestral pathway. In Pyrococcus horikoshii, only one pathway of the three has been identified in the genomic sequence, and PH1722 is the sole beta-decarboxylating dehydrogenase gene. The organism does not require leucine, glutamate, or lysine for growth; the single pathway might play multiple (i.e., ancestral) roles in amino acid biosynthesis. The PH1722 gene was cloned and expressed in Escherichia coli and the substrate specificity of the recombinant enzyme was investigated. It exhibited activities on isocitrate and homoisocitrate at near equal efficiency, but not on 3-isopropylmalate. PH1722 is thus a novel, bifunctional beta-decarboxylating dehydrogenase, which likely plays a dual role in glutamate and lysine biosynthesis in vivo.  相似文献   

10.
HACN (homoaconitase) is a member of a family of [4Fe-4S] cluster-dependent enzymes that catalyse hydration/dehydration reactions. The best characterized example of this family is the ubiquitous ACN (aconitase), which catalyses the dehydration of citrate to cis-aconitate, and the subsequent hydration of cis-aconitate to isocitrate. HACN is an enzyme from the alpha-aminoadipate pathway of lysine biosynthesis, and has been identified in higher fungi and several archaea and one thermophilic species of bacteria, Thermus thermophilus. HACN catalyses the hydration of cis-homoaconitate to (2R,3S)-homoisocitrate, but the HACN-catalysed dehydration of (R)-homocitrate to cis-homoaconitate has not been observed in vitro. We have synthesized the substrates and putative substrates for this enzyme, and in the present study report the first steady-state kinetic data for recombinant HACN from T. thermophilus using a (2R,3S)-homoisocitrate dehydrogenase-coupled assay. We have also examined the products of the reaction using HPLC. We do not observe HACN-catalysed 'homocitrate dehydratase' activity; however, we have observed that ACN can catalyse the dehydration of (R)-homocitrate to cis-homoaconitate, but HACN is required for subsequent conversion of cis-homoaconitate into homoisocitrate. This suggests that the in vivo process for conversion of homocitrate into homoisocitrate requires two enzymes, in simile with the propionate utilization pathway from Escherichia coli. Surprisingly, HACN does not show any activity when cis-aconitate is substituted for the substrate, even though other enzymes from the alpha-aminoadipate pathway can accept analogous tricarboxylic acid-cycle substrates. The enzyme shows no apparent feedback inhibition by L-lysine.  相似文献   

11.
Mutants of Pichia guilliermondii were isolated that lacked homoaconitate hydratase (lys4), homoisocitrate dehydrogenase (lys10) or α-aminoadipate reductase (lys2) and were able to excrete homocitrate into the culture medium. The effects of incubation time and lysine concentration in the medium on the excretion of homocitrate were examined. In the presence of 600 mg of L-lysine/1 in a minimum salt medium P. guilliermondii G75 (lys2) produced about 280 mg homocitrate/1 during 48 h of growth. A simple procedure to isolate homocitrate from the medium is described.  相似文献   

12.
Lin Y  West AH  Cook PF 《Biochemistry》2008,47(40):10809-10815
Potassium is an activator of the reaction catalyzed by homoisocitrate (HIc) dehydrogenase (HIcDH) from Saccharomyces cerevisiae with either the natural substrate, homoisocitrate, or the slow substrate isocitrate. On the basis of initial velocity studies, the selectivity of the activator site for monovalent ions was determined. Potassium is the best activator, and NH 4 (+) and Rb (+) are also activators of the reaction, while Cs (+), Li (+), and Na (+) are not. Chloride inhibits the reaction, while acetate is much less effective. Substitution of potassium acetate for KCl changes the kinetic mechanism of HIcDH from a steady state random to a fully ordered mechanism with the binding of MgHIc followed by K (+) and NAD. The change in mechanism likely reflects an apparent increase in the affinity of enzyme for MgHIc as a result of elimination of the inhibitory effect of Cl (-). The V/K NAD pH-rate profile in the absence of K (+) exhibits a >10-fold decrease in the affinity of enzyme for NAD upon deprotonation of an enzyme side chain with a p K a of about 5.5-6. On the other hand, the affinity for NAD is relatively constant at high pH in the presence of 200 mM KCl. Since the affinity of the dinucleotide decreases as the enzyme group is protonated and the effect is overcome by a monovalent cation, the enzyme residue may be a neutral acid, aspartate or glutamate. Data suggest that K (+) replaces the proton, and likely binds to the enzyme residue, the pyrophosphoryl moiety of NAD, or both. Viscosity and solvent deuterium isotope effects studies suggest the isomerization of E-MgHIc binary complex limits the rate in the absence of K (+).  相似文献   

13.
14.
Mycelial cultures of 76 strains of lignocellulose-degrading basidiomycete fungi were screened for the activity of pyranose dehydrogenase, a novel sugar oxidoreductase recently detected in Agaricus bisporus. Of these fungi, 37 strains belonging to seven phylogenetically related genera of mostly litter-decomposing Agaricales were positive for the dehydrogenase, based on activity assays towards D-glucose with 1,4-benzoquinone or ferricenium ion as electron acceptors, and on TLC/HPLC analyses of the reaction products. Lack of activity with O(2) as the oxidant, specificity for C-3 of D-glucose, and active extracellular secretion of the enzyme were used as criteria to differentiate pyranose dehydrogenase from pyranose 2-oxidase (EC 1.1.3.10), known to be produced by numerous wood-rotting fungi. Extracellular pyranose dehydrogenase from Macrolepiota rhacodes was heavily glycosylated. The enzyme was characterized as a 78-kDa flavoprotein under denaturing conditions and a 76-kDa native protein using gel filtration. This enzyme had a maximum extracellular activity of 4.1 U ml(-1) in 39-day liquid cultures. It exhibited broad selectivity for sugar substrates and oxidized D-glucose (K(m)=1.82) exclusively at C-3 to 3-dehydro-D-glucose (D-ribo-hexos-3-ulose), in contrast to pyranose dehydrogenases from Agaricus species, which acted at both C-3 and C-2 of D-glucose. The N-terminal sequence, AVVYRHPDEL, showed significant similarity with that reported for A. bisporus.  相似文献   

15.
The mode of glyoxylate production from acetyl-CoA was investigated in three strains of methylotrophic bacteria,Pseudomonas MA,Pseudomonas AM1 and organism PAR. This investigation was prompted by the recently reported discovery of a homoisocitrate lyase in methylotrophic bacteria and the suggested involvement of this novel enzyme in assimilation of C1 and C2 compounds as part of a homoisocitrate-glyoxylate cycle. We were unable to detect cleavage of any of the four stereoisomers of homoisocitric acid by cell-free extracts of C1-or C2-grown bacteria. Extracts of C1-grown bacteria did not catalyze condensation of glyoxylate with glutarate or production of glyoxylate from acetyl-CoA and 2-ketoglutarate. Extracts of C1-grownPseudomonas MA catalyzed cleavage of isocitrate;threo-homoisocitrate was a potent competitive inhibitor of this reaction. These results indicate that homoisocitrate cleavage does not occur in any of the methylotrophs tested. The pathway for oxidation of acetyl-CoA to glyoxylate inPseudomonas AM1 and organism PAR therefore remains obscure.  相似文献   

16.
Pyranose dehydrogenase (PDH) is a flavin-dependant sugar oxidoreductase found in the family Agaricaceae, basidiomycetes that degrade lignocellulose-rich forest litter, and is catalytically related to the fungal enzymes pyranose 2-oxidase and cellobiose dehydrogenase. It has broad substrate specificity and displays similar activities with most sugar constituents of lignocellulose including disaccharides and oligosaccharides, a number of (substituted) quinones, and metal ions are suitable electron acceptors rather than molecular oxygen. In contrast to pyranose 2-oxidase and cellobiose dehydrogenase, which oxidize regioselectively at C-2 and C-1, respectively, PDH is capable of oxidation on C-1 to C-4 as well as double oxidations, depending on the nature of the substrate. This makes it a very interesting enzyme for biocatalytic applications, as many of the reaction products are otherwise unaccessible by chemical or enzymatic means. PDH was characterized in detail in a limited number of fungi, and the first encoding genes were isolated only recently. We report here, for the first time, the heterologous expression of one of these genes, encoding the major PDH protein in Agaricus meleagris, in the filamentous fungi Aspergillus nidulans, and Aspergillus niger.  相似文献   

17.
Histochemical staining of alkaline phosphatase (ALP) and succinate dehydrogenase (SDH) activities in four arbuscular mycorrhizal fungi (Glomus intraradices, G. fasciculatum, G. monosporum and G. mosseae) and their relation to growth and metabolic activities of soybean plants were investigated in a greenhouse experiment. In general, mycorrhizal inoculation significantly increased the growth responses, phosphorus and nitrogen contents, acid and alkaline phosphatases as well as total soluble protein of soybean compared to non-mycorrhizal plants. Stimulation was related to the viability of each mycorrhizal fungus. The localization of succinate dehydrogenase (as a vital stain of metabolically active fungus) and alkaline phosphatase activity (as a potential marker of efficiency of the symbiosis) in the arbuscular mycorrhizal fungi were variable. The activity appeared in young arbuscles and intercellular hyphae, whereas the collapsed arbuscules were inactive. The histochemical staining results demonstrated that the activity of alkaline phosphatase fungi was lower than succinate dehydrogenase. The use of nitroblue tetrazolium chloride as a vital stain for SDH activity showed that all mycorrhizal infection revealed by trypan blue staining was not physiologically active. Thus, the possible utilization of these enzymes to assess the activity of mycorrhizal fungi and its relation with effectively for plant growth and mineral contents is discussed.  相似文献   

18.
The crystal structure of homoisocitrate dehydrogenase involved in lysine biosynthesis from Thermus thermophilus (TtHICDH) was determined at 1.85-A resolution. Arg85, which was shown to be a determinant for substrate specificity in our previous study, is positioned close to the putative substrate binding site and interacts with Glu122. Glu122 is highly conserved in the equivalent position in the primary sequence of ICDH and archaeal 3-isopropylmalate dehydrogenase (IPMDH) but interacts with main- and side-chain atoms in the same domain in those paralogs. In addition, a conserved Tyr residue (Tyr125 in TtHICDH) which extends its side chain toward a substrate and thus has a catalytic function in the related beta-decarboxylating dehydrogenases, is flipped out of the substrate-binding site. These results suggest the possibility that the conformation of the region containing Glu122-Tyr125 is changed upon substrate binding in TtHICDH. The crystal structure of TtHICDH also reveals that the arm region is involved in tetramer formation via hydrophobic interactions and might be responsible for the high thermotolerance. Mutation of Val135, located in the dimer-dimer interface and involved in the hydrophobic interaction, to Met alters the enzyme to a dimer (probably due to steric perturbation) and markedly decreases the thermal inactivation temperature. Both the crystal structure and the mutation analysis indicate that tetramer formation is involved in the extremely high thermotolerance of TtHICDH.  相似文献   

19.
20.
Polyols, or polyhydroxy alcohols, are produced by many fungi. Saccharomyces cerevisiae produces large amounts of glycerol, and several fungi that cause serious human infections produce D-arabinitol and mannitol. Glycerol functions as an intracellular osmolyte in S. cerevisiae, but the functions of D-arabinitol and mannitol in pathogenic fungi are not yet known. To investigate the functions of mannitol, we constructed a new mannitol biosynthetic pathway in S. cerevisiae. S. cerevisiae transformed with multicopy plasmids encoding the mannitol-1-phosphate dehydrogenase of Escherichia coli produced mannitol, whereas S. cerevisiae transformed with control plasmids did not. Although mannitol production had no obvious phenotypic effects in wild-type S. cerevisiae, it restored the ability of a glycerol-defective, osmosensitive osg1-1 mutant to grow in the presence of high NaCl concentrations. Moreover, osg1-1 mutants producing mannitol were more resistant to killing by oxidants produced by a cell-free H2O2-FeSO4-NaI system than were controls. These results indicate that mannitol can (i) function as an intracellular osmolyte in S. cerevisiae, (ii) substitute for glycerol as the principal intracellular osmolyte in S. cerevisiae, and (iii) protect S. cerevisiae from oxidative damage by scavenging toxic oxygen intermediates.  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