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.
Khalil S  Pawelek PD 《Biochemistry》2011,50(4):533-545
The Escherichia coli siderophore enterobactin is synthesized in response to iron starvation. 2,3-Dihydro-2,3-dihydroxybenzoate dehydrogenase (EntA) produces 2,3-dihydroxybenzoate (DHB), a biosynthetic intermediate. 2,3-Dihydroxybenzoate-AMP ligase (EntE) adenylates DHB, activating it for attachment to the NRPS substrate holo-EntB. Using analytical ultracentrifugation, we found that EntA undergoes concentration-dependent dimer-tetramer self-association (K(D) = 12.3 μM). We further found that EntA can form a specific complex with EntE. Pull-down assays revealed that recombinant EntA bait pulled down EntE from E. coli lysates, whereas recombinant EntE bait could pull down EntA. Addition of the SMCC cross-linker to a mixture of EntA and EntE resulted in a cross-linked product with a molecular mass of >250 kDa, suggesting a complex stoichiometry of one EntA tetramer and four EntE monomers. The effect of EntA on EntE activity was also examined. Addition of a 4-fold excess of EntA to an EntE assay mixture resulted in a 6-fold stimulation of EntE activity. EntA was also found to perturb the FRET signal between EntE donor residues and EntE-bound DHB. By following the EntA-dependent decrease in the magnitude of the EntE-DHB FRET signal, EntA-EntE binding behavior was found to be sigmoidal, suggesting the presence of both low- and high-affinity binding sites. The EntA-EntE interaction was also directly measured by isothermal titration calorimetry at 10 °C. The resulting binding isotherm fit well to a model describing two binding sites, supporting our AUC and fluorescence data. Taken together, our data show that tetrameric EntA optimally interacts with EntE, resulting in an enhancement of EntE activity.  相似文献   

2.
Degradation of 2-chlorobenzoate by Pseudomonas cepacia 2CBS   总被引:10,自引:0,他引:10  
A bacterium was isolated from water by enrichment on 2-chlorobenzoate as sole source of carbon and energy. Based on morphological and physiological properties, this microorganism was assigned to the species Pseudomonas cepacia. The organism was designated Pseudomonas cepacia 2CBS. During growth on 2-chlorobenzoate, the chlorine substituent was released quantitatively, and a small amount of 2,3-dihydroxybenzoate accumulated in the culture medium. Mutants of Pseudomonas cepacia 2CBS were induced by treatment with N-methyl-N'-nitro-N-nitrosoguanidine. Some of these mutants produced catechol from 2-chlorobenzoate. Other mutants accumulated the meta-cleavage product of catechol, 2-hydroxy-cis,cis-muconic acid semialdehyde. In crude cell-free extracts of Pseudomonas cepacia 2CBS, an enzyme was detected which catalysed the conversion of 2-chlorobenzoate to catechol. Molecular oxygen, NADH and exogenous Fe2+ were required for activity. Stoichiometric amounts of chloride were released. Experiments with 18O2 revealed that both oxygen atoms in the hydroxyl groups of the product were derived from molecular oxygen. Thus, the enzyme catalysing the conversion of 2-chlorobenzoate was identified as 2-chlorobenzoate 1,2-dioxygenase (1,2-hydroxylating, dehalogenating, decarboxylating). 2-Chlorobenzoate 1,2-dioxygenase from Pseudomonas cepacia 2CBS was shown to be a multicomponent enzyme system. The activities of catechol 2,3-dioxygenase and catechol 1,2-dioxygenase were detected in crude cell-free extracts. The activity of catechol 2,3-dioxygenase was 60 times higher than the activity of catechol 1,2-dioxygenase, indicating that catechol is mainly degraded via meta-cleavage in Pseudomonas cepacia 2CBS. No enzyme was found which converted 2,3-dihydroxybenzoate, suggesting that this compound is a dead-end metabolite of 2-chlorobenzoate catabolism. A pathway for the degradation of 2-chlorobenzoate by Pseudomonas cepacia 2CBS is proposed.  相似文献   

3.
Enzymes of the p-cymene pathway in Pseudomonas putida strains cometabolized the intermediate analogue 4-trifluoromethyl(TFM)benzoate. Three products, 4-TFM-2,3-dihydro-2,3-dihydroxybenzoate, 4-TFM-2,3-dihydroxy-benzoate and 2-hydroxy-6-oxo-7,7,7-trifluorohepta-2,4-dienoate (7-TFHOD) were identified chemically and by spectroscopic proterties.Certain TFM-substituted analogue metabolites of the p-cymene pathway were transformed at drastically reduced rates.Hammett type analysis of ring cleavage reactions of 4-substituted 2,3-dihydroxybenzoates revealed the negative inductive and especially mesomeric effect of substituents to be rate determining. Whereas decarboxylation of 3-carboxy-7-TFHOD was not affected by fluorine substitution the subsequent hydrolysis of 7-TFHOD proceeded very slowly. The negative inductive effect of the TFM-group probably inhibited heterolysis of the carbon bond between C5 and C6 of 7-TFHOD.Abbreviations DHB 1,2-Dihydroxy-2-hydrobenzoate - DHC 2,3-Dihydro-2,3-dihydroxybenzoate, this compound was termed DHC simply to distinguish it from the similar 1,2-dihydroxy-2-hydrobenzoate (DHB) as described in the preceeding paper (Engesser et al. 1988) - HMS 2-Hydroxymuconic semialdehyde - HOD 2-Hydroxy-6-oxohepta-2,4-dienoate - 7-TFHOD 2-Hydroxy-6-oxo-7,7,7-trifluorohepta-2,4-dienoate - TFM Trifluoromethyl This work was supported, in part, by the Gesellschaft für Strahlen- und Umweltforschung, Neuherberg/München, FRG  相似文献   

4.
Abstract Pseudomonas sp. strain RW611 utilized the ammonium salt of 2-sulfobenzoate as sole source of carbon, sulfur, nitrogen, and energy. The xenobiotic sulfo substituent was dioxygenolytically eliminated as sulfite, which was then slowly oxidized to sulfate. 2,3-Dihydroxybenzoate, which resulted from desulfonation underwent meta -cleavage, mediated by 2,3-dihydroxybenzoate 3,4-dioxygenase activity. This enzyme was inhibited by 3-chlorocatechol and 2,3,4-trihydroxybenzoate.  相似文献   

5.
Plasmid-carrying Pseudomonas putida strains degrade naphthalene through different biochemical pathways. The influence of various combinations of host bacteria and plasmids on growth characteristics and competitiveness of P. putida strains was studied in chemostat culture at a low dilution rate (D=0.05 h−1) with naphthalene as the sole source of carbon and energy. Under naphthalene limitation, the plasmid-bearing strains degrading naphthalene that use catechol 1,2-dioxygenase for catechol oxidation (ortho pathway), were the most competitive. The strains bearing plasmids that control naphthalene catabolism via catechol 2,3-dioxygenase (meta pathway), were less competitive. Under these conditions the strain carrying plasmid pBS4, which encodes for naphthalene catabolism via gentisic acid, was the least competitive. Received: 24 February 1997 / Received revision: 22 May 1997 / Accepted: 25 May 1997  相似文献   

6.
Proteome analysis of Bacillus subtilis cells grown at low and high salinity revealed the induction of 16 protein spots and the repression of 2 protein spots, respectively. Most of these protein spots were identified by mass spectrometry. Four of the 16 high-salinity-induced proteins corresponded to DhbA, DhbB, DhbC, and DhbE, enzymes that are involved in the synthesis of 2,3-dihydroxybenzoate (DHB) and its modification and esterification to the iron siderophore bacillibactin. These proteins are encoded by the dhbACEBF operon, which is negatively controlled by the central iron regulatory protein Fur and is derepressed upon iron limitation. We found that iron limitation and high salinity derepressed dhb expression to a similar extent and that both led to the accumulation of comparable amounts of DHB in the culture supernatant. DHB production increased linearly with the degree of salinity of the growth medium but could still be reduced by an excess of iron. Such an excess of iron also partially reversed the growth defect exhibited by salt-stressed B. subtilis cultures. Taken together, these findings strongly suggest that B. subtilis cells grown at high salinity experience iron limitation. In support of this notion, we found that the expression of several genes and operons encoding putative iron uptake systems was increased upon salt stress. The unexpected finding that high-salinity stress has an iron limitation component might be of special ecophysiological importance for the growth of B. subtilis in natural settings, in which bioavailable iron is usually scarce.  相似文献   

7.
Culture filtrates of virulent Nocardia asteroides GUH-2 after growth in acetate minimal medium displayed an absorbance maximum at 320 nm. After isolation by polyamide extraction and anion chromatography, a UV-active compound with this absorbance was shown to be 2,3-dihydroxybenzoic acid (DHB) by nuclear magnetic resonance, gas chromatographic, and mass spectrometric techniques. DHB production under several culture conditions was quantified by a standard high-pressure liquid chromatography assay. Under iron deficiency conditions, N. asteroides GUH-2 excreted up to 11 mg of DHB per liter into the culture medium. No DHB was detected when N. asteroides GUH-2 was grown in an iron-rich medium. With the less virulent strain N. asteroides 10905, DHB was not found under any condition tested.  相似文献   

8.
Biogenic amines (BAs) are produced primarily by microorganisms found in fermented foods and are often implicated in food poisoning. BA-producing bacteria found in fermented soybean pastes were isolated and characterized using a decarboxylating medium and multiplex PCR analysis. Two BA-producing bacteria were isolated from traditional soybean pastes: one was a histamine-producing Clostridium strain, and the other was a tyramine-producing Pseudomonas strain. The Clostridium strain was determined to be a potent histamine producer among the cultures tested. Synthesis of tyramine by Pseudomonas sp. T1 was observed for the first time in this study.  相似文献   

9.
Azotobacter vinelandii requires a high complement of iron and an efficient iron acquisition system to support nitrogen fixation. To circumvent problems inherent in batch culture trace metal studies, continuous cultures were used to measure the response of A. vinelandii to iron stress. Iron was found to be growth limiting for nitrogen-fixing A. vinelandii at a concentration as high as 12.5 muM; iron was growth sufficient at 25 muM. Iron-stressed A. vinelandii in continuous culture formed 2,3-hydroxybenzoic acid (DHB), 2-N,6-N-di-(2,3-dihydroxybenzoyl)-l-lysine (DHBL), and a chromophoric yellow-green fluorescent peptide (YGFP). At a fixed dilution rate of 0.1 h, steady-state growth occurred at growth-limiting iron concentrations. DHB and DHBL were quantitatively measured during iron-limited steady states and iron-sufficient states by Arnow colorimetric assays. YGFP was determined by absorbance measurements taken at 380 nm, and the concentration was calculated from the reported specific absorption coefficient. Biomass increased and DHBL, DHB, and YGFP concentrations decreased as the concentration of growth-limiting iron was increased in the culture vessel and medium reservoirs. DHBL was the major siderophore and YGFP was the minor siderophore species produced during iron-limited equilibrium growth. A low level of DHB and YGFP, but no DHBL, was formed under iron-sufficient conditions. These results provide further physiological evidence that DHB, YGFP, and especially DHBL may function as siderophores in nitrogen-fixing A. vinelandii.  相似文献   

10.
11.
Iron transport in Escherichia coli K-12   总被引:14,自引:0,他引:14  
The study of iron uptake promoted by 2,3-dihydroxybenzoate (DHB) into Escherichia coli K-12 aroB mutants allowed some dissection of outer and cytoplasmic membrane functions. These strains are unable to produce the iron-transporting chelate enterochelin, unless fed with a precursor such as DHB. When added to the medium, enterochelin and its natural breakdown products, the linear dimer and trimer of 2,3-dihydroxybenzoylserine (DBS), efficiently transported iron via the feuB, tonB and fep gene products. Thus mutants in these genes were defective in transport of the above chelates. However, feuB and tonB mutants were able to take up iron when DHB was added to the medium. Thus DHB-promoted iron uptake bypassed two functions required for the transport of ferric-enterochelin from the medium. One of these functions, feuB, has been shown to be an outer membrane protein. In contrast to three other iron transport systems including ferric-enterochelin uptake, DHB-promoted iron uptake was little affected by the uncoupler 2,4-dinitrophenol. Dissipation of the energized state of the cytoplasmic membrane apparently only affects those iron transport systems which require an outer membrane protein. Since DHB-promoted iron uptake bypasses the feuB outer membrane protein and the tonB function, it is concluded that, in ferricenterochelin transport, the tonB gene may function in coupling the energized state of the cytoplasmic membrane to the protein-dependent outer membrane permeability. DHB-promoted iron uptake required the synthesis and enzymatic breakdown of enterochelin as judged by the effects of the entF and fesB mutations. A fep mutant was not only deficient in the transport of the ferric chelates of enterochelin and its breakdown products, but was also deficient in DHB-promoted iron uptake. A scheme is presented in which iron diffuses as DHB-complex through the outer membrane, and is subsequently captured by enterochelin or DBS dimer or trimer and translocated across the cytoplasmic membrane.List of Abbreviations DHB 2,3-dihydroxybenzoate - DBS 2,3-dihydroxybenzoylserine - NTA nitrilotriacetate - DNP 2,4-dinitrophenol  相似文献   

12.
Itoic Acid Synthesis in Bacillus subtilis   总被引:8,自引:5,他引:3  
Under conditions of iron deficiency, strains of Bacillus subtilis produced 2,3-dihydroxybenzoic acid (DHB), 2,3-dihydroxybenzolyglycine (DHBG), or both of these compounds. DHB(G) production [production of DHB(G) refers to the production of DHB, or DHBG, or both] was proportional to the amount of iron present and occurred logarithmically, paralleling growth. Supplementation of media with more than 150 mug of iron per liter at zero-time inhibited DHB accumulation completely. In the presence of DHB, lower levels of iron inhibited DHB(G) production, so that the actual inhibitor of synthesis may involve the Fe(3+):[DHB(G)](3) complex. The strains producing DHBG also produced coproporphyrin III during iron-deficient growth, whereas a strain producing DHB did not produce coproporphyrin III under these conditions. Accumulation of DHB(G) was influenced by the levels of aromatic amino acids and anthranilic acid in the medium. In vivo experiments with strain B-1471 demonstrated that DHB was coupled to added glycine to form DHBG. Metabolism of DHB(G) was observed in two of the strains studied.  相似文献   

13.
The occurrence of ferrichrome-type hydroxamate siderophores in soil was confirmed. In the presence of the iron-scavenging chelator ethylenediamine[di(o-hydroxyphenylacetic)acid], soil extract stimulated the growth of an Escherichia coli strain possessing the ferrichrome transport protein (TonA) but did not stimulate growth of a strain lacking this protein (TonA). The siderophore concentration in a 1:1 (soil-water) extract was estimated to be approximately 78 nM. Specificity of the assay was supported by the absence of significant differential strain responses to ferric citrate, ferric 2,3-dihydroxybenzoate, enterochelin, ferrioxamine B, coprogen, and triacetylfusigen.  相似文献   

14.
Pseudomonas sp. strain PP2 isolated in our laboratory efficiently metabolizes phenanthrene at 0.3% concentration as the sole source of carbon and energy. The metabolic pathways for the degradation of phenanthrene, benzoate and p-hydroxybenzoate were elucidated by identifying metabolites, biotransformation studies, oxygen uptake by whole cells on probable metabolic intermediates, and monitoring enzyme activities in cell-free extracts. The results obtained suggest that phenanthrene degradation is initiated by double hydroxylation resulting in the formation of 3,4-dihydroxyphenanthrene. The diol was finally oxidized to 2-hydroxymuconic semialdehyde. Detection of 1-hydroxy-2-naphthoic acid, alpha-naphthol, 1,2-dihydroxy naphthalene, and salicylate in the spent medium by thin layer chromatography; the presence of 1,2-dihydroxynaphthalene dioxygenase, salicylaldehyde dehydrogenase and catechol-2,3-dioxygenase activity in the extract; O(2) uptake by cells on alpha-naphthol, 1,2-dihydroxynaphthalene, salicylaldehyde, salicylate and catechol; and no O(2) uptake on o-phthalate and 3,4-dihydroxybenzoate supports the novel route of metabolism of phenanthrene via 1-hydroxy-2-naphthoic acid --> [alpha-naphthol] --> 1,2-dihydroxy naphthalene --> salicylate --> catechol. The strain degrades benzoate via catechol and cis,cis-muconic acid, and p-hydroxybenzoate via 3,4-dihydroxybenzoate and 3-carboxy- cis,cis-muconic acid. Interestingly, the culture failed to grow on naphthalene. When grown on either hydrocarbon or dextrose, the culture showed good extracellular biosurfactant production. Growth-dependent changes in the cell surface hydrophobicity, and emulsification activity experiments suggest that: (1) production of biosurfactant was constitutive and growth-associated, (2) production was higher when cells were grown on phenanthrene as compared to dextrose and benzoate, (3) hydrocarbon-grown cells were more hydrophobic and showed higher affinity towards both aromatic and aliphatic hydrocarbons compared to dextrose-grown cells, and (4) mid-log-phase cells were significantly (2-fold) more hydrophobic than stationary phase cells. Based on these results, we hypothesize that growth-associated extracellular biosurfactant production and modulation of cell surface hydrophobicity plays an important role in hydrocarbon assimilation/uptake in Pseudomonas sp. strain PP2.  相似文献   

15.
Recent studies demonstrated that 2,3-dihydroxybiphenyl 1,2-dioxygenase from Burkholderia sp. strain LB400 (DHBDLB400; EC 1.13.11.39) cleaves chlorinated 2,3-dihydroxybiphenyls (DHBs) less specifically than unchlorinated DHB and is competitively inhibited by 2',6'-dichloro-2,3-dihydroxybiphenyl (2',6'-diCl DHB). To determine whether these are general characteristics of DHBDs, we characterized DHBDP6-I and DHBDP6-III, two evolutionarily divergent isozymes from Rhodococcus globerulus strain P6, another good polychlorinated biphenyl (PCB) degrader. In contrast to DHBDLB400, both rhodococcal enzymes had higher specificities for some chlorinated DHBs in air-saturated buffer. Thus, DHBDP6-I cleaved the DHBs in the following order of specificity: 6-Cl DHB > 3'-Cl DHB approximately DHB approximately 4'-Cl DHB > 2'-Cl DHB > 4-Cl DHB > 5-Cl DHB. It also cleaved its preferred substrate, 6-Cl DHB, three times more specifically than DHB. Interestingly, some of the worst substrates for DHBDP6-I were among the best for DHBDP6-III (4-Cl DHB > 5-Cl DHB approximately 6-Cl DHB approximately 3'-Cl DHB > DHB > 2'-Cl DHB approximately 4'-Cl DHB; DHBDP6-III cleaved 4-Cl DHB two times more specifically than DHB). Generally, each of the monochlorinated DHBs inactivated the enzymes more rapidly than DHB. The exceptions were 4-Cl DHB for DHBDP6-I and 2'-Cl DHB for DHBDP6-III. As observed in DHBDLB400, chloro substituents influenced the reactivity of the dioxygenases with O2. For example, the apparent specificities of DHBDP6-I and DHBDP6-III for O2 in the presence of 2'-Cl DHB were lower than those in the presence of DHB by factors of >60 and 4, respectively. DHBDP6-I and DHBDP6-III shared the relative inability of DHBDLB400 to cleave 2',6'-diCl DHB (apparent catalytic constants of 0.088 +/- 0.004 and 0.069 +/- 0.002 s(-1), respectively). However, these isozymes had remarkably different apparent K(m) values for this compound (0.007 +/- 0.001, 0.14 +/- 0.01, and 3.9 +/- 0.4 micro M for DHBDLB400, DHBDP6-I, and DHBDP6-III, respectively). The markedly different reactivities of DHBDP6-I and DHBDP6-III with chlorinated DHBs undoubtedly contribute to the PCB-degrading activity of R. globerulus P6.  相似文献   

16.
Four enzymic components, coded for by the entD, entE, entF and entG genes, involved in the biosynthesis of enterochelin from 2,3-dihydroxybenzoate have been separated from cell extracts of mutant strains of Escherichia coli K-12.The starting material for fractionation of the E, F and G components was a cell extract of an entD mutant strain, which yielded the E, F and G enzymic components uncontaminated by a functional D component. The D component was isolated from cell extracts of an entE mutant strain. The conversion of 2,3-dihydroxybenzoate and l-serine into enterochelin is dependent on the presence of all four enzymic components.The E and F components were shown to catalyze ATP-pyrophosphate exchange reactions dependent on 2,3-dihydroxybenzoate and l-serine, respectively, whereas fractionated extracts of the entE and entF mutant strains lacked these reactions. These data provide firm evidence that the E and F components are involved in the initial activation of the substrates. The D and G components are necessary for subsequent and, as yet, undefined reactions.  相似文献   

17.
Studies were carried out to determine the factors governing the induction of anthranilate hydroxylase and other enzymes in the pathway for the dissimilation of anthranilate by Aspergillus niger (UBC 814). The enzyme was induced by growth in the presence of tryptophan, kynurenine, anthranilate, and, surprisingly, by 3-hydroxyanthranilate, which was not an intermediate in the conversion of anthranilate to 2,3-dihydroxybenzoate. There was an initial lag in the synthesis of anthranilate hydroxylase when induced by tryptophan, anthranilate, and 3-hydroxyanthranilate. Cycloheximide inhibited the enzyme induction. Comparative studies on anthranilate hydroxylase, 2,3-dihydroxybenzoate carboxy-lyase, and catechol 1:2-oxygenase revealed that these enzymes were not coordinately induced by either anthranilate or 3-hydroxyanthranilate. Structural requirements for the induction of anthranilate hydroxylase were determined by using various analogues of anthranilate. The activity of the constitutive catechol oxygenase was increased threefold by exposure to anthranilate, 2,3-dihydroxybenzoate, or catechol. 3-Hydroxyanthranilate did not enhance the levels of catechol oxygenase activity.  相似文献   

18.
The entC and entA genes, coding for the enzymes isochorismate synthase and 2,3-dihydro-2,3-dihydroxybenzoate dehydrogenase, respectively, were subcloned behind the T7 promoter in the expression plasmid pGEM3Z. Their protein products were overproduced and partially purified for in vitro analysis of the conversion of chorismate to isochorismate. Whereas previous genetic experiments suggested that the EntA enzyme has a role in this conversion, this study clearly indicates that EntC alone catalyzes the reaction. Addition of EntA had no effect on isochorismate synthase activity. As a result, the mutation (previously designated entC401) in strain AN191 was characterized by nucleotide sequence analysis. The lesion is a single base substitution in the entA gene, resulting in a glutamic acid-for-glycine substitution at the penultimate amino acid (residue 247) of the EntA enzyme. The mutant protein was partially purified and shown to be devoid of 2,3-dihydro-2,3-dihydroxybenzoate dehydrogenase activity, whereas the entC gene product from strain AN191 exhibited normal isochorismate synthase function. These results conflict with the earlier characterization of the entC401 mutation in a different genetic background. The data presented herein establish that the EntA protein does not contribute to isochorismate synthase activity and that the mutant strain that led to this suggestion harbors a defective allele of entA rather than entC.  相似文献   

19.
3-Chlorobiphenyl is known to be mineralized by biphenyl-utilizing bacteria to 3-chlorobenzoate, which is further metabolized to 3-chlorocatechol. An extradiol dioxygenase, 2,3-dihydroxybiphenyl 1,2-dioxygenase (DHB12O; EC 1.13.11.39), which is encoded by the bphC gene, catalyzes the third step of the upper pathway of 3-chlorobiphenyl degradation. In this study, two full-length bphCs and nine partial fragments of bphCs fused to the 3' end of bphC in Pseudomonas pseudoalcaligenes KF707 were cloned from different biphenyl-utilizing soil bacteria and expressed in Escherichia coli. The enzyme activities of the expressed DHB12Os were inhibited to varying degrees by 3-chlorocatechol, and the E. coli cells overexpressing DHB12O could not grow or grew very slowly in the presence of 3-chlorocatechol. These sensitivities of enzyme activity and cell growth to 3-chlorocatechol were well correlated, and this phenomenon was employed in screening chimeric BphCs formed by family shuffling of bphC genes isolated from Comamonas testosteroni KF704 and C. testosteroni KF712. The resultant DHB12Os were more resistant by a factor of two to 3-chlorocatechol than one of the best parents, KF707 DHB12O.  相似文献   

20.
Excretion of the metal-chelating phenolic acid, 2,3-dihydroxybenzoate, by a tryptophan-requiring strain (M-13) of Bacillus subtilis was inversely proportional to the iron added to the medium. Addition of iron as the ferric chelates of two secondary hydroxamates (ferri-schizokinen and Desferal) markedly reduced excretion. Synthesis of 2,3-dihydroxybenzoate from chorismate by extracts of B. subtilis M-13, grown in low-iron medium, was unaltered by additions of FeSO(4), FeCl(3), ferrischizokinen, 2,3-dihydroxybenzoate, the 2,3-dihydroxybenzoate-iron complex, or by extracts of cells grown in high-iron medium (which contained no demonstrable 2,3-dihydroxybenzoate-synthesizing activity) to the extracts of "low-iron cells." Iron control seemed to involve repression of synthesis of the enzymes in the 2,3-dihydroxybenzoate pathway. Both ferri-schizokinen and 2,3-dihydroxybenzoate plus iron enhanced considerably the otherwise minimal repressive effects of iron at low concentrations. Ferri-schizokinen delayed derepression of the pathway in B. subtilis M-13, and reduced its rate of synthesis after derepression. Addition of FeSO(4) to derepressed cells of B. subtilis M-13 halted synthesis of the enzymes after a lag period. The effect of the ferric hydroxamates was related to the capacity of B. subtilis M-13 to incorporate (59)Fe(3+) from Desferal-(59)Fe(3+). Cellular accumulation of (59)Fe(3+) from Desferal-(59)Fe(3+) after 20 min was nearly double that incorporated from (59)FeCl(3).  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