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.
 The first step in the degradation of 2,4-dichlorophenoxyacetic acid (2,4-D) by Ralstonia eutropha JMP134 is catalyzed by the α-ketoglutarate (α-KG)-dependent dioxygenase TfdA. Previously, EPR and ESEEM studies on inactive Cu(II)-substituted TfdA suggested a mixture of nitrogen/oxygen coordination with two imidazole-like ligands. Differences between the spectra for Cu TfdA and α-KG- and 2,4-D-treated samples were interpreted as a rearrangement of the g–tensor principal axis system. Herein, we report the use of X-ray absorption spectroscopy (XAS) to further characterize the metal coordination environment of Cu TfdA as well as that in the active, wild-type Fe(II) enzyme. The EXAFS data are interpreted in terms of four N/O ligands (two imidazole-like) in the Cu TfdA sample and six N/O ligands (one or two imidazole-like) in the Fe TfdA sample. Addition of α-KG results in no significant structural change in coordination for Cu or Fe TfdA. However, addition of 2,4-D results in a decrease in the number of imidazole ligands in both Cu and Fe TfdA. Since this change is seen both in the Fe and Cu EXAFS, loss of one histidine ligand upon 2,4-D addition best describes the phenomenon. These XAS data clearly demonstrate that changes occur in the atomic environment of the metallocenter upon substrate binding. Received: 3 July 1998 / Accepted: 13 October 1998  相似文献   

2.
Botulinum neurotoxins are the most toxic proteins currently known. Based on a recently identified potent lead structure, 2,4-dichlorocinnamic acid hydroxamate, herein we report on the structure-activity relationship of a series of hydroxamate BoNT/A inhibitors. Among them, 2-bromo-4-chlorocinnamic acid hydroxamate, 2-methyl-4-chlorocinnamic acid hydroxamate, and 2-trifluoromethyl-4-chlorocinnamic acid hydroxamate displayed comparable inhibitory activity to that of the lead structure.  相似文献   

3.
TfdA is an Fe(II)- and alpha-ketoglutarate- (alphaKG-) dependent dioxygenase that hydroxylates the herbicide 2,4-dichlorophenoxyacetic acid (2,4-D) producing a hemiacetal that spontaneously decomposes to 2,4-dichlorophenol and glyoxylate. On the basis of a recently published TfdA structural model [Elkins et al. (2002) Biochemistry 41, 5185-5192], His214, Lys71, Arg278, and the backbone amide of Ser117 are suggested to bind the 2,4-D carboxylate; Lys95 and possibly Lys71 are hypothesized to interact with the 2,4-D ether atom; and Arg274 and Thr141 are suspected to bind alphaKG. TfdA variants with substitutions at these and other positions were purified and characterized in order to explore the roles of these residues in catalysis. The K71L, K71Q, K95L, K95Q, R274Q, R274L, and R278Q variants exhibited significantly increased 2,4-D K(m), alphaKG K(m), and alphaKG K(d) values, consistent with their proposed roles in substrate binding. A protease-sensitive site was successfully eliminated in the R78Q variant, which also exhibited decreased affinity for 2,4-D. In contrast, the Y81F, Y126F, T141V, Y169F, and Y244F variants showed only modest changes in their kinetics. An observed 4-fold lower K(m) of the K95L variant compared to wild-type protein with the alternative substrate 2,4-dichlorocinnamic acid provided additional evidence for an interaction between Lys95 and the 2,4-D ether atom. Phenylpropiolic acid was identified as a mechanism-based inactivator of the enzyme [K(i) = 38.1 +/- 6.0 microM and k(inact)(max) = 2.3 +/- 0.1 min(-1)]. This acetylenic compound covalently modifies a peptide (166-AEHYALNSR-174) that is predicted to form one side of the substrate-binding pocket. The K95L variant of TfdA was not inactivated by phenylpropiolic acid, providing added support that Lys95 is present at the active site. These results support the identity of suspected substrate-binding residues derived from structural modeling studies and extend our understanding of the oxidative chemistry carried out by TfdA.  相似文献   

4.
Purified 2,4-dichlorophenoxyacetic acid (2,4-D)/(alpha)-ketoglutarate dioxygenase (TfdA) was shown to use 4-nitrophenoxyacetic acid (K(infm) = 0.89 (plusmn) 0.04 mM, k(infcat) [catalytic constant] = 540 (plusmn) 10 min(sup-1)), producing intensely yellow 4-nitrophenol. This reagent was used to develop a rapid, continuous, colorimetric assay for the detection of TfdA and analogous activities in 2,4-D-degrading bacterial cells and extracts.  相似文献   

5.
Two α-ketoglutarate-dependent dioxygenases carrying enantiospecific activity for the etherolytic cleavage of racemic phenoxypropionate herbicides [(RS)-2-(2,4-dichlorophenoxy)propionate and (RS)-2-(4-chloro-2-methylphenoxy)propionate] from Delftia acidovorans MC1 were characterized with respect to protein and sequence data. The (S)-phenoxypropionate/α-ketoglutarate-dioxygenase (SdpA) appeared as a monomeric enzyme with a molecular weight of 32 kDa in the presence of SDS. N-terminal sequences revealed relationship to α-ketoglutarate-dependent taurine dioxygenase (TauD) and to 2,4-dichlorophenoxyacetate/α-ketoglutarate-dioxygenase (TfdA). The (R)-phenoxypropionate/α-ketoglutarate-dioxygenase (RdpA) referred to 36 kDa in the presence of SDS and to 108 kDa under native conditions. Internal sequences of fragments obtained after digestion made evident relationship to TfdA and TauD. Two-dimensional electrophoretic separation resulted in the resolution of up to 3 individual spots with almost identical molecular weights but different isoelectric points with both RdpA and SdpA. The structural differences of these isoenzyme forms are not yet clear.  相似文献   

6.
1. Simultaneous oxidation of C14-methyl-labeled acetate, and unlabeled malate or fumarate and α-ketoglutarate results in entrapment of labeled carbon in the C4-dicarboxylic acids, but not in α-ketoglutarate, although all substrates are utilized at comparable rates. 2. A large endogenous reduction of all C4-dicarboxylic acids (fumarate, oxalacetate, and malate) to succinate is observed under aerobic conditions, and when vigorous oxidation is proceeding. This effect occurs with both freshly harvested young (18 hour) cells and stored (2 week) cells. 3. This reduction can be considerably minimized under high oxygen tensions. 4. The quantitative concordance of these results with a Thunberg-Knoop cyclic mechanism for acetate oxidation is shown. Possible alternative C4 products formed prior to succinate are not completely excluded, but it appears that the cells can utilize the succinate condensation as a major pathway in acetate oxidation.  相似文献   

7.
MenD as a versatile catalyst for asymmetric synthesis   总被引:1,自引:0,他引:1  
The thiamine diphosphate (ThDP)-dependent enzyme 2-succinyl-5-enolpyruvyl-6-hydroxy-3-cyclohexene-1-carboxylate synthase (MenD) from Escherichia coli K12, formerly known as SHCHC-synthase, catalyses the decarboxylation of α-ketoglutarate and the subsequent addition of the resulting succinyl-THDP to isochorismate. Here, the enzyme is tested for unphysiologial C–C bond-forming reactions.Condensation of α-ketoglutarate after decarboxylation to a broad range of aldehydes gave α-hydroxyketones with isolated yields from 26 to 87% and 94 to 98% ee for addition to aromatic aldehydes. MenD accepts a wide range of aldehydes as acceptor substrates to produce chiral α-hydroxyketones with conserved regioselectivity where the activated succinylsemialdehyde serves selectively as the donor. Regioselectivity is inverted only for condensation of α-ketoglutarate with pyruvate (activated acetaldehyde) as donor. Besides α-ketoglutarate, pyruvate and oxalacetate are accepted as donors in combination with benzaldehyde and 2-fluorobenzaldehyde as acceptors, however with decreased activity of C–C bond formation.The physiological 1,4-addition of α-ketoglutarate to isochorismate was investigated for acceptor substrate variability. (2S,3S)-2,3-Dihydroxy-2,3-dihydrobenzoate (2,3-CHD), which lacks the pyruvyl found in isochorismate, is converted to (5S,6S)-2-succinyl-5,6-dihydroxycyclohex-2-enecarboxylate. In contrast to the addition to carbonyls, the active site of MenD does appear to impose specific constraints on the acceptor substrate for 1,4-addition with α,β-unsaturated carboxylic acids.  相似文献   

8.
2,4-dichlorophenoxyacetic acid (2,4-D)/alpha-ketoglutarate (alpha-KG) dioxygenase (TfdA) is an Fe(II)-dependent enzyme that catalyzes the first step in degradation of the herbicide 2,4-D. The active site structures of a small number of enzymes within the alpha-KG-dependent dioxygenase superfamily have been characterized and shown to have a similar HXDX(50-70)HX(10)RXS arrangement of residues that make up the binding sites for Fe(II) and alpha-KG. TfdA does not have obvious homology to the dioxygenases containing the above motif but is related in sequence to eight other enzymes in the superfamily that form a distinct consensus sequence (HX(D/E)X(138-207) HX(10)R/K). Variants of TfdA were created to examine the roles of putative metal-binding residues and the functions of the other seven histidines in this protein. The H167A, H200A, H213A, H245A, and H262A forms of TfdA formed inclusion bodies when overproduced in Escherichia coli DH5alpha; however, these proteins were soluble when fused to the maltose-binding protein (MBP). MBP-TfdA exhibited kinetic parameters similar to the native enzyme. The H8A and H235A variants were catalytically similar to wild-type TfdA. MBP-H213A and H216A TfdA have elevated K(m) values for 2,4-D, and the former showed a decreased k(cat), suggesting these residues may affect substrate binding or catalysis. The H113A, D115A, MBP-H167A, MBP-H200A, MBP-H245A and MBP-H262A variants of TfdA were inactive. Gel filtration analysis revealed that the latter two proteins were highly aggregated. The remaining four inactive variants were examined in their Cu(II)-substituted forms by EPR and electron spin-echo envelope modulation (ESEEM) spectroscopic methods. Changes in EPR spectra upon addition of substrates indicated that copper was present at the active site in the H113A and D115A variants. ESEEM analysis revealed that two histidines are bound equatorially to the copper in the D115A and MBP-H167A TfdA variants. The experimental data and sequence analysis lead us to conclude that His-113, Asp-115, and His-262 are likely metal ligands in TfdA and that His-213 may aid in catalysis or binding of 2,4-D.  相似文献   

9.
α-Ketoglutarate-dependent (R)-dichlorprop dioxygenase (RdpA) and α-ketoglutarate-dependent (S)-dichlorprop dioxygenase (SdpA), which are involved in the degradation of phenoxyalkanoic acid herbicides in Sphingomonas herbicidovorans MH, were expressed and purified as His6-tagged fusion proteins from Escherichia coli BL21(DE3)(pLysS). RdpA and SdpA belong to subgroup II of the α-ketoglutarate-dependent dioxygenases and share the specific motif HXDX24TX131HX10R. Amino acids His-111, Asp-113, and His-270 and amino acids His-102, Asp-104, and His 257 comprise the 2-His-1-carboxylate facial triads and were predicted to be involved in iron binding in RdpA and SdpA, respectively. RdpA exclusively transformed the (R) enantiomers of mecoprop [2-(4-chloro-2-methylphenoxy)propanoic acid] and dichlorprop [2-(2,4-dichlorophenoxy)propanoic acid], whereas SdpA was specific for the (S) enantiomers. The apparent Km values were 99 μM for (R)-mecoprop, 164 μM for (R)-dichlorprop, and 3 μM for α-ketoglutarate for RdpA and 132 μM for (S)-mecoprop, 495 μM for (S)-dichlorprop, and 20 μM for α-ketoglutarate for SdpA. Both enzymes had high apparent Km values for oxygen; these values were 159 μM for SdpA and >230 μM for RdpA, whose activity was linearly dependent on oxygen at the concentration range measured. Both enzymes had narrow cosubstrate specificity; only 2-oxoadipate was able to replace α-ketoglutarate, and the rates were substantially diminished. Ferrous iron was necessary for activity of the enzymes, and other divalent cations could not replace it. Although the results of growth experiments suggest that strain MH harbors a specific 2,4-dichlorophenoxyacetic acid-converting enzyme, tfdA-, tfdAα-, or cadAB-like genes were not discovered in a screening analysis in which heterologous hybridization and PCR were used.  相似文献   

10.
The first step of amino acid degradation in lactococci is a transamination, which requires an α-keto acid as the amino group acceptor. We have previously shown that the level of available α-keto acid in semihard cheese is the first limiting factor for conversion of amino acids to aroma compounds, since aroma formation is greatly enhanced by adding α-ketoglutarate to cheese curd. In this study we introduced a heterologous catabolic glutamate dehydrogenase (GDH) gene into Lactococcus lactis so that this organism could produce α-ketoglutarate from glutamate, which is present at high levels in cheese. Then we evaluated the impact of GDH activity on amino acid conversion in in vitro tests and in a cheese model by using radiolabeled amino acids as tracers. The GDH-producing lactococcal strain degraded amino acids without added α-ketoglutarate to the same extent that the wild-type strain degraded amino acids with added α-ketoglutarate. Interestingly, the GDH-producing lactococcal strain produced a higher proportion of carboxylic acids, which are major aroma compounds. Our results demonstrated that a GDH-producing lactococcal strain could be used instead of adding α-ketoglutarate to improve aroma development in cheese.  相似文献   

11.
2,4-Dichlorophenoxyacetic acid (2,4-D)/alpha-ketoglutarate (alphaKG) dioxygenase, TfdA, couples the oxidative decarboxylation of alphaKG to the oxidation of the herbicide 2,4-D using a mononuclear non-heme Fe(II) active site. The intrinsic tryptophan fluorescence associated with the four Trp residues in TfdA allows for the use of fluorescence spectroscopy to monitor the binding of iron and alphaKG to the enzyme. The fluorescence spectrum of TfdA is quenched by 50-85% upon addition of Fe(II) or alphaKG, allowing determination of their binding affinities (K(d)=7.45+/-0.61 and 3.35+/-0.35 microM, respectively). Cu, Zn, Mn, Co, Mg, and Ca dictations also quench the TfdA fluorescence with affinities similar to that of Fe(II), whereas monovalent cations such as Na, K, and Li do not. H114A and D116A mutant forms of TfdA, lacking either a histidine or aspartate metallocenter ligand, exhibit weaker affinity for both Fe(II) and alphaKG based on the fluorescence changes. Trp256 is predicted to lie within 5 A of the metal and alphaKG binding sites; however, its substitution by Phe or Leu has negligible effects on the Fe(II)- and alphaKG-dependent fluorescence quenching. Because Trp195 is predicted to be quite distant ( approximately 15 A) from the active site, we conclude that some combination of Trp113 and Trp248 serves as the reporter that senses metal and cofactor binding to TfdA.  相似文献   

12.
Unlike other lactic acid bacteria, Lactococcus lactis subsp. lactis NCDO 2118 was able to grow in a medium lacking glutamate and the amino acids of the glutamate family. Growth in such a medium proceeded after a lag phase of about 2 days and with a reduced growth rate (0.11 h−1) compared to that in the reference medium containing glutamate (0.16 h−1). The enzymatic studies showed that a phosphoenolpyruvate carboxylase activity was present, while the malic enzyme and the enzymes of the glyoxylic shunt were not detected. As in most anaerobic bacteria, no α-ketoglutarate dehydrogenase activity could be detected, and the citric acid cycle was restricted to a reductive pathway leading to succinate formation and an oxidative branch enabling the synthesis of α-ketoglutarate. The metabolic bottleneck responsible for the limited growth rate was located in this latter pathway. As regards the synthesis of glutamate from α-ketoglutarate, no glutamate dehydrogenase was detected. While the glutamate synthase-glutamine synthetase system was detected at a low level, high transaminase activity was measured. The conversion of α-ketoglutarate to glutamate by the transaminase, the reverse of the normal physiological direction, operated with different amino acids as nitrogen donor. All of the enzymes assayed were shown to be constitutive.  相似文献   

13.
Human platelet-type 12-lipoxygenase (12-LOX) has recently been shown to play an important role in regulation of human platelet function by reacting with arachidonic acid (AA). However, a number of other fatty acids are present on the platelet surface that, when cleaved from the phospholipid, can be oxidized by 12-LOX. We sought to characterize the substrate specificity of 12-LOX against six essential fatty acids: AA, dihomo-γ-linolenic acid (DGLA), eicosapentaenoic acid (EPA), α-linolenic acid (ALA), eicosadienoic acid (EDA), and linoleic acid (LA). Three fatty acids were comparable substrates (AA, DGLA, and EPA), one was 5-fold slower (ALA), and two showed no reactivity with 12-LOX (EDA and LA). The bioactive lipid products resulting from 12-LOX oxidation of DGLA, 12-(S)-hydroperoxy-8Z,10E,14Z-eicosatrienoic acid [12(S)-HPETrE], and its reduced product, 12(S)-HETrE, resulted in significant attenuation of agonist-mediated platelet aggregation, granule secretion, αIIbβ3 activation, Rap1 activation, and clot retraction. Treatment with DGLA similarly inhibited PAR1-mediated platelet activation as well as platelet clot retraction. These observations are in surprising contrast to our recent work showing 12(S)-HETE is a prothrombotic bioactive lipid and support our hypothesis that the overall effect of 12-LOX oxidation of fatty acids in the platelet is dependent on the fatty acid substrates available at the platelet membrane.  相似文献   

14.
NADH oxidation by plasma membrane vesicles purified from hypocotyls of etiolated soybean seedlings by two-phase partition was stimulated 2- to 3-fold by auxins, indole-3-acetic acid, 2,4-dichlorophenoxy acetic acid (2,4-D), and α-naphthaleneacetic acid. The stimulation was concentration dependent in the presence or absence of detergent with a maximum for 2,4-D at 1 micromolar. The NADH oxidation activity was solubilized with the zwitterionic detergent CHAPS and purified by ion exchange chromatography and gel filtration approximately 2000-fold over the total homogenate. Both the partially purified fraction and an active band from nondenaturing gel electrophoresis revealed the same three bands when analyzed by denaturing gel electrophoresis. When obtained from plasma membrane vesicles from the region of rapid cell elongation, the NADH oxidase complex retained auxin responsiveness throughout purification (3- to 5-fold stimulation by 1 micromolar 2,4-D).  相似文献   

15.
Wheat (Triticum aestivum L. cv Etoile de Choisy) microsomes catalyzed the cytochrome P-450-dependent oxidation of the herbicide diclofop to three hydroxy-diclofop isomers. Hydroxylation was predominant at carbon 4, with migration of chlorine to carbon 5 (67%) and carbon 3 (25%). The 2,4-dichloro-5-hydroxy isomer was identified as a minor reaction product (8%). Substrate-specificity studies showed that the activity was not inhibited or was weakly inhibited by a range of xenobiotic or physiological cytochrome P-450 substrates, with the exception of lauric acid. Wheat microsomes also catalyze the metabolism of the herbicides chlorsulfuron, chlortoluron, and 2,4-dichlorophenoxyacetic acid and of the model substrate ethoxycoumarin, as well as the hydroxylation of the endogenous substrates cinnamic and lauric acids. Treatments of wheat seedlings with phenobarbital or the safener naphthalic acid anhydride enhanced the cytochrome P-450 content of the microsomes and all related activities except that of cinnamic acid 4-hydroxylase, which was reduced. The stimulation patterns of diclofop aryl hydroxylase and lauric acid hydroxylase were similar, in contrast with the other activities tested. Lauric acid inhibited competitively (Ki = 9 μm) the oxidation of diclofop and reciprocally. The similarity of diclofop aryl hydroxylase and lauric acid hydroxylase was further investigated by alternative substrate kinetics, autocatalytic inactivation, and computer-aided molecular modelisation studies, and the results suggest that both reactions are catalyzed by the same cytochrome P-450 isozyme.  相似文献   

16.
Moreau F  Romani R 《Plant physiology》1982,70(5):1385-1390
After preparation on self-generated Percoll gradients, avocado (Persea americana Mill, var. Fuerte and Hass) mitochondria retain a high proportion of cyanide-insensitive respiration, especially with α-ketoglutarate and malate as substrates. Whereas α-ketoglutarate oxidation remains unchanged, the rate of malate oxidation increases as ripening advances through the climacteric. An enhancement of mitochondrial malic enzyme activity, measured by the accumulation of pyruvate, closely parallels the increase of malate oxidation. The capacity for cyanide-insensitive respiration is also considerably enhanced while respiratory control decreases (from 3.3 to 1.7), leading to high state 4 rates.

Both malate dehydrogenase and malic enzyme are functional in state 3, but malic enzyme appears to predominate before the addition of ADP and after its depletion. In the presence of cyanide, a membrane potential is generated when the alterntive pathway is operating. Cyanide-insensitive malate oxidation can be either coupled to the first phosphorylation site, sensitive to rotenone, or by-pass this site. In the absence of phosphate acceptor, malate oxidation is mainly carried out via malic enzyme and the alternative pathway. Experimental modification of the external mitochondrial environment in vitro (pH, NAD+, glutamade) results in changes in malate dehydrogenase and malic enzyme activities, which also modify cyanide resistance. It appears that a functional connection exists between malic enzyme and the alternative pathway via a rotenone-insensitive NADH dehydrogenase and that this pathway is responsible, in part, for nonphosphorylating respiratory activity during the climacteric.

  相似文献   

17.
The effect of octanoate on the oxidative decarboxylation of 14C-labeled amino acids has been studied in perfused hindquarter and liver of rat. Regulation of the branched-chain α-keto acid dehydrogenase has been further studied with α-[14C-1]ketoisovalerate in isolated rat muscle and liver mitochondria. (1) Octanoate has a stimulatory effect on the oxidation of branched-chain amino acids in perfused hindquarter. The oxidative decarboxylation of other amino acids are inhibited. Octanoate inhibits the oxidative decarboxylation of all amino acids in perfused liver. (2) The oxidation of valine is stimulated by octanoate and hexanoate also in isolated muscle mitochondria. The stimulatory effect is probably related to activation of the fatty acids since acyl-carnitines inhibit the oxidation. (3) The oxidation of α-ketoisovalerate in mitochondria is inhibited by competing substrates (pyruvate, α-ketoglutarate and succinate). This inhibition is counteracted by octanoate and ADP. (4) Low concentrations (1–5 μM) of 2,4-dinitrophenol (DNP) activates wheras higher concentrations inactivates the branched-chain α-keto acid dehydrogenase in intact but not in solubilized muscle mitochondria. The inactivation is counteracted by ATP, but is increased by octanoate. (5) The observations seem to suggest that the activation (like the inactivation) of branched-chain α-keto acid dehydrogenase in skeletal muscle is dependent on the mitochondrial energy state which therefore may regulate both activation and inactivation of the dehydrogenase.  相似文献   

18.
12-Ketochenodeoxycholic acid, an essential intermediate in the synthesis of chenodeoxycholic acid, has been enzymatically prepared from cholic acid. The specific oxidation of the 12α-hydroxyl group of cholic acid with NADP+ was catalysed by 12α-hydroxysteroid dehydrogenase (12α-hydroxysteroid: NAD+ oxidoreductase, EC 1.1.1.176), and the regeneration of NADP+ was obtained through the glutamate dehydrogenase (l-glutamate:NADP+ oxidoreductase, EC 1.4.1.4) catalysed reduction of α-ketoglutarate. The two enzymes were immobilized onto Sepharose CL-4B activated with tresyl chloride. The coimmobilized enzymes showed a cycling efficiency for the coenzyme similar to that of the free enzymes. High concentrations of cholic acid (up to 4%, w/v) were completely and specifically transformed into the 12-keto derivative using amounts of cofactor about 1600 times lower on a molar basis. The immobilized enzymes maintained 70% of the initial activity after 2 months of continuous use.  相似文献   

19.
Thirty-four thermophilic Bacillus sp. strains were isolated from decayed wood bark and a hot spring water sample based on their ability to degrade vanillic acid under thermophilic conditions. It was found that these bacteria were able to degrade a wide range of aromatic acids such as cinnamic, 4-coumaric, 3-phenylpropionic, 3-(p-hydroxyphenyl)propionic, ferulic, benzoic, and 4-hydroxybenzoic acids. The metabolic pathways for the degradation of these aromatic acids at 60°C were examined by using one of the isolates, strain B1. Benzoic and 4-hydroxybenzoic acids were detected as breakdown products from cinnamic and 4-coumaric acids, respectively. The β-oxidative mechanism was proposed to be responsible for these conversions. The degradation of benzoic and 4-hydroxybenzoic acids was determined to proceed through catechol and gentisic acid, respectively, for their ring fission. It is likely that a non-β-oxidative mechanism is the case in the ferulic acid catabolism, which involved 4-hydroxy-3-methoxyphenyl-β-hydroxypropionic acid, vanillin, and vanillic acid as the intermediates. Other strains examined, which are V0, D1, E1, G2, ZI3, and H4, were found to have the same pathways as those of strain B1, except that strains V0, D1, and H4 had the ability to transform 3-hydroxybenzoic acid to gentisic acid, which strain B1 could not do.  相似文献   

20.
The effect of cyanide and rotenone on malate (pH 6.8), malate plus glutamate (pH 7.8), citrate, α-ketoglutarate, and succinate oxidation by cauliflower (Brassica oleracea L.) bud, sweet potato (Ipomoea batatis L.) tuber, and spinach (Spinacia oleracea and Kalanchoë daigremontiana leaf mitochondria was investigated. Cyanide inhibited all substrates equally with the exception of malate plus glutamate; in this case, inhibition of O2 uptake was more severe due to an effect of cyanide on aspartate aminotransferase. Azide and antimycin A gave similar inhibitions with all substrates. Subsequent addition of NAD had no effect with any substrate. Providing that oxalacetate accumulation was prevented, rotenone inhibited all NAD-linked substrates equally and caused ADP:O ratios to decrease by one-third. Addition of succinate to mitochondria oxidizing malate stimulated oxygen uptake, but adding citrate and α-ketoglutarate did not. These results indicate that there is no direct link between malic enzyme and the rotenone- and cyanide-resistant respiratory pathways, and that there is no need to postulate separate compartmentation of malic enzyme and the other NAD-linked enzymes in the matrix.  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