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.
Ribulose 1,5-bisphosphate carboxylase/oxygenase (RubisCO) catalyzes the assimilation of atmospheric CO2 into organic matter and is thus central to the existence of life on earth. The beginning of the 2000s was marked by the discovery of a new family of proteins, the RubisCO-like proteins (RLPs), which are structural homologs of RubisCO. RLPs are unable to catalyze CO2 fixation. The RLPs from Chlorobaculum tepidum, Bacillus subtilis, Geobacillus kaustophilus, and Microcystis aeruginosa have been shown to participate in sulfur metabolism. Whereas the precise function of C. tepidum RLP is unknown, the B. subtilis, G. kaustophilus, and M. aeruginosa RLPs function as tautomerases/enolases in a methionine salvage pathway (MSP). Here, we show that the form II RubisCO enzyme from the nonsulfur purple bacterium Rhodospirillum rubrum is also able to function as an enolase in vivo as part of an MSP, but only under anaerobic conditions. However, unlike B. subtilis RLP, R. rubrum RLP does not catalyze the enolization of 2,3-diketo-5-methylthiopentyl-1-phosphate. Instead, under aerobic growth conditions, R. rubrum RLP employs another intermediate of the MSP, 5-methylthioribulose-1-phosphate, as a substrate, resulting in the formation of different products. To further determine the interrelationship between RubisCOs and RLPs (and the potential integration of cellular carbon and sulfur metabolism), the functional roles of both RubisCO and RLP have been examined in vivo via the use of specific knockout strains and complementation studies of R. rubrum. The presence of functional, yet separate, MSPs in R. rubrum under both aerobic (chemoheterotrophic) and anaerobic (photoheterotrophic) growth conditions has not been observed previously in any organism. Moreover, the aerobic and anaerobic sulfur salvage pathways appear to be differentially controlled, with novel and previously undescribed steps apparent for sulfur salvage in this organism.Ribulose 1,5-bisphosphate (RuBP) carboxylase/oxygenase (RubisCO) is the key enzyme of the Calvin-Benson-Bassham (CBB) reductive pentose phosphate pathway. This enzyme catalyzes the primary CO2 fixation reaction and is found in diverse organisms, including plants, most photosynthetic and chemoautotrophic microorganisms, and many archaea (25). On the basis of amino acid sequence similarities, the RubisCO family of proteins has been classified into four groups, i.e., form I, form II, form III, and form IV (Fig. (Fig.1).1). The enzymes classified under forms I, II, and III are all able to catalyze the RubisCO reaction, i.e., carboxylation/oxygenation of RuBP. The most recently discovered group of enzymes in the RubisCO family are the form IV or RubisCO-like proteins (RLPs). These proteins have thus far been identified in proteobacteria, cyanobacteria, archaea, and algae (2, 4, 8, 11, 12, 21, 25, 26). RLPs have been further divided into six different subgroups based on sequence similarities within the group: IV-Photo, IV-Nonphoto, IV-YkrW, IV-DeepYrkW, IV-GOS (Global Ocean Sequencing), and IV-AMC (Acid Mine Consortium) (25, 26). Despite sharing a level of sequence similarity with the bonafide RubisCOs, the RLPs are unable to carry out CO2/O2 fixation because their sequences contain dissimilar residues at positions analogous to RubisCO''s active-site residues (25). The structures of the Geobacillus kaustophilus and Chlorobaculum tepidum RLPs have now been solved, and there are indeed differences between the tertiary structures of these two proteins and the bonafide RubisCO enzymes (14, 17, 25). Moreover, distinct patterns of active-site residue identities among the different clades of the RLP lineage suggest that these subgroups of RLPs are likely to utilize different substrates and perform dissimilar reactions (23, 25, 26).Open in a separate windowFIG. 1.Summary of the different classes of RubisCO found in nature so far (25). Forms I, II, and III catalyze bonafide CO2/O2 fixation reactions by using RuBP as the substrate. Form IV RubisCO (RLP) does not catalyze RuBP-dependent CO2/O2 fixation and is divided into six known clades (25), with only representatives of the type IV-YkrW and IV-DeepYkrW subgroups shown to catalyze defined, yet distinct, reactions (Fig. (Fig.22).Previous studies performed with the Chlorobaculum tepidum RLP (of the IV-Photo group) gave the first indication that the RLPs may be involved in some aspect of sulfur metabolism (12, 13). This was later substantiated when the precise function was established for the Bacillus subtilis (2), Microcystis aeruginosa (4), and Geobacillus kaustophilus (14) RLPs of the IV-YkrW group. All three proteins catalyze a tautomerase/enolase reaction of a methionine salvage pathway (MSP) in which the substrate 2,3-diketo-5-methylthiopentanyl-1-phosphate (DK-MTP 1P) is converted to 2-hydroxy-3-keto-5-thiomethylpent-1-ene 1-phosphate (HK-MTP 1P) (Fig. (Fig.2).2). This reaction is very reminiscent of the enolization of RuBP catalyzed by RubisCO. Moreover, form II RubisCO from Rhodospirillum rubrum was shown to complement an RLP mutant strain of B. subtilis, with the ability to catalyze the identical tautomerase/enolase reaction (2). Interestingly, in addition to the presence of a form II RubisCO gene (cbbM), the genome of R. rubrum also encodes an RLP that clusters with the IV-DeepYkrW group (25). The function of this protein was recently determined, and it was shown to catalyze a distinct reaction that uses 5-methylthioribulose-1-phosphate as the substrate (15). Via an unprecedented 1,3-proton transfer, with two successive 1,2-proton transfers from its substrate, R. rubrum RLP catalyzes the formation of two products, i.e., 1-thiomethyl-d-xylulose-5-phosphate and 1-thiomethyl-d-ribulose-5-phosphate, at a 3:1 ratio (15) (Fig. (Fig.2).2). The novel reaction catalyzed by this RLP suggests that R. rubrum likely uses a different pathway to salvage sulfur compounds.Open in a separate windowFIG. 2.Distinct reactions catalyzed by type IV-YkrW (A) and type IV-DeepYkrW (B) classes of form IV RubisCO/RLP, exemplified by the proteins from B. subtilis and R. rubrum, respectively.The presence of an RLP-encoding gene triggered the search for additional genes in the R. rubrum genome that might be homologs of known enzymes that participate in a conventional MSP. Several genes were indeed identified to encode homologs of MSP enzymes. However, to this point there is no experimental evidence for the existence of a functional MSP (21) in R. rubrum. Thus, in this study, we sought to determine the role of the RLP and RubisCO protein in sulfur salvage since each protein catalyzes different reactions and RubisCO is known to be synthesized only under anaerobic conditions (6, 7). Moreover, it is well appreciated that R. rubrum possesses a versatile metabolic capacity and is able to grow under both anaerobic and aerobic conditions, using a variety of carbon sources. The involvement of RLP and RubisCO in sulfur salvage was thus determined and found to be associated with aerobic and anaerobic metabolism, respectively.  相似文献   

2.
A significant fraction of the proteome of Chlorobium tepidum is altered in a mutant strain of the green sulfur bacterium C. tepidum (Omega::RLP) lacking the Rubisco-like protein (RLP). Additionally, a number of stress proteins display altered abundance or migration in strain Omega::RLP, including a thioredoxin, a putative Hsp20 family chaperonin, and GroEL. Changes in protein abundance are closely correlated to mRNA abundance in the case of two other stress proteins, a thiol-specific antioxidant protein homolog (Tsa/AhpC) and an iron only superoxide dismutase (Fe-SOD). Strain Omega::RLP is more resistant to hydrogen peroxide exposure than strain WT2321, providing evidence that the stress proteins are functional. Strain Omega::RLP is also defective in thiosulfate oxidation, but is able to oxidize sulfide as well as the wild-type strain. Based on studies with periplasm-enriched extracts of strain Omega::RLP, the loss of thiosulfate oxidation capability correlates with undetectable levels of the Sox Y protein, a component of the predicted thiosulfate oxidation complex. These results provide further indications that sulfur oxidation capacity and the response to stress are linked in C. tepidum, with the RLP playing a major role.  相似文献   

3.
Ribulose-1,5-bisphosphate carboxylase/oxygenase (RuBisCO) is the key enzyme in the fixation of CO(2) in the Calvin cycle of plants. Many genome projects have revealed that bacteria, including Bacillus subtilis, possess genes for proteins that are similar to the large subunit of RuBisCO. These RuBisCO homologues are called RuBisCO-like proteins (RLPs) because they are not able to catalyse the carboxylase or the oxygenase reactions that are catalysed by photosynthetic RuBisCO. It has been demonstrated that B. subtilis RLP catalyses the 2,3-diketo-5-methylthiopentyl-1-phosphate (DK-MTP-1-P) enolase reaction in the methionine salvage pathway. The structure of DK-MTP-1-P is very similar to that of ribulose-1,5-bisphosphate (RuBP) and the enolase reaction is a part of the reaction catalysed by photosynthetic RuBisCO. In this review, functional and evolutionary relationships between B. subtilis RLP of the methionine salvage pathway, other RLPs, and photosynthetic RuBisCO are discussed. In addition, the fundamental question, 'How has RuBisCO evolved?' is also considered, and evidence is presented that RuBisCOs evolved from RLPs.  相似文献   

4.
About 30 years have now passed since it was discovered that microbes synthesize RubisCO molecules that differ from the typical plant paradigm. RubisCOs of forms I, II, and III catalyze CO(2) fixation reactions, albeit for potentially different physiological purposes, while the RubisCO-like protein (RLP) (form IV RubisCO) has evolved, thus far at least, to catalyze reactions that are important for sulfur metabolism. RubisCO is the major global CO(2) fixation catalyst, and RLP is a somewhat related protein, exemplified by the fact that some of the latter proteins, along with RubisCO, catalyze similar enolization reactions as a part of their respective catalytic mechanisms. RLP in some organisms catalyzes a key reaction of a methionine salvage pathway, while in green sulfur bacteria, RLP plays a role in oxidative thiosulfate metabolism. In many organisms, the function of RLP is unknown. Indeed, there now appear to be at least six different clades of RLP molecules found in nature. Consideration of the many RubisCO (forms I, II, and III) and RLP (form IV) sequences in the database has subsequently led to a coherent picture of how these proteins may have evolved, with a form III RubisCO arising from the Methanomicrobia as the most likely ultimate source of all RubisCO and RLP lineages. In addition, structure-function analyses of RLP and RubisCO have provided information as to how the active sites of these proteins have evolved for their specific functions.  相似文献   

5.
D M Schmidt  B K Hubbard  J A Gerlt 《Biochemistry》2001,40(51):15707-15715
The members of the mechanistically diverse enolase superfamily catalyze different overall reactions by using a common catalytic strategy and structural scaffold. In the muconate lactonizing enzyme (MLE) subgroup of the superfamily, abstraction of a proton adjacent to a carboxylate group initiates reactions, including cycloisomerization (MLE), dehydration [o-succinylbenzoate synthase (OSBS)], and 1,1-proton transfer (catalyzed by an OSBS that also catalyzes a promiscuous N-acylamino acid racemase reaction). The realization that a member of the MLE subgroup could catalyze a 1,1-proton transfer reaction, albeit poorly, led to a search for other enzymes which might catalyze a 1,1-proton transfer as their physiological reaction. YcjG from Escherichia coli and YkfB from Bacillus subtilis, proteins of previously unknown function, were discovered to be L-Ala-D/L-Glu epimerases, although they also catalyze the epimerization of other dipeptides. The values of k(cat)/K(M) for L-Ala-D/L-Glu for both proteins are approximately 10(4) M(-1) s(-1). The genomic context and the substrate specificity of both YcjG and YkfB suggest roles in the metabolism of the murein peptide, of which L-Ala-D-Glu is a component. Homologues possessing L-Ala-D/L-Glu epimerase activity have been identified in at least two other organisms.  相似文献   

6.
AIMS: To develop alternative approaches for medical and environmental control of pathogenic Acanthamoeba spp. by means of photodynamic treatment with a tetracationic Zn(II)-phthalocyanine (RLP068). METHODS AND RESULTS: Incubation of cyst cultures with RLP068 for 1 h caused an accumulation of readily detectable concentrations of the phthalocyanine, even at doses as low as 0.5 micromol l(-1). RLP068 exhibited no dark toxicity towards cysts up to 5 micromol l(-1) concentration. A decrease of c. 50% in cyst survival in comparison with controls was measured upon incubation of the cysts with 0.5 micromol l(-1) RLP068, followed by exposure to light (600-700 nm) for 20 min at a fluence rate of 50 mW cm(-2) (60 J cm(-2)). After incubation with 3 and 5 micromol l(-1) RLP068 and irradiation, the cysts lost their excystment ability as early as day 5 and up to day 10, and were clearly damaged when observed under an interference contrast microscope. CONCLUSIONS: These data indicate the promising use of RLP068 in phototreatment of diseases caused by pathogenic amoebae and in initial disinfection of wastewaters. SIGNIFICANCE AND IMPACT OF THE STUDY: Rapid and extensive photodamage may be induced in the highly resistant cystic stages by means of 600- to 700-nm light sources.  相似文献   

7.
The nuclear spin-spin and spin-lattice relaxation times of the C epsilon 1-proton of His-35 and the C delta 2-proton of His-46 of reduced Pseudomonas aeruginosa azurin have been determined at 298 and 320 K and at pH 4.5 and 9.0 at various concentrations of total azurin and in the presence of varying amounts of oxidized azurin. The relaxation times appear strongly influenced by the electron self-exchange reaction between oxidized and reduced protein. The T1 data of the His-35 proton have been analyzed according to the "fast-exchange limit," while the "slow-exchange limit" appears to obtain for the T2 data of the His-46 proton. Analysis of the proton relaxation data yields values of the electron self-exchange rate constants of (9.6 +/- 0.7) X 10(5) M-1 S-1 (pH 4.5) and (7.0 +/- 1.3) X 10(5) M-1 S-1 (pH 9.0) at 298 K. The dipolar correlation time amounts to 1-2.5 ns in the temperature range of 298-320 K. A Fermi-contact interaction of about 100 mG for the C delta 2-proton of His-46 is compatible with the experimental observations. The pH-induced conformational changes lead to variations on the order of about 1 A in the distance from the copper to the His-35 protons. The data implicate the "hydrophobic patch" around His-117 as the site of electron transfer in the self-exchange reaction of the azurin.  相似文献   

8.
The decomposition of l-tyrosine and its α-deuterated analogue under the action of extracts from Escherichia intermedia A-21 with high tyrosine phenol-lyase [l-tyrosine phenol-lyase (deaminating), EC 4.1.99.2] activity has been studied. The mass spectrometric data for samples of phenol produced by decomposition of deutero-l-tyrosine in water and D2O-water (10:1) mixture, and by decomposition of normal l-tyrosine in D2O-water (10:1), show that the process is accompanied by the intramolecular transfer of D or H to the leaving phenol group. The degree of transfer is 7–10%. Thus, the abstraction of α-proton and the subsequent protonation of the aromatic ring are accomplished by the same functional group of the enzyme. This is indicative of the cis-orientation of α-proton and the phenol fragment in relation to the plane of Schiff's base of α-aminoacrylate with pyridoxal phosphate during α,β-elimination. The isotope effect of the studied enzymic reaction is ≈3, which allows us to consider the α-proton abstraction as the limiting stage of the process.  相似文献   

9.
Bacterial phosphopentomutases (PPMs) are alkaline phosphatase superfamily members that interconvert α-D-ribose 5-phosphate (ribose 5-phosphate) and α-D-ribose 1-phosphate (ribose 1-phosphate). We investigated the reaction mechanism of Bacillus cereus PPM using a combination of structural and biochemical studies. Four high resolution crystal structures of B. cereus PPM revealed the active site architecture, identified binding sites for the substrate ribose 5-phosphate and the activator α-D-glucose 1,6-bisphosphate (glucose 1,6-bisphosphate), and demonstrated that glucose 1,6-bisphosphate increased phosphorylation of the active site residue Thr-85. The phosphorylation of Thr-85 was confirmed by Western and mass spectroscopic analyses. Biochemical assays identified Mn(2+)-dependent enzyme turnover and demonstrated that glucose 1,6-bisphosphate treatment increases enzyme activity. These results suggest that protein phosphorylation activates the enzyme, which supports an intermolecular transferase mechanism. We confirmed intermolecular phosphoryl transfer using an isotope relay assay in which PPM reactions containing mixtures of ribose 5-[(18)O(3)]phosphate and [U-(13)C(5)]ribose 5-phosphate were analyzed by mass spectrometry. This intermolecular phosphoryl transfer is seemingly counter to what is anticipated from phosphomutases employing a general alkaline phosphatase reaction mechanism, which are reported to catalyze intramolecular phosphoryl transfer. However, the two mechanisms may be reconciled if substrate encounters the enzyme at a different point in the catalytic cycle.  相似文献   

10.
A putative long-chain mannitol-1-phosphate 5-dehydrogenase from Aspergillus fumigatus (AfM1PDH) was overexpressed in Escherichia coli to a level of about 50% of total intracellular protein. The purified recombinant protein was a approximately 40-kDa monomer in solution and displayed the predicted enzymatic function, catalyzing NAD(H)-dependent interconversion of d-mannitol 1-phosphate and d-fructose 6-phosphate with a specific reductase activity of 170 U/mg at pH 7.1 and 25 degrees C. NADP(H) showed a marginal activity. Hydrogen transfer from formate to d-fructose 6-phosphate, mediated by NAD(H) and catalyzed by a coupled enzyme system of purified Candida boidinii formate dehydrogenase and AfM1PDH, was used for the preparative synthesis of d-mannitol 1-phosphate or, by applying an analogous procedure using deuterio formate, the 5-[2H] derivative thereof. Following the precipitation of d-mannitol 1-phosphate as barium salt, pure product (>95% by HPLC and NMR) was obtained in isolated yields of about 90%, based on 200 mM of d-fructose 6-phosphate employed in the reaction. In situ proton NMR studies of enzymatic oxidation of d-5-[2H]-mannitol 1-phosphate demonstrated that AfM1PDH was stereospecific for transferring the deuterium to NAD+, producing (4S)-[2H]-NADH. Comparison of maximum initial rates for NAD+-dependent oxidation of protio and deuterio forms of D-mannitol 1-phosphate at pH 7.1 and 25 degrees C revealed a primary kinetic isotope effect of 2.9+/-0.2, suggesting that the hydride transfer was strongly rate-determining for the overall enzymatic reaction under these conditions.  相似文献   

11.
Pulse radiolysis studies of the hydrogen atom transfer ("repair") reaction from the sulfhydryl-containing (RSH) compound dithiothreitol (DTT) to the DNA sugar deoxyribose and to several related sugars have been undertaken. The H transfer reaction is measured by monitoring the transient absorbance of the radical-anion RSSR-. The H atom transfer reactions for some sugars were fitted by a single time exponential function, but other sugars exhibited both a fast and a slow component (approximately 10-fold difference in rates) to the reaction. The reaction rates for the slow stage of the reaction between DTT and the sugars ranged from 0.5 X 10(7) dm3 mole-1 sec-1 for ribose-5-phosphate to 9 X 10(7) dm3 mole-1 sec-1 for 2-deoxyglucose. The maximum extent of total repair varied from 60% for ribose-5-phosphate to 100% for 2-deoxyglucose. The rate of repair, the extent of repair, and the appearance of more than one component of repair seem to depend on several factors: The occurrence of radical-radical reactions in the system is indicated by the demonstration of a dose dependence of the reaction kinetics, and this affects the observed rate of formation of RSSR-. Sugars with a deoxy group on the 2-carbon atom seem to have enhanced rates and extents of repair and to exhibit both fast and slow components to the reaction. The presence of a phosphate group on the sugar causes a decrease in the rate and extent of repair. The biological relevance of the reactions studied herein is discussed and the rates obtained are compared with rates for repair of damage in certain radiobiological systems.  相似文献   

12.
1-Deoxy-D-xylulose-5-phosphate is converted into 2-C-methyl-D-erythritol-4-phosphate by the catalytic action of 1-deoxy-D-xylulose-5-phosphate reductoisomerase (Dxr protein) using NADPH as cofactor. The stereochemical features of this reaction were investigated in in vitro experiments with the recombinant Dxr protein of Escherichia coli using (4R)- or (4S)-[4-(2)H(1)]NADPH as coenzyme. The enzymatically formed 2-C-methyl-D-erythritol-4-phosphate was isolated and converted into 1,2:3,4-di-O-isopropylidene-2-C-methyl-D-erythritol; NMR spectroscopic investigation of this derivative indicated that only (4S)-[4-(2)H(1)]NADPH affords 2-C-methyl-D-erythritol-4-phosphate labelled exclusively in the H(Re) position of C-1. Stereospecific transfer of H(Si) from C-4 of the cofactor identifies the Dxr protein of E. coli as a class B dehydrogenase.  相似文献   

13.
Zolli M  Kobric DJ  Brown ED 《Biochemistry》2001,40(16):5041-5048
CDP-ribitol synthase is a bifunctional reductase and cytidylyltransferase that catalyzes the transformation of D-ribulose 5-phosphate, NADPH, and CTP to CDP-ribitol, a repeating unit present in the virulence-associated polysaccharide capsules of Haemophilus influenzae types a and b [Follens, A., et al. (1999) J. Bacteriol. 181, 2001]. In the work described here, we investigated the order of the reactions catalyzed by CDP-ribitol synthase and conducted experiments to resolve the question of substrate channeling in this bifunctional enzyme. It was determined that the synthase first catalyzed the reduction of D-ribulose 5-phosphate followed by cytidylyl transfer to D-ribitol 5-phosphate. Steady state kinetic measurements revealed a 650-fold kinetic preference for cytidylyl transfer to D-ribitol 5-phosphate over D-ribulose 5-phosphate. Rapid mixing studies indicated quick reduction of D-ribulose 5-phosphate with a lag in the cytidylyl transfer reaction, consistent with a requirement for the accumulation of K(m) quantities of D-ribitol 5-phosphate. Signature motifs in the C-terminal and N-terminal sequences of the enzyme (short chain dehydrogenase/reductase and nucleotidyltransferase motifs, respectively) were targeted with site-directed mutagenesis to generate variants that were impaired for only one of the two activities (K386A and R18A impaired for reduction and cytidylyl transfer, respectively). Release and free diffusion of the metabolic intermediate D-ribitol 5-phosphate was indicated by the finding that equimolar mixtures of K386A and R18A variants were efficient for bifunctional catalysis. Taken together, these findings suggest that bifunctional turnover occurs in distinct active sites of CDP-ribitol synthase with reduction of D-ribulose 5-phosphate and release and free diffusion of the metabolic intermediate D-ribitol 5-phosphate followed by cytidylyl transfer.  相似文献   

14.
In an earlier investigation (Shanmugam, K. T., Buchanan, B. B., and Arnon, D. I. (1972) Biochim. Biophys. Acta 256, 477-486) the extraction of ferredoxin from Rhodospirillum rubrum cells with the aid of a detergent (Triton X-100) and acetone revealed the existence of two types of ferredoxin (I and II) and led to the conclusion that both are membrane-bound. In the present investigation, ferredoxin and acid-labile sulfur analyses of photosynthetic membranes (chromatophores) and soluble protein extracts of the photosynthetic bacteria R. rubrum and Rhodopseudomonas spheroides showed that ferredoxins I and II are primarily components of the soluble protein fraction. After their removal, washed R. rubrum chromatophores were found to contain a considerable amount of tightly bound iron-sulfur protein(s), as evidenced by acid-labile sulfur and electron paramagnetic resonance analyses. Thus, like all other photosynthetic cells examined to date, R. rubrum cells contain both soluble ferredoxins and iron-sulfur proteins tightly bound to photosynthetic membranes. The molecular weights of ferredoxins I and II from photosynthetically grown R. rubrum cells were found to be 8,800 and 14,500, respectively. Using these molecular weights, the molar extinction coefficients at 390 nm for ferredoxins I and II were determined to be 30.3 and 17.2 mM-1 CM-1, respectively. Ferredoxin I contains 8 non-heme iron and 8 acid-labile sulfur atoms per molecule; ferredoxin II contains 4 non-heme iron and 4 acid-labile sulfur atoms per molecule. Ferredoxin I was found only in photosynthetically grown cells whereas ferredoxin II was present in both light- and dark-grown cells. Ferredoxin II from both light- and dark-grown cells has the same molecular weight (14,500) and absorption spectrum and has 4 iron and 4 acid-labile sulfur atoms per molecule. Low temperature electron paramagnetic resonance spectra of oxidized and photoreduced ferredoxins I and II from R. rubrum were recorded. The EPR spectrum of oxidized ferredoxin II exhibited a single resonance line at g = 2.012. Oxidized ferredoxin I, however, exhibited a spectrum that may arise from the superimposition of two resonance lines near g = 2.012. Photoreduced ferredoxin II displayed a rhombic EPR spectrum with a g value of 1.94. Photoreduced ferredoxin I exhibited a similar EPR spectrum at a temperature of 16 K, but when the temperature was lowered to 4.5 K the spectrum of ferredoxin I changed. This temperature-dependent spectrum may result from a weak spin-spin interaction between two iron-sulfur clusters. These results are consistent with the conclusion that R. rubrum ferredoxins I and II are, respectively, 8 iron/8 sulfur and 4 iron/4sulfur proteins.  相似文献   

15.
The stereochemistry of the hydrogen transfer to NAD catalyzed by ribitol dehydrogenase (ribitol:NAD 2-oxidoreductase, EC 1.1.1.56) from Klebsiella pneumoniae and D-mannitol-1-phosphate dehydrogenase (D-mannitol-1-phosphate:NAD 2-oxidoreductase, EC 1.1.1.17) from Escherichia coli was investigated. [4-3H]NAD was enzymatically reduced with nonlabelled ribitol in the presence of ribitol dehydrogenase and with nonlabelled D-mannitol 1-phosphate and D-mannitol 1-phosphate dehydrogenase, respectively. In both cases the [4-3H]-NADH produced was isolated and the chirality at the C-4 position determined. It was found that after the transfer of hydride, the label was in both reactions exclusively confined to the (4R) position of the newly formed [4-3H]NADH. In order to explain these results, the hydrogen transferred from the nonlabelled substrates to [4-3H]NAD must have entered the (4S) position of the nicotinamide ring. These data indicate for both investigated inducible dehydrogenases a classification as B or (S) type enzymes. Ribitol also can be dehydrogenated by the constitutive A-type L-iditol dehydrogenase (L-iditol:NAD 5-oxidoreductase, EC 1.1.1.14) from sheep liver. When L-iditol dehydrogenase utilizes ribitol as hydrogen donor, the same A-type classification for this oxidoreductase, as expected, holds true. For the first time, opposite chirality of hydrogen transfer to NAD in one organic reaction--ribitol + NAD = D-ribu + NADH + H--is observed when two different dehydrogenases, the inducible ribitol dehydrogenase from K. pneumoniae and the constitutive L-iditol dehydrogenase from sheep liver, are used as enzymes. This result contradicts the previous generalization that the chirality of hydrogen transfer to the coenzyme for the same reaction is independent of the source of the catalyzing enzyme.  相似文献   

16.
The stereospecificity of the reaction catalysed by the spinach chloroplast enzyme NADP-dependent glyceraldehyde-3-phosphate dehydrogenase (D-glyceraldehyde-3-phosphate: NADP+ oxidoreductase (phosphorylating), EC 1.2.1.13) with respect to the C4 nicotinamide hydrogen transfer was investigated. NADPH deuterated at the C4 HA position was synthesized using aldehyde dehydrogenase. 1H-NMR spectroscopy was used to examine the NADP+ product of the GPDH reaction for the presence or absence of the C4 deuterium atom. Chloroplast NADP-dependent glyceraldehyde-3-phosphate dehydrogenase retains the deuterium at the C4 HA position (removing the hydrogen atom), and is therefore a B (pro-S) specific dehydrogenase.  相似文献   

17.
Phosphoribulokinase (EC 2.7.1.19, ATP: d-ribulose-5-phosphate-1-phosphotransferase) resembles the NADPH-dependent glyceraldehyde-3-phosphate dehydrogenase (EC 1.2.1.13, d-glyceraldehyde-3-phosphate: NADPH(+) oxidoreductase [phosphorylating]) of chloroplasts in that the activation of both of these enzymes involves the dissociation of oligomers (apparently tetrameric forms) with low catalytic activity to give protomers which possess higher catalytic activity. Gel filtration on Sepharose 6B has shown that the molecular weights of the oligomer and active protomer of phosphoribulokinase are, respectively, about 6.8 x 10(5) and 1.7 x 10(5), whereas the corresponding values for glyceraldehyde-3-phosphate dehydrogenase are 8.2 x 10(5) and 2.2 x 10(5). Activation of both enzymes occurs in response to either ATP, dithiothreitol, or cholate while the glyceraldehyde-3-phosphate dehydrogenase is also activated by NADPH. Activation/dissociation of these enzymes may involve conformational changes resulting from nucleotide binding, the reduction of sulfur bridges, and the cholate induced loosening of hydrophobic interactions.  相似文献   

18.
Pattern recognition receptors (PRRs) sense ligands in pattern-triggered immunity (PTI). Plant PRRs include numerous receptor-like proteins (RLPs), but many RLPs remain functionally uncharacterized. Here, we examine an Arabidopsis thaliana RLP, RLP53, which positively regulates immune signaling. Our forward genetic screen for suppressors of enhanced disease resistance1 (edr1) identified a point mutation in RLP53 that fully suppresses disease resistance and mildew-induced cell death in edr1 mutants. The rlp53 mutants showed enhanced susceptibility to virulent pathogens, including fungi, oomycetes, and bacteria, indicating that RLP53 is important for plant immunity. The ectodomain of RLP53 contains leucine-rich repeat (LRR) motifs. RLP53 constitutively associates with the LRR receptor-like kinase SUPPRESSOR OF BRASSINOSTEROID INSENSITIVE1-ASSOCIATED KINASE (BAK1)-INTERACTING RECEPTOR KINASE1 (SOBIR1) and interacts with the co-receptor BAK1 in a pathogen-induced manner. The double mutation sobir1-12 bak1-5 suppresses edr1-mediated disease resistance, suggesting that EDR1 negatively regulates PTI modulated by the RLP53–SOBIR1–BAK1 complex. Moreover, the glycosylphosphatidylinositol (GPI)-anchored protein LORELEI-LIKE GPI-ANCHORED PROTEIN1 (LLG1) interacts with RLP53 and mediates RLP53 accumulation in the plasma membrane. We thus uncovered the role of a novel RLP and its associated immune complex in plant defense responses and revealed a potential new mechanism underlying regulation of RLP immune function by a GPI-anchored protein.  相似文献   

19.
Glycogen phosphorylase contains firmly bound pyridoxal 5′-phosphate (PLP), and catalyzes the reversible transfer of a glucosyl moiety between glucose-1-phosphate (G-1-P) and α-1,4-glucan. X-ray crystallographic studies revealed that PLP is located in a pocket where the phosphate group of PLP is pointed toward the G-1-P binding site. We have synthesized pyridoxal(5′)diphospho(1)-α-d-glucose, as a model compound for the phosphate-phosphate interaction between PLP and G-1-P, and reconstituted the enzyme with this compound. The resulting enzyme is catalytically inactive in itself, but, in the presence of glucan, the glycosyl moiety of this compound is transferred to the glucan forming a new α-1,4-glucosidic linkage along with the production of pyridoxal 5′-diphosphate. This glucosyltransfer is similar to the normal catalytic reaction in various aspects, although the rate is smaller in the order of three. AMP accelerates the transfer about 24 times compared with the reaction in its absence. We have more recently used pyridoxal(5′)triphospho(1)-α-D-glucose to reconstitute the enzyme. In the presence of glucan, the compound bound to enzyme is gradually degraded to pyridoxal 5′-triphosphate. This reaction is essentially dependent on AMP, and proceeds several times more slowly than the glucosyltransfer from the diphospho compound. These results provide evidence for the direct phosphate-phosphate interaction between the coenzyme and the substrate in the normal enzyme reaction, and seem to reflect a rather wide allowance in regard to this interaction.  相似文献   

20.
The isotopic discrimination, diastereotopic specificity and intramolecular hydrogen transfer characterizing the reaction catalyzed by phosphomannoisomerase are examined. During the monodirectional conversion of D-[2-3H]mannose 6-phosphate to D-fructose 6-phosphate and D-fructose 1,6-bisphosphate, the reaction velocity is one order of magnitude lower than with D-[U-14C]mannose 6-phosphate and little tritium (less than 6%) is transferred intramolecularly. Inorganic phosphate decreases the reaction velocity but favours the intramolecular transfer of tritium. Likewise, when D-[1-3H]fructose 6-phosphate prepared from D-[1-3H]glucose is exposed solely to phosphomannoisomerase, the generation of tritiated metabolites is virtually restricted to 3H2O and occurs at a much lower rate than the production of D-[U-14C]mannose 6-phosphate from D-[U-14C]fructose 6-phosphate. However, no 3H2O is formed when D-[1-3H]fructose 6-phosphate generated from D-[2-3H]glucose is exposed to phosphomannoisomerase, indicating that the diastereotopic specificity of the latter enzyme represents a mirror image of that of phosphoglucoisomerase. Advantage is taken of such a contrasting enzymic behaviour to assess the back-and-forth flow through the reaction catalyzed by phosphomannoisomerase in intact cells exposed to D-[1-3H]glucose, D-[5-3H]glucose or D-[6-3H]glucose. Relative to the rate of glycolysis, this back-and-forth flow amounted to approx. 4% in human erythrocytes and rat parotid cells, 9% in tumoral cells of the RINm5F line and 47% in rat pancreatic islets.  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