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 31 毫秒
1.
The pyruvate dehydrogenase complex of Escherichia coli contains two lipoic acid residues per dihydrolipoamide acetyltransferase chain, and these are known to engage in the part-reactions of the enzyme. The enzyme complex was treated with trypsin at pH 7.0, and a partly proteolysed complex was obtained that had lost almost 60% of its lipoic acid residues although it retained 80% of its pyruvate dehydrogenase-complex activity. When this complex was treated with N-ethylmaleimide in the presence of pyruvate and the absence of CoASH, the rate of modification of the remaining S-acetyldihydrolipoic acid residues was approximately equal to the accompanying rate of loss of enzymic activity. This is in contrast with the native pyruvate dehydrogenase complex, where under the same conditions modification proceeds appreciably faster than the loss of enzymic activity. The native pyruvate dehydrogenase complex was also treated with lipoamidase prepared from Streptococcus faecalis. The release of lipoic acid from the complex followed zero-order kinetics for most of the reaction, whereas the accompanying loss of pyruvate dehydrogenase-complex activity lagged substantially behind. These results eliminate a model for the enzyme mechanism in which specifically one of the two lipoic acid residues on each dihydrolipoamide acetyltransferase chain is essential for the reaction. They are consistent with a model in which the dihydrolipoamide acetyltransferase component contains more lipoic acid residues than are required to serve the pyruvate decarboxylase subunits under conditions of saturating substrates, enabling the function of an excised or inactivated lipoic acid residue to be taken over by another one. Unusual structural properties of the enzyme complex might permit this novel feature of the enzyme mechanism.  相似文献   

2.
African trypanosomes contain a membrane-bound enzyme capable of removing dimyristylglycerol from the membrane-attached form of the variable surface glycoprotein (mfVSG; Ferguson, M. A. J., K. Halder, and G. A. M. Cross, 1985, J. Biol Chem., 260:4963-4968). Although mfVSG phospholipase-C has been implicated in the removal of the VSG from the trypanosome surface (Cardoso de Almeida, M. L., and M. J. Turner, 1983, Nature (Lond.)., 302:349-352; Ferguson, M. A. J., K. Halder, and G. A. M. Cross, 1985, J. Biol Chem., 260:4963-4968), its precise function and subcellular location have not been determined. We have developed a procedure for the separation of the cell fractions and organelles of Trypanosoma brucei brucei (and other trypanosome species) by differential sucrose and isopycnic PercollR centrifugation. These fractions were tested for mfVSG phospholipase activity using Trypanosoma brucei mfVSG labeled with 3H-myristic acid as substrate. The highest enzyme-specific activity was associated with the flagella and evidence is presented to suggest that it is localized in the flagellar pocket. Some activity was also associated with the Golgi complex. These results suggest that the mfVSG phospholipase is localized primarily in the membrane of the flagella pocket and possibly other membrane organelles derived from and associated with this structure, and may be part of the VSG-membrane recycling system in African trypanosomes. The activity of mfVSG phospholipase amongst various trypanosome species was determined. We show that, in contrast to the bloodstream forms of Trypanosoma brucei, cultured procyclic Trypanosoma brucei and bloodstream Trypanosoma vivax had little or no mfVSG phospholipase activity. The activity found in bloodstream forms of Trypanosoma congolense was intermediate between Trypanosoma vivax and Trypanosoma brucei.  相似文献   

3.
When purified D-amino acid dehydrogenase [Olsiewski, P. J., Kaczorowski, G. J., & Walsh, C. T. (1980) J. Biol. Chem. 255, 4487] is incubated with right-side-out membrane vesicles from Escherichia coli, the enzyme binds to the membrane in a time- and concentration-dependent manner. As a result, the vesicles acquire the ability to oxidize D-alanine and catalyze D-alanine-dependent active transport. Similarly, incubation of D-amino acid dehydrogenase with inside-out vesicles results in binding of enzyme and D-alanine oxidase activity. Antibody inhibition studies indicate that the enzyme is bound exclusively to the inner cytoplasmic surface of the membrane in native vesicles (i.e., membrane vesicles prepared from cells induced for D-amino acid dehydrogenase). In contrast, similar studies with reconstituted vesicles demonstrate that enzyme binds to the surface exposed to the medium regardless of the orientation of the membrane. Thus, enzyme bound to right-side-out vesicles is located on the opposite side of the membrane from where it is normally found. Remarkably, in the presence of D-alanine, reconstituted right-side-out and inside-out vesicles generate electrochemical proton gradients of similar magnitude but opposite polarity, indicating that enzyme bound to either surface of the membrane is physiologically functional. The results suggest that vectorial proton translocation via the respiratory chain occurs at a point distal to the site where electrons enter the respiratory chain from the primary dehydrogenase, a conclusion that is inconsistent with the notion that the dehydrogenase forms part of a proton-translocating loop.  相似文献   

4.
In an earlier study (S. W. Jordan and J. E. Cronan, Jr., J. Biol. Chem. 272:17903-17906, 1997) we reported a new enzyme, lipoyl-[acyl carrier protein]-protein N-lipoyltransferase, in Escherichia coli and mitochondria that transfers lipoic acid from lipoyl-acyl carrier protein to the lipoyl domains of pyruvate dehydrogenase. It was also shown that E. coli lipB mutants lack this enzyme activity, a finding consistent with lipB being the gene that encoded the lipoyltransferase. However, it remained possible that lipB encoded a positive regulator required for lipoyltransferase expression or action. We now report genetic and biochemical evidence demonstrating that lipB encodes the lipoyltransferase. A lipB temperature-sensitive mutant was shown to produce a thermolabile lipoyltransferase and a tagged version of the lipB-encoded protein was purified to homogeneity and shown to catalyze the transfer of either lipoic acid or octanoic acid from their acyl carrier protein thioesters to the lipoyl domain of pyruvate dehydrogenase. In the course of these experiments the ATG initiation codon commonly assigned to lipB genes in genomic databases was shown to produce a nonfunctional E. coli LipB protein, whereas initiation at an upstream TTG codon gave a stable and enzymatically active protein. Prior genetic results (T. W. Morris, K. E. Reed, and J. E. Cronan, Jr., J. Bacteriol. 177:1-10, 1995) suggested that lipoate protein ligase (LplA) could also utilize (albeit poorly) acyl carrier protein substrates in addition to its normal substrates lipoic acid plus ATP. We have detected a very slow LplA-catalyzed transfer of lipoic acid and octanoic acid from their acyl carrier protein thioesters to the lipoyl domain of pyruvate dehydrogenase. A nonhydrolyzable lipoyl-AMP analogue was found to competitively inhibit both ACP-dependent and ATP-dependent reactions of LplA, suggesting that the same active site catalyzes two chemically diverse reactions.  相似文献   

5.
This study describes the enzymatic characterization of dihydrolipoamide dehydrogenase (DLDH) from Streptococcus pneumoniae and is the first characterization of a DLDH that carries its own substrate (a lipoic acid covalently attached to a lipoyl protein domain) within its own sequence. Full-length recombinant DLDH (rDLDH) was expressed and compared with enzyme expressed in the absence of lipoic acid (rDLDH(-LA)) or with enzyme lacking the first 112 amino acids constituting the lipoyl protein domain (rDLDH(-LIPOYL)). All three proteins contained 1 mol of FAD/mol of protein, had a higher activity for the conversion of NAD(+) to NADH than for the reaction in the reverse direction, and were unable to use NADP(+) and NADPH as substrates. The enzymes had similar substrate specificities, with the K(m) for NAD(+) being approximately 20 times higher than that for dihydrolipoamide. The kinetic pattern suggested a Ping Pong Bi Bi mechanism, which was verified by product inhibition studies. The protein expressed without lipoic acid was indistinguishable from the wild-type protein in all analyses. On the other hand, the protein without a lipoyl protein domain had a 2-3-fold higher turnover number, a lower K(I) for NADH, and a higher K(I) for lipoamide compared with the other two enzymes. The results suggest that the lipoyl protein domain (but not lipoic acid alone) plays a regulatory role in the enzymatic characteristics of pneumococcal DLDH.  相似文献   

6.
Avidin can form intermolecular cross-links between particles of the pyruvate dehydrogenase multienzyme complex from various sources. Avidin does this by binding to lipoic acid-containing regions of the dihydrolipoamide acetyltransferase polypeptide chains that comprise the structural core of the complex. It is inferred that the lipoyl domains of the acetyltransferase chain extend outwards from the interior of the enzyme particle, interdigitating between the subunits of the other two enzymes bound peripherally in the assembled structure, with the lipoyl-lysine residues capable of reaching to within at least 1-2 nm of the outer surface of the enzyme complex (diameter ca. 37 nm). The distribution of enzymic activities between different domains of the dihydrolipoamide acetyltransferase chain implies that considerable movement of the lipoyl domains is a feature of the catalytic activity of the enzyme complex. There is evidence that the lipoyl domain of the 2-oxo acid dehydrogenase complexes is similar in structure to a domain that binds the cofactor biotin, also in amide linkage with a specific lysine residue, in the biotin-dependent class of carboxylases.  相似文献   

7.
The overexpression of a subgene encoding a hybrid lipoyl domain of the dihydrolipoamide acetyltransferase component of the pyruvate dehydrogenase complex of Escherichia coli has previously been shown to result in the formation of lipoylated and unlipoylated products. Overexpression of the same subgene in a lipoic acid biosynthesis mutant growing under lipoate-deficient conditions has now been shown to produce domains modified by octanoylation as well as unmodified domains. It was concluded from the mass of a lipoyl-binding-site peptide that the modification involves N6-octanoylation of the lysine residue (Lys244) that is normally lipoylated, and this was confirmed by the trypsin-insensitivity of the corresponding Lys244-Ala-245 bond, and the absence of modification in a mutant domain in which Lys244 is replaced by Gln. This novel protein modification raises interesting questions concerning the pathway of lipoic acid biosynthesis and the mechanism of enzyme lipoylation.  相似文献   

8.
Dihydrolipoamide dehydrogenase (DLD) is a multifunctional protein well characterized as the E3 component of the pyruvate dehydrogenase and α-ketoglutarate dehydrogenase complexes. Previously, conditions predicted to destabilize the DLD dimer revealed that DLD could also function as a diaphorase and serine protease. However, the relevance of these cryptic activities remained undefined. We analyzed human DLD mutations linked to strikingly different clinical phenotypes, including E340K, D444V, R447G, and R460G in the dimer interface domain that are responsible for severe multisystem disorders of infancy and G194C in the NAD(+)-binding domain that is typically associated with milder presentations. In vitro, all of these mutations decreased to various degrees dihydrolipoamide dehydrogenase activity, whereas dimer interface mutations also enhanced proteolytic and/or diaphorase activity. Human DLD proteins carrying each individual mutation complemented fully the respiratory-deficient phenotype of yeast cells lacking endogenous DLD even when residual dihydrolipoamide dehydrogenase activity was as low as 21% of controls. However, under elevated oxidative stress, expression of DLD proteins with dimer interface mutations greatly accelerated the loss of respiratory function, resulting from enhanced oxidative damage to the lipoic acid cofactor of pyruvate dehydrogenase and α-ketoglutarate dehydrogenase and other mitochondrial targets. This effect was not observed with the G194C mutation or a mutation that disrupts the proteolytic active site of DLD. As in yeast, lipoic acid cofactor was damaged in human D444V-homozygous fibroblasts after exposure to oxidative stress. We conclude that the cryptic activities of DLD promote oxidative damage to neighboring molecules and thus contribute to the clinical severity of DLD mutations.  相似文献   

9.
The pyruvate dehydrogenase complex of Bacillus stearothermophilus was treated with Staphylococcus aureus V8 proteinase, causing cleavage of the dihydrolipoamide acetyltransferase polypeptide chain (apparent Mr 57 000), inhibition of the enzymic activity and disassembly of the complex. Fragments of the dihydrolipoamide acetyltransferase chains with apparent Mr 28 000, which contained the acetyltransferase activity, remained assembled as a particle ascribed the role of an inner core of the complex. The lipoic acid residue of each dihydrolipoamide acetyltransferase chain was found as part of a small but stable domain that, unlike free lipoamide, was able still to function as a substrate for reductive acetylation by pyruvate in the presence of intact enzyme complex or isolated pyruvate dehydrogenase (lipoamide) component. The lipoyl domain was acidic and had an apparent Mr of 6500 (by sedimentation equilibrium), 7800 (by sodium dodecyl sulphate/polyacrylamide-gel electrophoresis) and 10 000 and 20 400 (by gel filtration in the presence and in the absence respectively of 6M-guanidinium chloride). 1H-n.m.r. spectroscopy of the dihydrolipoamide acetyltransferase inner core demonstrated that it did not contain the segments of highly mobile polypeptide chain found in the pyruvate dehydrogenase complex. 1H-n.m.r. spectroscopy of the lipoyl domain demonstrated that it had a stable and defined tertiary structure. From these and other experiments, a model of the dihydrolipoamide acetyltransferase chain is proposed in which the small, folded, lipoyl domain comprises the N-terminal region, and the large, folded, core-forming domain that contains the acetyltransferase active site comprises the C-terminal region. These two regions are separated by a third segment of the chain, which includes a substantial region of polypeptide chain that enjoys high conformational mobility and facilitates movement of the lipoyl domain between the various active sites in the enzyme complex.  相似文献   

10.
Spermatozoa are cells distinctly different from other somatic cells of the body, capacitation being one of the unique phenomena manifested by this gamete. We have shown earlier that dihydrolipoamide dehydrogenase, a post-pyruvate metabolic enzyme, undergoes capacitation-dependent tyrosine phosphorylation, and the functioning of the enzyme is required for hyperactivation (enhanced motility) and acrosome reaction of hamster spermatozoa (Mitra, K., and Shivaji, S. (2004) Biol. Reprod. 70, 887-899). In this report we have investigated the localization of this mitochondrial enzyme in spermatozoa revealing non-canonical extra-mitochondrial localization of the enzyme in mammalian spermatozoa. In hamster spermatozoa, dihydrolipoamide dehydrogenase along with its host complex, the pyruvate dehydrogenase complex, are localized in the acrosome and in the principal piece of the sperm flagella. The localization of dihydrolipoamide dehydrogenase, however, appears to be in the mitochondria in the spermatocytes, but in spermatids it appears to show a juxtanuclear localization (like Golgi). The capacitation-dependent time course of tyrosine phosphorylation of dihydrolipoamide dehydrogenase appears to be different in the principal piece of the flagella and the acrosome in hamster spermatozoa. Activity assays of this bi-directional enzyme suggest a strong correlation between the tyrosine phosphorylation and the bi-directional enzyme activity. This is the first report of a direct correlation of the localization, tyrosine phosphorylation, and activity of the important metabolic enzyme, dihydrolipoamide dehydrogenase, implicating dual involvement and regulation of the enzyme during sperm capacitation.  相似文献   

11.
The glycine decarboxylase complex consists of four different component enzymes (P-, H-, T- and L-proteins). The 14-kDa lipoamide-containing H-protein plays a pivotal role in the complete sequence of reactions as its prosthetic group (lipoic acid) interacts successively with the three other components of the complex and undergoes a cycle of reductive methylamination, methylamine transfer and electron transfer. With the aim to understand the interaction between the H-protein and its different partners, we have previously determined the crystal structure of the oxidized and methylaminated forms of the H-protein. In the present study, we have crystallized the H-protein in its reduced state and the L-protein (lipoamide dehydrogenase or dihydrolipoamide dehydrogenase). The L-protein has been overexpressed in Escherichia coli and refolded from inclusion bodies in an active form. Crystals were obtained from the refolded L-protein and the structure has been determined by X-ray crystallography. This first crystal structure of a plant dihydrolipoamide dehydrogenase is similar to other known dihydrolipoamide dehydrogenase structures. The crystal structure of the H-protein in its reduced form has been determined and compared to the structure of the other forms of the protein. It is isomorphous to the structure of the oxidized form. In contrast with methylaminated H-protein where the loaded lipoamide arm was locked into a cavity of the protein, the reduced lipoamide arm appeared freely exposed to the solvent. Such a freedom is required to allow its targeting inside the hollow active site of L-protein. Our results strongly suggest that a direct interaction between the H- and L-proteins is not necessary for the reoxidation of the reduced lipoamide arm bound to the H-protein. This hypothesis is supported by biochemical data [Neuburger, M., Polidori, A.M., Piètre, E., Faure, M., Jourdain, A., Bourguignon, J., Pucci, B. & Douce, R. (2000) Eur. J. Biochem. 267, 2882-2889] and by small angle X-ray scattering experiments reported herein.  相似文献   

12.
Enterococcus faecalis lipoamidase was discovered almost 50 years ago (Reed, L. J., Koike, M., Levitch, M. E., and Leach, F. R. (1958) J. Biol. Chem. 232, 143-158) as an enzyme activity that cleaved lipoic acid from small lipoylated molecules and from pyruvate dehydrogenase thereby inactivating the enzyme. Although the partially purified enzyme was a key reagent in proving the crucial role of protein-bound lipoic acid in the reaction mechanism of the 2-oxoacid dehydrogenases, the identity of the lipoamidase protein and the encoding gene remained unknown. We report isolation of the lipoamidase gene by screening an expression library made in an unusual cosmid vector in which the copy number of the vector is readily varied from 1-2 to 40-80 in an appropriate Escherichia coli host. Although designed for manipulation of large genome segments, the vector was also ideally suited to isolation of the gene encoding the extremely toxic lipoamidase. The gene encoding lipoamidase was isolated by screening for expression in E. coli and proved to encode an unexpectedly large protein (80 kDa) that contained the sequence signature of the Ser-Ser-Lys triad amidohydrolase family. The hexa-histidine-tagged protein was expressed in E. coli and purified to near-homogeneity. The purified enzyme was found to cleave both small molecule lipoylated and biotinylated substrates as well as lipoic acid from two 2-oxoacid dehydrogenases and an isolated lipoylated lipoyl domain derived from the pyruvate dehydrogenase E2 subunit. Lipoamidase-mediated inactivation of the 2-oxoacid dehydrogenases was observed both in vivo and in vitro. Mutagenesis studies showed that the residues of the Ser-Ser-Lys triad were required for activity on both small molecule and protein substrates and confirmed that lipoamidase is a member of the Ser-Ser-Lys triad amidohydrolase family.  相似文献   

13.
The major surface antigen of the bloodstream form of Trypanosoma brucei, the variant surface glycoprotein, is attached to the plasma membrane via a glycosylphosphatidylinositol anchor. The biosynthesis of the glycosylphosphatidylinositol anchor, as well as the assembly of the asparagine-linked oligosaccharide chains found on the variant surface glycoproteins, involves polyisoprenoid lipids that act as sugar carriers. Preliminary observations (Menon, A.K., Schwarz, R.T., Mayor, and Cross, G.A.M. (1990) J. Biol. Chem. 265, 9033-9042) suggested that the sugar carriers in T. brucei were short-chain polyisoprenoids containing substantially fewer isoprene residues than polyisoprenols in mammalian cells. In this paper we describe metabolic labeling experiments with [3H]mevalonate, as well as chromatographic and mass spectrometric analyses of products of the mevalonate pathway in T. brucei. We report that cells of the bloodstream form of T. brucei contain a limited spectrum of short chain dolichols and dolichol phosphates (11 and 12 isoprene residues). The total dolichol content was estimated to be 0.28 nmol/10(9) cells; the dolichyl phosphate content was 0.07 nmol/10(9) cells. The same spectrum of dolichol chain lengths was also found in a polar lipid that could be labeled with [3H]mevalonate, [3H]glucosamine, and [3H]mannose, and which was characterized as Man5GlcNAc2-PP-dolichol. The most abundant product of the mevalonate pathway identified in T. brucei was cholesterol (140 nmol/10(9) cells). Ubiquinone (0.09 nmol/10(9) cells) with a solanesol side chain was also identified.  相似文献   

14.
Cellular location of ganglioside-sialidase activity was determined in confluent hamster embryo fibroblasts transformed with herpes simplex virus type 2. Approximately equal specific activities of ganglioside-sialidase activity were found to be associated with the crude lysosomal and crude plasma membrane fractions isolated from whole cell homogenates. Whole transformed cells hydrolyzed exogenous ganglioside substrate, suggesting a partial location of the cellular sialidase on the outer surface of the plasma membrane of these cells. Intact cells were treated with the diazonium salt of sulfanilic acid, a nonpenetrating reagent inhibitory to ecto-enzymes (DePierre, J.W., and M. L. Karnovsky. 1974. J. Biol. Chem. 249:7111-7120). Cytoplasmic lactate dehydrogenase activity was not inhibited by this treatment, and mitochondrial succinate dehydrogenase activity was inhibited only 10%, indicating that intracellular enzymes were not affected. 5'-Nucleotidase activity was diminished 90%, and sialidase very rapidly lost 40% of its exogenously directed activity. These results show that, in herpes simplex virus-transformed fibroblasts, ganglioside-sialidase is both a lysosomal and a plasma membrane enzyme. The plasma membrane sialidase is capable of acting on endogenous plasma membrane sialolipids and also functions in the cultured transformed cell as an ecto-enzyme which can attack exogenous substrates.  相似文献   

15.
A cDNA clone encoding the entire transacylase (E2b) precursor of the bovine branched-chain alpha-keto acid dehydrogenase complex (Griffin, T. A., Lau, K. S., and Chuang, D. T. (1988) J. Biol. Chem. 263, 14008-14014) was used to construct a prokaryotic expression vector for recombinant mature E2b. The overexpression in Escherichia coli correlates with the presence near the 5'-terminus of the mature E2b coding region (nucleotides 20 to 28) of the sequence 5'-TCAAACT-CT-3'. It has been proposed that this sequence is involved in secondary mRNA recognition through interaction with the 5'-terminus of the bacterial 16 S rRNA. The mature E2b protein has transacylase activity when assayed with exogenous dihydrolipoamide and [1-14C] isovaleryl-CoA as substrates. However, the recombinant protein has no attached lipoic acid. This was established by the absence of radiolabel incorporation when transformed E. coli cells were grown in a medium containing DL-[2-3H]lipoic acid. The recombinant mature E2b protein was purified to greater than 95% purity in one step using Sepharose 4B column chromatography. The purified recombinant protein was shown to have a cubic 24-mer structure by electron microscopy and to possess a specific activity similar to that of the purified natural bovine E2b. The purified recombinant mature E2b was lipoylated in vitro in the presence of 2 mM ATP using a mitochondrial extract prepared from bovine liver. The above results provide the first evidence that the proper folding and assembly of mature bovine E2b is independent of the attachment of lipoyl moieties and that mammalian lipoylation activity is present in mitochondria.  相似文献   

16.
A common diagnostic feature of glycosylinositol phospholipid (GPI)-anchored proteins is their release from the membrane by a phosphatidylinositol-specific phospholipase C (PI-PLC). However, some GPI-anchored proteins are resistant to this enzyme. The best characterized example of this subclass is the human erythrocyte acetylcholinesterase, where the structural basis of PI-PLC resistance has been shown to be the acylation of an inositol hydroxyl group(s) (Roberts, W. L., Myher, J. J., Kuksis, A., Low, M. G., and Rosenberry, T. L. (1988) J. Biol. Chem. 263, 18766-18775). Both PI-PLC-sensitive and resistant GPI-anchor precursors (P2 and P3, respectively) have been found in Trypanosoma brucei, where the major surface glycoprotein is anchored by a PI-PLC-sensitive glycolipid anchor. The accompanying paper (Mayor, S., Menon, A. K., Cross, G. A. M., Ferguson, M. A. J., Dwek, R. A., and Rademacher, T. W. (1990) J. Biol. Chem. 265, 6164-6173) shows that P2 and P3 have identical glycans, indistinguishable from the common core glycan found on all the characterized GPI protein anchors. This paper shows that the single difference between P2 and P3, and the basis for the PI-PLC insusceptibility of P3, is a fatty acid, ester-linked to the inositol residue in P3. The inositol-linked fatty acid can be removed by treatment with mild base to restore PI-PLC sensitivity. Biosynthetic labeling experiments with [3H]palmitic acid and [3H]myristic acid show that [3H]palmitic acid specifically labels the inositol residue in P3 while [3H]myristic acid labels the diacylglycerol portion. Possible models to account for the simultaneous presence of PI-PLC-resistant and sensitive glycolipids are discussed in the context of available information on the biosynthesis of GPI-anchors.  相似文献   

17.
Full-length cDNA clones for succinyltransferase of the rat alpha-ketoglutarate dehydrogenase complex were isolated from rat heart cDNA libraries in lambda gt11. The cDNA clones were identified as those for rat succinyltransferase by the identity of their predicted amino acid sequence with the NH2-terminal amino acid sequence of rat succinyltransferase determined by protein chemical analysis and the known amino acid sequence of bovine succinyltransferase. The clone with the longest cDNA consisted of 2747 base pairs and coded for a leader peptide of 56 amino acid residues and a mature protein of 386 amino acid residues. The primary structure of rat succinyltransferase showed close similarity to Escherichia coli and Azotobacter vinelandii succinyltransferases, in the COOH-terminal part forming the lipoyl-binding domain and the NH2-terminal part forming the inner core-catalytic domain. However, the rat succinyltransferase did not contain a sequence motif that has been found as an E3- and/or E1-binding site in the dihydrolipoamide acyltransferases of three alpha-ketoacid dehydrogenase complexes (Hummel, K. B., Litwer, S., Bradford, A. P., Aitken, A., Danner, D. J., and Yeaman, S. J. (1988) J. Biol. Chem. 263, 6165-6168, Reed, L. J., and Hackert, M. L. (1990) J. Biol. Chem. 265, 8971-8974). The absence of this sequence was confirmed by direct sequencing of the polymerase chain reaction product of rat heart mRNA and by computer analysis. These results show that the rat succinyltransferase does not have the sequence motif of the putative E3- and/or E1-binding site.  相似文献   

18.
The glycine-utilizing bacterium Clostridium litoralis contained two enzyme systems for oxidizing dihydrolipoamide. The first one was found to be a genuine dihydrolipoamide dehydrogenase, present only in low amounts. This enzyme had the typical dimeric structure with a subunit molecular mass of about 53 kDa; however, it reacted with both NADP (Km 0.11 mM) and NAD (Km 0.5 mM). The reduction of pyridine nucleotides by dihydrolipoamide was the strongly preferred reaction. A second dihydrolipoamide-oxidizing enzyme system consisted of the interaction of two proteins, the previously described NADP(H)-dependent electron-transferring flavoprotein (D. Dietrichs, M. Meyer, B. Schmidt, and J. R. Andreesen, J. Bacteriol. 172:2088-2095, 1990) and a thioredoxin. This enzyme system was responsible for most of the dihydrolipoamide dehydrogenase activity in cell extracts. The thioredoxin did not bind to DEAE, was heat stable, and had a molecular mass of about 15 kDa. N-terminal amino acid analysis of the first 38 amino acid residues resulted in 38% homology to Escherichia coli thioredoxin and about 76% homology to a corresponding protein isolated from the physiologically close related Eubacterium acidaminophilum. The protein of the latter organism had a molecular mass of about 14 kDa and stimulated the low dihydrolipoamide dehydrogenase activity of the corresponding flavoprotein. By this interaction with NADPH-dependent flavoproteins, a new assay system for thioredoxin was established. A function of thioredoxin in glycine metabolism of some anaerobic bacteria is proposed.  相似文献   

19.
Previous studies in Trypanosoma brucei have shown that intracellular pH homeostasis is affected by inhibitors of H+-ATPases, suggesting a major role for these pumps in this process (Vander-Heyden, N., Wong, J., and Docampo, R., (2000) Biochem. J. 346, 53-62). Here, we report the cloning and sequencing of three genes (TbHA1, TbHA2, and TbHA3) present in the genome of T. brucei that encode proteins with homology to fungal and plant P-type proton-pumping ATPases. Northern and Western blot analyses revealed that these genes are up-regulated in procyclic trypomastigotes. TbHA1, TbHA2, and TbHA3 complemented a Saccharomyces cerevisiae strain deficient in P-type H+-ATPase activity, providing genetic evidence for their function. Indirect immunofluorescence analysis showed that TbHA proteins are localized mainly in the plasma membrane of procyclic forms and in the plasma membrane and flagellum of bloodstream forms. T. brucei H+-ATPase genes were functionally characterized using double-stranded RNA interference methodology. The induction of double-stranded RNA (RNA interference) caused growth inhibition, which was more accentuated in procyclic forms and when expression of all TbHA proteins was decreased. Knockdown of TbHA1 and TbHA3, but not of TbHA2, resulted in cells with a lower steady-state pH(i) and a slower rate of pH(i) recovery from acidification. No evidence was found of an intracellular P-type H+-ATPase activity. These results establish that T. brucei H+-ATPases are plasma membrane enzymes essential for parasite viability.  相似文献   

20.
Pleomorphic Trypanosoma brucei strains are characterized by their ability to differentiate from replicating long slender forms into non-dividing short stumpy forms in the mammalian host. The differentiation process can be efficiently induced in vitro by treatment with the membrane-permeable cAMP derivative 8-(4-chlorophenylthio)-cAMP (pCPTcAMP). In contrast, monomorphic T. brucei strains do not differentiate to stumpy forms in the host. Here, we show that exposure of monomorphic, culture-adapted T. brucei bloodstream forms to pCPTcAMP allowed their subsequent differentiation into short stumpy forms. The stumpy nature of pCPTcAMP-treated parasites was confirmed by (1) morphological change, (2) inhibition of growth and DNA synthesis, (3) cell cycle arrest in the G(1)/G(0) phase, (4) expression of NADH diaphorase activity and dihydrolipoamide dehydrogenase, (5) disappearance of the small subunit of ribonucleotide reductase, (6) up-regulation of the major lysosomal membrane protein, and (7) efficient transformation into replicating procyclic insect forms after induction with citrate/cis-aconitate. Our results indicate that the inability of monomorphic T. brucei bloodstream forms to differentiate into short stumpy forms in the host may be due to a failure in the signalling pathway rather than in the differentiation process itself. Treatment of monomorphic bloodstream trypanosomes with pCPTcAMP could be a useful method for identifying the genes involved in the slender-to-stumpy differentiation process.  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