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 fragmentary form of alanine racemase from Bacillus stearothermophilus is composed of two sets of two different polypeptides corresponding to the two domains of the subunit of wild-type enzyme. It was denatured with 6 M guanidium hydrochloride to be separated into pieces, and renatured by dilution with about 50% recovery of the activity. The two kinds of polypeptides (i.e., large and a small fragments) were isolated by gel filtration in the presence of 4 M guanidium hydrochloride. The CD spectra obtained by summation of the spectra of the refolded fragments were closely similar to that of the native fragmentary enzyme. The lysine residue to which PLP is bound in the wild-type enzyme occurs in the large peptide of the fragmentary enzyme containing the amino terminus of the wild-type enzyme. The visible spectrum of the large peptide refolded, however, indicated that PLP was not bound to it. The large peptide alone had no significant activity, but it was activated by incubation with the small peptide. Accordingly, co-existene of both peptide fragments is necessary for folding of a complex of the two kinds of peptide to form an active structure.  相似文献   

2.
Limited proteolysis studies on alanine racemase suggested that the enzyme subunit is composed of two domains (Galakatos, N. G., and Walsh, C. T. (1987) Biochemistry 26, 8475-8480). We have constructed a mutant gene that tandemly encodes the two polypeptides of the Bacillus stearothermophilus enzyme subunit cleaved at the position corresponding to the predicted hinge region. The mutant gene product purified was shown to be composed of two sets of the two polypeptide fragments and was immunologically identical to the wild-type enzyme. The mutant enzyme, i.e. the fragmentary alanine racemase, was active in both directions of the racemization of alanine. The maximum velocity (Vmax) was about half that of the wild-type enzyme, and the Km value was about double. Absorption and circular dichroism spectra of the fragmentary enzyme were similar to those of the wild-type enzyme. An attempt was made to separately express in Escherichia coli a single polypeptide corresponding to each domain, but no protein reactive with the antibody against the wild-type alanine racemase was produced. Therefore, it is suggested that the two polypeptide fragments can fold into an active structure only when they are co-translated and that they correspond to structural folding units in the parental polypeptide chain.  相似文献   

3.
NAD+ facilitates high-yield reactivation of clostridial glutamate dehydrogenase (GDH) after unfolding in urea. The specificity of this effect has been explored by using analogues and fragments of NAD+. The adenine portion, unlike the nicotinamide portion, is important for reactivation. Alteration in the nicotinamide portion, in acetylpyridine adenine dinucleotide, has little effect, whereas loss of the 6-NH2 substitution on the adenine ring, in 6-deamino NAD, diminishes the effectiveness of the nucleotide in promoting refolding. Also ADP-ribose, lacking nicotinamide, promotes reactivation whereas NMN-phosphoribose, lacking the adenine, does not. Of the smaller fragments, those containing an adenosine moiety, and especially those with one or more phosphate groups, impede the refolding ability of NAD+, and are able to bind to the folding intermediate though unable to facilitate refolding. These results are interpreted in terms of the known 3D structure for clostridial glutamate dehydrogenase. It is assumed that the refolding intermediate has a more or less fully formed NAD+-binding domain but a partially disordered substrate-binding domain and linking region. Binding of NAD+ or ADP-ribose appears to impose new structural constraints that result in completion of the correct folding of the second domain, allowing association of enzyme molecules to form the native hexamer.  相似文献   

4.
Proton-translocating transhydrogenases, reducing NADP+ by NADH through hydride transfer, are membrane proteins utilizing the electrochemical proton gradient for NADPH generation. The enzymes have important physiological roles in the maintenance of e.g. reduced glutathione, relevant for essentially all cell types. Following X-ray crystallography and structural resolution of the soluble substrate-binding domains, mechanistic aspects of the hydride transfer are beginning to be resolved. However, the structure of the intact enzyme is unknown. Key questions regarding the coupling mechanism, i.e., the mechanism of proton translocation, are addressed using the separately expressed substrate-binding domains. Important aspects are therefore which functions and properties of mainly the soluble NADP(H)-binding domain, but also the NAD(H)-binding domain, are relevant for proton translocation, how the soluble domains communicate with the membrane domain, and the mechanism of proton translocation through the membrane domain.  相似文献   

5.
L-threonine dehydrogenase (TDH) is an enzyme that catalyzes the oxidation of L-threonine to 2-amino-3-ketobutyrate. We solved the first crystal structure of a medium chain L-threonine dehydrogenase from a hyperthermophilic archaeon, Pyrococcus horikoshii (PhTDH), by the single wavelength anomalous diffraction method using a selenomethionine-substituted enzyme. This recombinant PhTDH is a homo-tetramer in solution. Three monomers of PhTDHs were located in the crystallographic asymmetric unit, however, the crystal structure exhibits a homo-tetramer structure with crystallographic and non-crystallographic 222 symmetry in the cell. Despite the low level of sequence identity to a medium-chain NAD(H)-dependent alcohol dehydrogenase (ADH) and the different substrate specificity, the overall folds of the PhTDH monomer and tetramer are similar to those of the other ADH. Each subunit is composed of two domains: a nicotinamide cofactor (NAD(H))-binding domain and a catalytic domain. The NAD(H)-binding domain contains the alpha/beta Rossmann fold motif, characteristic of the NAD(H)-binding protein. One molecule of PhTDH contains one zinc ion playing a structural role. This metal ion exhibits coordination with four cysteine ligands and some of the ligands are conserved throughout the structural zinc-containing ADHs and TDHs. However, the catalytic zinc ion that is coordinated at the bottom of the cleft in the case of ADH was not observed in the crystal of PhTDH. There is a significant difference in the orientation of the catalytic domain relative to the coenzyme-binding domain that results in a larger interdomain cleft.  相似文献   

6.
There is growing evidence that metabolic enzymes may act as multifunctional proteins performing diverse roles in cellular metabolism. Among these functions are the RNA-binding activities of NAD(+)-dependent dehydrogenases. Previously, we have characterized the glycolytic enzyme glyceraldehyde-3-phosphate dehydrogenase (GAPDH) as an RNA-binding protein with preference to adenine-uracil-rich sequences. In this study, we used GST-GAPDH fusion proteins generated by deletion mutagenesis to search for the RNA binding domain. We established that the N-terminal 43 amino acid residues of GAPDH, which correspond to the first mononucleotide-binding domain of the NAD(+)-binding fold is sufficient to confer RNA-binding. We also provide evidence that this single domain, although it retains most of the RNA-binding activity, loses sequence specificity. Our results suggest a molecular basis for RNA-recognition by NAD(+)-dependent dehydrogenases and (di)nucleotide-binding metabolic enzymes that had been reported to have RNA-binding activity with different specificity. To support this prediction we also identified other members of the family of NAD(+)-dependent dehydrogenases with no previous history of nucleic acid binding as RNA binding proteins in vitro. Based on our findings we propose the addition of the NAD(+)-binding domain to the list of RNA binding domains/motifs.  相似文献   

7.
The Bacillus methanolicus methanol dehydrogenase (MDH) is a decameric nicotinoprotein alcohol dehydrogenase (family III) with one Zn(2+) ion, one or two Mg(2+) ions, and a tightly bound cofactor NAD(H) per subunit. The Mg(2+) ions are essential for binding of cofactor NAD(H) in MDH. A B. methanolicus activator protein strongly stimulates the relatively low coenzyme NAD(+)-dependent MDH activity, involving hydrolytic removal of the NMN(H) moiety of cofactor NAD(H) (Kloosterman, H., Vrijbloed, J. W., and Dijkhuizen, L. (2002) J. Biol. Chem. 277, 34785-34792). Members of family III of NAD(P)-dependent alcohol dehydrogenases contain three unique, conserved sequence motifs (domains A, B, and C). Domain C is thought to be involved in metal binding, whereas the functions of domains A and B are still unknown. This paper provides evidence that domain A constitutes (part of) a new magnesium-dependent NAD(P)(H)-binding domain. Site-directed mutants D100N and K103R lacked (most of the) bound cofactor NAD(H) and had lost all coenzyme NAD(+)-dependent MDH activity. Also mutants G95A and S97G were both impaired in cofactor NAD(H) binding but retained coenzyme NAD(+)-dependent MDH activity. Mutant G95A displayed a rather low MDH activity, whereas mutant S97G was insensitive to activator protein but displayed "fully activated" MDH reaction rates. The various roles of these amino acid residues in coenzyme and/or cofactor NAD(H) binding in MDH are discussed.  相似文献   

8.
Aspartic acid 244 that occurs at the putative NAD(+)-binding site of rat liver S-adenosylhomocysteinase was replaced by glutamic acid by oligonucleotide-directed mutagenesis. The mutant enzyme was purified to homogeneity as judged by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Gel permeation chromatography showed that the purified mutant enzyme was a tetramer as is the wild-type enzyme. In contrast to the wild-type enzyme, which possesses 1 mol of tightly bound NAD+ per mol of enzyme subunit, the mutant enzyme had only 0.05 mol of NAD+ but contained about 0.6 mol each of NADH and adenine per mol of subunit. The mutant enzyme, after removal of the bound compounds by acid-ammonium sulfate treatment, exhibited S-adenosylhomocysteinase activity when assayed in the presence of NAD+. From the appearance of activity as a function of NAD+ concentration, the enzyme was shown to bind NAD+ with a Kd of 23.0 microM at 25 degrees C, a value greater than 280-fold greater than that of the wild-type enzyme. In the presence of a saturating concentration of NAD+, the mutant enzyme showed apparent Km values for substrates similar to those of the wild-type enzyme. Moderate decreases of 8- and 15-fold were observed in Vmax values for the synthetic and hydrolytic directions, respectively. These results indicate the importance of Asp-244 in binding NAD+, and are consistent with the idea that the region of S-adenosylhomocysteinase from residues 213 to 244 is part of the NAD+ binding site. This region has structural features characteristic of the dinucleotide-binding domains of NAD(+)- and FAD-binding proteins (Ogawa, H., Gomi, T., Mueckler, M. M., Fujioka, M., Backlund, P.S., Jr., Aksamit, R.R., Unson, C.G., and Cantoni, G.L. (1987) Proc. Natl. Acad. Sci. U.S.A. 84, 719-723).  相似文献   

9.
Vallon O 《Proteins》2000,38(1):95-114
We describe two new sequence motifs, present in several families of flavoproteins. The "GG motif" (RxGGRxxS/T) is found shortly after the betaalphabetadinucleotide-binding motif (DBM) in L-amino acid oxidases, achacin and aplysianin-A, monoamine oxidases, corticosteroid-binding proteins, and tryptophan 2-monooxygenases. Other disperse sequence similarities between these families suggest a common origin. A GG motif is also found in protoporphyrinogen oxidase and carotenoid desaturases and, reduced to the central GG doublet, in the THI4 protein, dTDP-4-dehydrorhamnose reductase, soluble fumarate reductase, steroid dehydrogenases, Rab GDP-dissociation inhibitor, and in most flavoproteins with two dinucleotide-binding domains (glutathione reductase, glutamate synthase, flavin-containing monooxygenase, trimethylamine dehydrogenase...). In the latter families, an "ATG motif" (oxhhhATG) is found in both the FAD- and NAD(P)H-binding domains, forming the fourth beta-strand of the Rossman fold and the connecting loop. On the basis of these and previously described motifs, we present a classification of dinucleotide-binding proteins that could also serve as an evolutionary scheme. Like the DBM, the ATG motif appears to predate the divergence of NAD(P)H- and FAD-binding proteins. We propose that flavoproteins have evolved from a well-differentiated NAD(P)H-binding protein. The bulk of the substrate-binding domain was formed by an insertion after the fourth beta-strand, either of a closely related NAD(P)H-binding domain or of a domain of completely different origin.  相似文献   

10.
11.
Drosophila alcohol dehydrogenase (ADH), an NAD(+)-dependent dehydrogenase, shares little sequence similarity with horse liver ADH. However, these two enzymes do have substantial similarity in their secondary structure at the NAD(+)-binding domain [Benyajati, C., Place, A. P., Powers, D. A. & Sofer, W. (1981) Proc. Natl Acad. Sci. USA 78, 2717-2721]. Asp38, a conserved residue between Drosophila and horse liver ADH, appears to interact with the hydroxyl groups of the ribose moiety in the AMP portion of NAD+. A secondary-structure comparison between the nucleotide-binding domain of NAD(+)-dependent enzymes and that of NADP(+)-dependent enzymes also suggests that Asp38 could play an important role in cofactor specificity. Mutating Asp38 of Drosophila ADH into Asn38 decreases Km(app)NADP 62-fold and increases kcat/Km(app)NADP 590-fold at pH 9.8, when compared with wild-type ADH. These results suggest that Asp38 is in the NAD(+)-binding domain and its substituent, Asn38, allows Drosophila ADH to use both NAD+ and NADP+ as its cofactor. The observations from the experiments of thermal denaturation and kinetic measurement with pH also confirm that the repulsion between the negative charges of Asp38 and 2'-phosphate of NADP+ is the major energy barrier for NADP+ to serve as a cofactor for Drosophila ADH.  相似文献   

12.
The open reading frame TM1643 of Thermotoga maritima belongs to a large family of proteins, with homologues in bacteria, archaea, and eukaryotes. TM1643 is found in an operon with two other genes that encode enzymes involved in the biosynthesis of NAD. In several bacteria, the gene in the position occupied by TM1643 encodes an aspartate oxidase (NadB), which synthesizes iminoaspartate as a substrate for NadA, the next enzyme in the pathway. The amino acid sequence of TM1643 does not share any recognizable homology with aspartate oxidase or with other proteins of known functions or structures. To help define the biological functions of TM1643, we determined its crystal structure at 2.6A resolution and performed a series of screens for enzymatic function. The structure reveals the presence of an N-terminal Rossmann fold domain with a bound NAD(+) cofactor and a C-terminal alpha+beta domain. The structural information suggests that TM1643 may be a dehydrogenase and the active site of the enzyme is located at the interface between the two domains. The enzymatic characterization of TM1643 revealed that it possesses NAD or NADP-dependent dehydrogenase activity toward l-aspartate but no aspartate oxidase activity. The product of the aspartate dehydrogenase activity is also iminoaspartate. Therefore, our studies demonstrate that two different enzymes, an oxidase and a dehydrogenase, may have evolved to catalyze the first step of NAD biosynthesis in prokaryotes. TM1643 establishes a new class of amino acid dehydrogenases.  相似文献   

13.
BACKGROUND: Bacillus stearothermophilus glycerol dehydrogenase (GlyDH) (glycerol:NAD(+) 2-oxidoreductase, EC 1.1.1.6) catalyzes the oxidation of glycerol to dihydroxyacetone (1,3-dihydroxypropanone) with concomitant reduction of NAD(+) to NADH. Analysis of the sequence of this enzyme indicates that it is a member of the so-called iron-containing alcohol dehydrogenase family. Despite this sequence similarity, GlyDH shows a strict dependence on zinc for activity. On the basis of this, we propose to rename this group the family III metal-dependent polyol dehydrogenases. To date, no structural data have been reported for any enzyme in this group. RESULTS: The crystal structure of B. stearothermophilus glycerol dehydrogenase has been determined at 1.7 A resolution to provide structural insights into the mechanistic features of this family. The enzyme has 370 amino acid residues, has a molecular mass of 39.5 kDa, and is a homooctamer in solution. CONCLUSIONS: Analysis of the crystal structures of the free enzyme and of the binary complexes with NAD(+) and glycerol show that the active site of GlyDH lies in the cleft between the enzyme's two domains, with the catalytic zinc ion playing a role in stabilizing an alkoxide intermediate. In addition, the specificity of this enzyme for a range of diols can be understood, as both hydroxyls of the glycerol form ligands to the enzyme-bound Zn(2+) ion at the active site. The structure further reveals a previously unsuspected similarity to dehydroquinate synthase, an enzyme whose more complex chemistry shares a common chemical step with that catalyzed by glycerol dehydrogenase, providing a striking example of divergent evolution. Finally, the structure suggests that the NAD(+) binding domain of GlyDH may be related to that of the classical Rossmann fold by switching the sequence order of the two mononucleotide binding folds that make up this domain.  相似文献   

14.
Reger AS  Carney JM  Gulick AM 《Biochemistry》2007,46(22):6536-6546
The adenylate-forming enzymes, including acyl-CoA synthetases, the adenylation domains of non-ribosomal peptide synthetases (NRPS), and firefly luciferase, perform two half-reactions in a ping-pong mechanism. We have proposed a domain alternation mechanism for these enzymes whereby, upon completion of the initial adenylation reaction, the C-terminal domain of these enzymes undergoes a 140 degrees rotation to perform the second thioester-forming half-reaction. Structural and kinetic data of mutant enzymes support this hypothesis. We present here mutations to Salmonella enterica acetyl-CoA synthetase (Acs) and test the ability of the enzymes to catalyze the complete reaction and the adenylation half-reaction. Substitution of Lys609 with alanine results in an enzyme that is unable to catalyze the adenylate reaction, while the Gly524 to leucine substitution is unable to catalyze the complete reaction yet catalyzes the adenylation half-reaction with activity comparable to the wild-type enzyme. The positions of these two residues, which are located on the mobile C-terminal domain, strongly support the domain alternation hypothesis. We also present steady-state kinetic data of putative substrate-binding residues and demonstrate that no single residue plays a dominant role in dictating CoA binding. We have also created two mutations in the active site to alter the acyl substrate specificity. Finally, the crystallographic structures of wild-type Acs and mutants R194A, R584A, R584E, K609A, and V386A are presented to support the biochemical analysis.  相似文献   

15.
Proton-translocating nicotinamide nucleotide transhydrogenase is a conformationally driven pump which catalyzes the reversibel reduction of NADP(+) by NADH. Transhydrogenases contain three domains, i.e., the hydrophilic NAD(H)-binding domain I and the NADP(H)-binding domain III, and the hydrophobic domain II containing the proton channel. Domains I and III have been separately expressed and characterized structurally by, e.g. X-ray crystallography and NMR. These domains catalyze transhydrogenation in the absence of domain II. However, due to the absence of the latter domain, the reactions catalyzed by domains I and III differ significantly from those catalyzed by the intact enzyme. Mutagenesis of residues in domain II markedly affects the activity of the intact enzyme. In order to resolve the structure-function relationships of the intact enzyme, and the molecular mechanism of proton translocation, it is therefore essential to establish the structure and function of domain II and its interactions with domains I and III. This review describes some relevant recent results in this field of research.  相似文献   

16.
Crystal structure of S-adenosylhomocysteine hydrolase from rat liver.   总被引:5,自引:0,他引:5  
The crystal structure of rat liver S-adenosyl-L-homocysteine hydrolase (AdoHcyase, EC 3.3.1.1) which catalyzes the reversible hydrolysis of S-adenosylhomocysteine (AdoHcy) has been determined at 2.8 A resolution. AdoHcyase from rat liver is a tetrameric enzyme with 431 amino acid residues in each identical subunit. The subunit is composed of the catalytic domain, the NAD+-binding domain, and the small C-terminal domain. Both catalytic and NAD+-binding domains are folded into an ellipsoid with a typical alpha/beta twisted open sheet structure. The C-terminal section is far from the main body of the subunit and extends into the opposite subunit. An NAD+ molecule binds to the consensus NAD+-binding cleft of the NAD+-binding domain. The peptide folding pattern of the catalytic domain is quite similar to the patterns observed in many methyltransferases. Although the crystal structure does not contain AdoHcy or its analogue, there is a well-formed AdoHcy-binding crevice in the catalytic domain. Without introducing any major structural changes, an AdoHcy molecule can be placed in the catalytic domain. In the structure described here, the catalytic and NAD+-binding domains are quite far apart from each other. Thus, the enzyme appears to have an "open" conformation in the absence of substrate. It is likely that binding of AdoHcy induces a large conformational change so as to place the ribose moiety of AdoHcy in close proximity to the nicotinamide moiety of NAD+. A catalytic mechanism of AdoHcyase has been proposed on the basis of this crystal structure. Glu155 acts as a proton acceptor from the O3'-H when the proton of C3'-H is abstracted by NAD+. His54 or Asp130 acts as a general acid-base catalyst, while Cys194 modulates the oxidation state of the bound NAD+. The polypeptide folding pattern of the catalytic domain suggests that AdoHcy molecules can travel freely to and from AdoHcyase and methyltransferases to properly regulate methyltransferase activities. We believe that the crystal structure described here can provide insight into the molecular architecture of this important regulatory enzyme.  相似文献   

17.
The gene for leucine dehydrogenase (EC 1.4.1.9) from Bacillus stearothermophilus was cloned and expressed in Escherichia coli. The selection for the cloned gene was based upon activity staining of the replica printed E. coli cells. A transformant showing high leucine dehydrogenase activity was found to carry an about 9 kilobase pair plasmid, which contained 4.6 kilobase pairs of B. stearothermophilus DNA. The nucleotide sequence including the 1287 base pair coding region of the leucine dehydrogenase gene was determined by the dideoxy chain termination method. The translated amino acid sequence was confirmed by automated Edman degradation of several peptide fragments produced from the purified enzyme by trypsin digestion. The polypeptide contained 429 amino acid residues corresponding to the subunit (Mr 49,000) of the hexameric enzyme. Comparison of the amino acid sequence of leucine dehydrogenase with those of other pyridine nucleotide dependent oxidoreductases registered in a protein data bank revealed significant sequence similarity, particularly between leucine and glutamate dehydrogenases, in the regions containing the coenzyme binding domain and certain specific residues with catalytic importance.  相似文献   

18.
(S)-1-Phenylethanol dehydrogenase (PED) from the denitrifying bacterium strain EbN1 catalyzes the NAD+-dependent, stereospecific oxidation of (S)-1-phenylethanol to acetophenone and the biotechnologically interesting reverse reaction. This novel enzyme belongs to the short-chain alcohol dehydrogenase/aldehyde reductase family. The coding gene (ped) was heterologously expressed in Escherichia coli and the purified protein was crystallized. The X-ray structures of the apo-form and the NAD+-bound form were solved at a resolution of 2.1 and 2.4 A, respectively, revealing that the enzyme is a tetramer with two types of hydrophobic dimerization interfaces, similar to beta-oxoacyl-[acyl carrier protein] reductase (FabG) from E. coli. NAD+-binding is associated with a conformational shift of the substrate binding loop of PED from a crystallographically unordered "open" to a more ordered "closed" form. Modeling the substrate acetophenone into the active site revealed the structural prerequisites for the strong enantioselectivity of the enzyme and for the catalytic mechanism. Studies on the steady-state kinetics of PED indicated a highly positive cooperativity of both catalytic directions with respect to the substrates. This is contrasted by the behavior of FabG. Moreover, PED exhibits extensive regulation on the enzyme level, being inhibited by elevated concentrations of substrates and products, as well as the wrong enantiomer of 1-phenylethanol. These regulatory properties of PED are consistent with the presence of a putative "transmission module" between the subunits. This module consists of the C-terminal loops of all four subunits, which form a special interconnected structural domain and mediate close contact of the subunits, and of a phenylalanine residue in each subunit that reaches out between substrate-binding loop and C-terminal domain of an adjacent subunit. These elements may transmit the substrate-induced conformational change of the substrate binding loop from one subunit to the others in the tetrameric complex and thus mediate the cooperative behavior of PED.  相似文献   

19.
Native rat liver methylmalonate semialdehyde dehydrogenase was proteolyzed by lysylendopeptidase C, chymotrypsin, and trypsin to generate different cleavage fragments of molecular masses: 50, 8, 55, 44, 39, 53, 45, and 40 kDa. A proteolytic cleavage map of MMSDH was constructed based on sequencing data and a comparison of appearance and degradation rates of the different protein fragments as shown by SDS-PAGE. NAD+ was highly effective as a protector against proteolysis in both the N-terminal and the C-terminal parts of the intact enzyme. NADH did not efficiently protect the intact enzyme; however, it stabilized proteolytic fragment L50 from further degradation. This suggests that the NAD(+)-binding domain is not destroyed by cleavage of the N-terminal part of MMSDH. CoA had no effect on the proteolytic cleavage patterns of MMSDH. However, CoA esters reduced the protective effect of NAD+ with an order of effectiveness of acetyl-CoA greater than propionyl-CoA greater than butyryl-CoA. p-Nitrophenyl acetate, substrate for esterase activity by the enzyme, partially prevented the protective effect of NAD+ against proteolysis. These results suggest that S-acylation of the enzyme prevents a stabilizing conformational change induced in MMSDH by NAD+ binding.  相似文献   

20.
Kinetic studies of Morris 7777 hepatoma mitochondrial NAD(P) malic enzyme were consistent with an ordered mechanism where NAD adds to the enzyme before malate and dissociation of NADH from the enzyme is rate-limiting. In addition to its active site, malate apparently also associates with a lower affinity with an activator site. The activator fumarate competes with malate at the activator site and facilitates dissociation of NADH from the enzyme. The ratio of NAD(P) malic enzyme to malate dehydrogenase activity in the hepatoma mitochondrial extract was found to be too low, even in the presence of known inhibitors of malate dehydrogenase, to account for the known ability of NAD(P) malic enzyme to intercept exogenous malate from malate dehydrogenase in intact tumor mitochondria (Moreadith, R.W., and Lehninger, A.L. (1984) J. Biol. Chem. 259, 6215-6221). However, NAD(P) malic enzyme may be able to intercept exogenous malate because according to the present results, it can associate with the pyruvate dehydrogenase complex, which could localize NAD(P) malic enzyme in the vicinity of the inner mitochondrial membrane. The activity levels of some key metabolic enzymes were found to be different in Morris 7777 mitochondria than in liver or mitochondria of other rapidly dividing tumors. These results are discussed in terms of differences among tumors in their ability to utilize malate, glutamate, and citrate as respiratory fuels.  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