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 796 毫秒
1.
The degradation of hemoglobin by the malaria parasite, Plasmodium falciparum, produces free ferriprotoporphyrin IX (FP) as a toxic by-product. In the presence of FP-binding drugs such as chloroquine, FP detoxification is inhibited, and the build-up of free FP is thought to be a key mechanism in parasite killing. In an effort to identify parasite proteins that might interact preferentially with FP, we have used a mass spectrometry approach. Proteins that bind to FP immobilized on agarose include P. falciparum glyceraldehyde-3-phosphate dehydrogenase (PfGAPDH), P. falciparum glutathione reductase (PfGR), and P. falciparum protein disulfide isomerase. To examine the potential consequences of FP binding, we have examined the ability of FP to inhibit the activities of GAPDH and GR from P. falciparum and other sources. FP inhibits the enzymic activity of PfGAPDH with a Ki value of 0.2 microm, whereas red blood cell GAPDH is much less sensitive. By contrast, PfGR is more resistant to FP inhibition (Ki > 25 microm) than its human counterpart. We also examined the ability of FP to inhibit the activities of the additional antioxidant enzymes, P. falciparum thioredoxin reductase, which exhibits a Ki value of 1 microm, and P. falciparum glutaredoxin, which shows more moderate sensitivity to FP. The exquisite sensitivity of PfGAPDH to FP may indicate that the glycolytic pathway of the parasite is particularly susceptible to modulation by FP stress. Inhibition of this pathway may drive flux through the pentose phosphate pathway ensuring sufficient production of reducing equivalents to counteract the oxidative stress induced by FP build-up.  相似文献   

2.
Glyceraldehyde 3-phosphate dehydrogenase is a tetramer of four chemically identical subunits which requires the cofactor nicotinamide adenine dinucleotide (NAD) for activity. The structure of the holo-enzyme from Bacillus stearothermophilus has recently been refined using X-ray data to 2.4 A resolution. This has facilitated the structure determination of both the apo-enzyme and the enzyme with one molecule of NAD bound to the tetramer. These structures have been refined at 4 A resolution using the constrained-restrained parameter structure factor least-squares refinement program CORELS. When combined with individual atomic temperature factors from the holo-enzyme, these refined models give crystallographic R factors of 30.2% and 30.4%, respectively, for data to 3 A resolution. The apo-enzyme has 222 molecular symmetry, and the subunit structure is related to that of the holo-enzyme by an approximate rigid-body rotation of the coenzyme binding domain by 4.3 degrees with respect to the catalytic domains, which form the core of the tetramer. The effect of this rotation is to shield the coenzyme and active site from solvent in the holo-enzyme. In addition to the rigid-body rotation, there is a rearrangement of several residues involved in NAD binding. The structure of the 1 NAD enzyme is asymmetric. The subunit which contains the bound NAD adopts a conformation very similar to that of a holo-enzyme subunit, while the other three unliganded subunits are very similar to the apo-enzyme conformation. This result provides unambiguous evidence for ligand-induced sequential conformational changes in B. stearothermophilus glyceraldehyde 3-phosphate dehydrogenase.  相似文献   

3.
Glyceraldehyde-3-phosphate dehydrogenase (D-glyceraldehyde-3-phosphate: NAD+ oxidoreductase (phosphorylating), EC 1.2.1.12) was isolated from a sturgeon, Huso huso, from the Caspian Sea. It is closely related to the enzyme from a Pacific sturgeon, Acipenser transmontanus, with respect to amino acid composition, steady-state kinetics and coenzyme binding. The latter, as studied by means of a spin-labeled derivative of NAD+, is negatively cooperative exhibiting a Hill coefficient of 0.84 at 12 degrees C. Two derivatives of NAD+ spin-labeled at N6 or C8 of the adenine ring were found to be active coenzymes with maximum velocities reaching 35 or 45% of the value for NAD+ itself. When more than two equivalents of either spin-labeled NAD+ are bound to the enzyme spin-spin interactions are observed in the ESR spectra. Distances between the nitroxide radicals (8--9 A) calculated from the observed splittings are in excellent agreement with data predicted from the crystal structure of the lobster enzyme when the coenzyme is bound in an anti-conformation of the adenine moiety about the glycosidic bond to all four subunits.  相似文献   

4.
The chloroplast NADH dehydrogenase-like complex (NDH) was first discovered based on its similarity to complex I in respiratory electron transport, and is involved in electron transport from photoproduced stromal reductants such as NADPH and ferredoxin to the intersystem plastoqunone pool. However, a recent study suggested that it is a ferredoxin-dependent plastoquinone reductase rather than an NAD(P)H dehydrogenase. Furthermore, recent advances in subunit analysis of NDH have revealed the presence of a novel hydrophilic subcomplex on the stromal side of the thylakoid membrane, as well as an unexpected lumenal subcomplex. This review discusses these new studies on the structure of NDH, and proposes a unified nomenclature for newly discovered NDH subunits.  相似文献   

5.
On site heterogeneity in sturgeon muscle GPDH: a kinetic approach   总被引:1,自引:0,他引:1  
The kinetics and stoichiometry of the reaction of sturgeon muscle glyceraldchyde-3-PO4-dehydrogenase (GPDH) with the disulfide interchange reagent bis(2,2' dithio-bis(5-nitrobenzoate) (DTNB) has been studied in detail. The native enzyme, a tetramer of covalently identical subunits, reacts relatively rapidly with precisely four equivalents of reagent, although there are three cysteine residues per subunit (12 per tetratner). Reaction of these four cystcines leads to total catalytic inactivation; the extent of inactivation is proportional to the fractional reaction. The rate of reaction is dependent on the extent of bound NAD: reactivity being very much greater at unliganded sites. The reaction with apo-enzyme is fastest, bimolecular and monophasic. Over a wide range of NAD concentration, however, the reaction of enzyme with a large molar excess of reagent is precisely biphasic, and each individual kinetic experiment can be analytically described by two pseudo first-order (NAD concentration-dependent) rate constants and two unequal NAD concentration-insensitive amplitudes. The biphasicity in rate is quantitatively explainable on the basis of a C2 symmetry for the tetrameric subunits with a tighter binding of NAD at two of the four sites, if high reactivity is exclusively dependent on the absence of bound NAD. The inequality in the two amplitudes, however, requires either a more complex or a more dynamic model. Arguments are presented for the appropriateness of a C2 symmetry model in which intramolecular transconformational isomerization of tight and loose NAD binding sites is possible. The equilibrium constant for the isomerization is estimable from the macroscopic specific rates and amplitudes. This “flip-over” C2 symmetry model is apropos to all situations of negative cooperativity in ligand binding to tetramers, as is discussed.  相似文献   

6.
In this study we intended to examine the extent of genetic diversity of Plasmodium falciparum parasites causing severe malaria (SM). For this purpose, 100 parasite isolates were obtained from patients with SM and uncomplicated malaria, from an area of low and unstable malaria transmission in Sudan. The diversity of infection (DOI) was estimated by relating the number of the different parasite genotypes that were detected to the total number of parasites that were genotyped (parasite population/subpopulation). We used different molecular markers individually (pfcrt-76, pfmr1-86, GLURP size and MSP2 family and size) and as a group to set a multilocus genetic profile for each parasite isolate. The DOI as estimated by MSP2 and GLURP was 0.553 and 0.435, respectively. However, combination of all four molecular markers (multilocus genetic profile) revealed a fingerprint pattern of genetic diversity with a DOI of 0.936, indicating that in SM infection, diversity is the rule and homogeny is the exception. Furthermore, our clinical data suggest that the virulence markers might also be more diverse than expected. In conclusion, the results are unexpected and overturn the assumption that parasites causing SM are a limited subpopulation of virulent parasites or of a clonal nature. However, it was more likely that there was a genetically unique parasite in each infection.  相似文献   

7.
Barker CD  Reda T  Hirst J 《Biochemistry》2007,46(11):3454-3464
Complex I (NADH:ubiquinone oxidoreductase) from bovine heart mitochondria contains 45 different subunits and nine redox cofactors. NADH is oxidized by a noncovalently bound flavin mononucleotide (FMN), then seven iron-sulfur clusters transfer the two electrons to quinone, and four protons are pumped across the inner mitochondrial membrane. Here, we use protein film voltammetry to investigate the mechanisms of NADH oxidation and NAD+ reduction in the simplest catalytically active subcomplex of complex I, the flavoprotein (Fp) subcomplex. The Fp subcomplex was prepared using chromatography and contained the 51 and 24 kDa subunits, the FMN, one [4Fe-4S] cluster, and one [2Fe-2S] cluster. The reduction potential of the FMN in the enzyme's active site is lower than that of free FMN (thus, the oxidized state of the FMN is most strongly bound) and close to the reduction potential of NAD+. Consequently, the catalytic transformation is reversible. Electrocatalytic NADH oxidation by subcomplex Fp can be explained by a model comprising substrate mass transport, the Michaelis-Menten equation, and interfacial electron transfer kinetics. The difference between the "catalytic" potential and the FMN potential suggests that the flavin is reoxidized before NAD+ is released or that intramolecular electron transfer from the flavin to the [4Fe-4S] cluster influences the catalytic rate. NAD+ reduction displays a marked activity maximum, below which the catalytic rate decreases sharply as the driving force increases. Two possible models reproduce the observed catalytic waveshapes: one describing an effect from reducing the proximal [2Fe-2S] cluster and the other the enhanced catalytic ability of the semiflavin state.  相似文献   

8.
Cucumber mosaic virus (CMV), the type member of the genus Cucumovirus (family Bromoviridae), is transmitted by aphids in a nonpersistent manner. Mutagenesis experiments identified the betaH-betaI loop of the capsid subunit as a potential key motif responsible for interactions with the insect vector. To further examine the functional characteristics of this motif, we generated monoclonal antibodies that bound to native virions but not to betaH-betaI mutants. Fab fragments from these antibodies were complexed with wild-type CMV and the virus-Fab structure was determined to 12-A resolution by using electron cryomicroscopy and image reconstruction techniques. The electron density attributed to the bound antibody has a turret-like appearance and protrudes from each of the 12 fivefold axes of the icosahedral virus. Thus, the antibody binds only to the pentameric clusters (pentons) of A subunits of the T=3 quasisymmetric virus and does not appear to bind to any of the B and C subunits that occur as hexameric clusters (hexons) at the threefold (quasi-sixfold) axes. Modeling and electron density comparisons were used to analyze the paratope-epitope interface and demonstrated that the antibody binds to three betaH-betaI loops in three adjacent A subunits in each penton. This antibody can discriminate between A and B/C subunits even though the betaH-betaI loop adopts the same structure in all 180 capsid subunits and is therefore recognizing differences in subunit arrangements. Antibodies with such character have potential use as probes of viral assembly. Our results may provide an additional rationale for designing synthetic vaccines by using symmetrical viral particles.  相似文献   

9.
The fluorescence of the natural coenzyme, NADH, is used to monitor the environment of the nicotinamide moiety at the active centre of rabbit muscle glyceraldehyde-3-phosphate dehydrogenase (EC 1.2.1.12). Changes of the fluorescence quantum yield and polarization of a small amount of NADH, totally bound by an excess of enzyme, show that at half-saturation of the oligomer with NAD a conformational change is induced which affects the active centre regions of the remaining subunits. This conformational transition is not effected by adenosine diphosphoribose, suggesting that the binding of the nicotinamide moiety of NAD to two subunits is essential for the change of tertiary structure of the remaining subunits that causes the observed changes of the fluorescence properties of the ADH "tracer probe". It is suggested that this conformational transition of the oligomer is responsible for the major decrease of affinity for NAD which occurs at half-saturation, and possibly for the activation by NAD+ of the reductive dephosphorylation reaction catalysed by the enzyme. It is also suggested, by analogy with haemoglobin, that the molecular basis of the negative cooperativity may be the creation of additional intersubunit bonds during the binding of the first two NAD molecules to the tetramer, and a change from a "relaxed" quaternary structure to a "tense" structure at half-saturation.  相似文献   

10.
The aquaglyceroporin of Plasmodium falciparum (PfAQP) is a bi-functional channel with permeability for water and solutes. Its functions supposedly are in osmotic protection of parasites and in facilitation of glycerol permeation for glycerolipid biosynthesis. Here, we show PfAQP permeability for the glycolysis-related metabolites methylglyoxal, a cytotoxic byproduct, and dihydroxyacetone, a ketotriose. AQP3, the red cell aquaglyceroporin, also passed dihydroxacetone but excluded methylglyoxal. Proliferation of malaria parasites was inhibited by methylglyoxal with an IC50 around 200 microM. Surprisingly, also dihydroxyacetone, which is an energy source in human cells, was antiproliferative in chloroquine-sensitive and resistant strains with an IC50 around 3 mM. We expressed P. falciparum glyceraldehyde 3-phosphate dehydrogenase (PfGAPDH) to examine whether it is inhibited by either carbonyl compound. Methylglyoxal did not affect PfGAPDH on incubation with 2.5 mM for 20 h. Treatment with 2.5 mM dihydroxyacetone, however, abolished PfGAPDH activity within 6 h. Aquaglyceroporin permeability for glycolytic metabolites may thus be of physiological significance.  相似文献   

11.
The binding of a spin-labeled AMP analog to tetrameric glyceraldehyde-3-phosphate dehydrogenase from rabbit muscle is described. The spin label, perdeuterated and 15N-substituted 4-amino-2,2,6,6-tetramethylpiperidine-1-oxyl, was attached to C-8 of AMP (C8-SL-AMP). Up to 8 equivalents of C8-SL-AMP bind per enzyme tetramer, i.e., 2 per monomer. Combining sites are the adenine subsite of the coenzyme-binding domain and the phosphate site. Glyceraldehyde 3-phosphate causes a conformational change in the enzyme that brings C8-SL-AMP molecules bound to adjacent R-axis-related subunits closer to one another by 0.2-0.3 nm and allows for spin-spin interaction between the nitroxide radicals. Similar, but less pronounced structural changes take place upon lowering the pH from 8 to 7. Addition of a single equivalent of NAD+ to a complex of the enzyme with 7.6 equivalents of C8-SL-AMP leads to the release of almost 4 C8-SL-AMP molecules. This supports our previous findings that binding of just one NAD+ molecule induces conformational changes in all four subunits.  相似文献   

12.
Perdeuterated spin label (DSL) analogs of NAD+, with the spin label attached at either the C8 or N6 position of the adenine ring, have been employed in an EPR investigation of models for negative cooperativity binding to tetrameric glyceraldehyde-3-phosphate dehydrogenase and conformational changes of the DSL-NAD+-enzyme complex during the catalytic reaction. C8-DSL-NAD+ and N6-DSL-NAD+ showed 80 and 45% of the activity of the native NAD+, respectively. Therefore, these spin-labeled compounds are very efficacious for investigations of the motional dynamics and catalytic mechanism of this dehydrogenase. Perdeuterated spin labels enhanced spectral sensitivity and resolution thereby enabling the simultaneous detection of spin-labeled NAD+ in three conditions: (1) DSL-NAD+ freely tumbling in the presence of, but not bound to, glyceraldehyde-3-phosphate dehydrogenase, (2) DSL-NAD+ tightly bound to enzyme subunits remote (58 A) from other NAD+ binding sites, and (3) DSL-NAD+ bound to adjacent monomers and exhibiting electron dipolar interactions (8-9 A or 12-13 A, depending on the analog). Determinations of relative amounts of DSL-NAD+ in these three environments and measurements of the binding constants, K1-K4, permitted characterization of the mathematical model describing the negative cooperativity in the binding of four NAD+ to glyceraldehyde-3-phosphate dehydrogenase. For enzyme crystallized from rabbit muscle, EPR results were found to be consistent with the ligand-induced sequential model and inconsistent with the pre-existing asymmetry models. The electron dipolar interaction observed between spin labels bound to two adjacent glyceraldehyde-3-phosphate dehydrogenase monomers (8-9 or 12-13 A) related by the R-axis provided a sensitive probe of conformational changes of the enzyme-DSL-NAD+ complex. When glyceraldehyde-3-phosphate was covalently bound to the active site cysteine-149, an increase in electron dipolar interaction was observed. This increase was consistent with a closer approximation of spin labels produced by steric interactions between the phosphoglyceryl residue and DSL-NAD+. Coenzyme reduction (DSL-NADH) or inactivation of the dehydrogenase by carboxymethylation of the active site cysteine-149 did not produce changes in the dipolar interactions or spatial separation of the spin labels attached to the adenine moiety of the NAD+. However, coenzyme reduction or carboxymethylation did alter the stoichiometry of binding and caused the release of approximately one loosely bound DSL-NAD+ from the enzyme. These findings suggest that ionic charge interactions are important in coenzyme binding at the active site.  相似文献   

13.
The major protein component at the surface of merozoites, the infectious form of blood stage malaria parasites, is the merozoite surface protein 1 (MSP-1) complex. In the human malaria parasite Plasmodium falciparum, this complex is generated by proteolytic cleavage of a 190-kDa glycosylphosphatidylinositol-anchored precursor into four major fragments, which remain non-covalently associated. Here, we describe the in vitro reconstitution of the MSP-1 complex of P. falciparum strain 3D7 from its heterologously produced subunits. We provide evidence for the arrangement of the subunits within the complex and show how they interact with each other. Our data indicate that the conformation assumed by the reassembled complex as well as by the heterologously produced 190-kDa precursor corresponds to the native one. Based on these results we propose a first structural model for the MSP-1 complex. Together with access to faithfully produced material, this information will advance further structure-function studies of MSP-1 that plays an essential role during invasion of erythrocytes by the parasite and that is considered a promising candidate for a malaria vaccine.  相似文献   

14.
The crystal structure of heterotetrameric sarcosine oxidase (TSOX) from Pseudomonas maltophilia has been determined at 1.85 A resolution. TSOX contains three coenzymes (FAD, FMN and NAD+), four different subunits (alpha, 103 kDa; beta, 44 kDa; gamma, 21 kDa; delta, 11 kDa) and catalyzes the oxidation of sarcosine (N-methylglycine) to yield hydrogen peroxide, glycine and formaldehyde. In the presence of tetrahydrofolate, the oxidation of sarcosine is coupled to the formation of 5,10-methylenetetrahydrofolate. The NAD+ and putative folate binding sites are located in the alpha-subunit. The FAD binding site is in the beta-subunit. FMN is bound at the interface of the alpha and beta-subunits. The FAD and FMN rings are separated by a short segment of the beta-subunit with the closest atoms located 7.4 A apart. Sulfite, an inhibitor of oxygen reduction, is bound at the FMN site. 2-Furoate, a competitive inhibitor with respect to sarcosine, is bound at the FAD site. The sarcosine dehydrogenase and 5,10-methylenetetrahydrofolate synthase sites are 35 A apart but connected by a large internal cavity (approximately 10,000 A3). An unexpected zinc ion, coordinated by three cysteine and one histidine side-chains, is bound to the delta-subunit. The N-terminal half of the alpha subunit of TSOX (alphaA) is closely similar to the FAD-binding domain of glutathione reductase but with NAD+ replacing FAD. The C-terminal half of the alpha subunit of TSOX (alphaB) is similar to the C-terminal half of dimethylglycine oxidase and the T-protein of the glycine cleavage system, proteins that bind tetrahydrofolate. The beta-subunit of TSOX is very similar to monomeric sarcosine oxidase. The gamma-subunit is similar to the C-terminal sub-domain of alpha-TSOX. The delta-subunit shows little similarity with any PDB entry. The alphaA domain/beta-subunit sub-structure of TSOX closely resembles the alphabeta dimer of L-proline dehydrogenase, a heteroctameric protein (alphabeta)4 that shows highest overall similarity to TSOX.  相似文献   

15.
The conformation of NAD bound to diphtheria toxin (DT), an ADP-ribosylating enzyme, has been compared to the conformations of NAD(P) bound to 23 distinct NAD(P)-binding oxidoreductase enzymes, whose structures are available in the Brookhaven Protein Data Bank. For the oxidoreductase enzymes, NAD(P) functions as a cofactor in electron transfer, whereas for DT, NAD is a labile substrate in which the N-glycosidic bond between the nicotinamide ring and the N-ribose is cleaved. All NAD(P) conformations were compared by (1) visual inspection of superimposed molecules, (2) RMSD of atomic positions, (3) principal component analysis, and (4) analysis of torsion angles and other conformational parameters. Whereas the majority of oxidoreductase-bound NAD(P) conformations are found to be similar, the conformation of NAD bound to DT is found to be unusual. Distinctive features of the conformation of NAD bound to DT that may be relevant to DT''s function as an ADP-ribosylating enzyme include (1) an unusually short distance between the PN and N1N atoms, reflecting a highly folded conformation for the nicotinamide mononucleotide (NMN) portion of NAD, and (2) a torsion angle chi N approximately 0 degree about the scissile N-glycosidic bond, placing the nicotinamide ring outside of the preferred anti and syn orientations. In NAD bound to DT, the highly folded NMN conformation and torsion angle chi N approximately 0 degree could contribute to catalysis, possibly by orienting the C1''N atom of NAD for nucleophilic attack, or by placing strain on the N-glycosidic bond, which is cleaved by DT. The unusual overall conformation of NAD bound to DT is likely to reflect the structure of DT, which is unusual among NAD(P)-binding enzymes. In DT, the NAD binding site is formed at the junction of two antiparallel beta-sheets. In contrast, although the 24 oxidoreductase enzymes belong to at least six different structural classes, almost all of them bind NAD(P) at the C-terminal end of a parallel beta-sheet. The structural alignments and principal component analysis show that enzymes of the same structural class bind to particularly similar conformations of NAD(P), with few exceptions. The conformation of NAD bound to DT superimposes closely with that of an NAD analogue bound to Pseudomonas exotoxin A, an ADP-ribosylating toxin that is structurally homologous to DT. This suggests that all of the ADP-ribosylating enzymes that are structurally homologous to DT and ETA will bind a highly similar conformation of NAD.  相似文献   

16.
The aquaglyceroporin of Plasmodium falciparum (PfAQP) is a bi-functional channel with permeability for water and solutes. Its functions supposedly are in osmotic protection of parasites and in facilitation of glycerol permeation for glycerolipid biosynthesis. Here, we show PfAQP permeability for the glycolysis-related metabolites methylglyoxal, a cytotoxic byproduct, and dihydroxyacetone, a ketotriose. AQP3, the red cell aquaglyceroporin, also passed dihydroxacetone but excluded methylglyoxal. Proliferation of malaria parasites was inhibited by methylglyoxal with an IC50 around 200 μM. Surprisingly, also dihydroxyacetone, which is an energy source in human cells, was antiproliferative in chloroquine-sensitive and resistant strains with an IC50 around 3 mM. We expressed P. falciparum glyceraldehyde 3-phosphate dehydrogenase (PfGAPDH) to examine whether it is inhibited by either carbonyl compound. Methylglyoxal did not affect PfGAPDH on incubation with 2.5 mM for 20 h. Treatment with 2.5 mM dihydroxyacetone, however, abolished PfGAPDH activity within 6 h. Aquaglyceroporin permeability for glycolytic metabolites may thus be of physiological significance.  相似文献   

17.
Uronate dehydrogenase from Agrobacterium tumefaciens (AtUdh) belongs to the short-chain dehydrogenase/reductase superfamily and catalyzes the oxidation of D-galacturonic acid and D-glucuronic acid with NAD(+) as a cofactor. We have determined the crystal structures of an apo-form of AtUdh, a ternary form in complex with NADH and product (substrate-soaked structure), and an inactive Y136A mutant in complex with NAD(+). The crystal structures suggest AtUdh to be a homohexamer, which has also been observed to be the major form in solution. The monomer contains a Rossmann fold, essential for nucleotide binding and a common feature of the short-chain dehydrogenase/reductase family enzymes. The ternary complex structure reveals a product, D-galactaro-1,5-lactone, which is bound above the nicotinamide ring. This product rearranges in solution to D-galactaro-1,4-lactone as verified by mass spectrometry analysis, which agrees with our previous NMR study. The crystal structure of the mutant with the catalytic residue Tyr-136 substituted with alanine shows changes in the position of Ile-74 and Ser-75. This probably altered the binding of the nicotinamide end of NAD(+), which was not visible in the electron density map. The structures presented provide novel insights into cofactor and substrate binding and the reaction mechanism of AtUdh. This information can be applied to the design of efficient microbial conversion of D-galacturonic acid-based waste materials.  相似文献   

18.
The spatial distribution of four subunits specifically associated to the yeast DNA-dependent RNA polymerase I (RNA pol I) was studied by electron microscopy. A structural model of the native enzyme was determined by cryo-electron microscopy from isolated molecules and was compared with the atomic structure of RNA pol II Delta 4/7, which lacks the specific polypeptides. The two models were aligned and a difference map revealed four additional protein densities present in RNA pol I, which were characterized by immunolabelling. A protruding protein density named stalk was found to contain the RNA pol I-specific subunits A43 and A14. The docking with the atomic structure showed that the stalk protruded from the structure at the same site as the C-terminal domain (CTD) of the largest subunit of RNA pol II. Subunit A49 was placed on top of the clamp whereas subunit A34.5 bound at the entrance of the DNA binding cleft, where it could contact the downstream DNA. The location of the RNA pol I-specific subunits is correlated with their biological activity.  相似文献   

19.
Lipoate scavenging from the human host is essential for malaria parasite survival. Scavenged lipoate is covalently attached to three parasite proteins: the H‐protein and the E2 subunits of branched chain amino acid dehydrogenase (BCDH) and α‐ketoglutarate dehydrogenase (KDH). We show mitochondrial localization for the E2 subunits of BCDH and KDH, similar to previously localized H‐protein, demonstrating that all three lipoylated proteins reside in the parasite mitochondrion. The lipoate ligase 1, LipL1, has been shown to reside in the mitochondrion and it catalyses the lipoylation of the H‐protein; however, we show that LipL1 alone cannot lipoylate BCDH or KDH. A second mitochondrial protein with homology to lipoate ligases, LipL2, does not show ligase activity and is not capable of lipoylating any of the mitochondrial substrates. Instead, BCDH and KDH are lipoylated through a novel mechanism requiring both LipL1 and LipL2. This mechanism is sensitive to redox conditions where BCDH and KDH are exclusively lipoylated under strong reducing conditions in contrast to the H‐protein which is preferentially lipoylated under less reducing conditions. Thus, malaria parasites contain two different routes of mitochondrial lipoylation, an arrangement that has not been described for any other organism.  相似文献   

20.
Transhydrogenase (TH) is a dimeric integral membrane enzyme in mitochondria and prokaryotes that couples proton translocation across a membrane with hydride transfer between NAD(H) and NADP(H) in soluble domains. Crystal structures of the NAD(H) binding alpha1 subunit (domain I) of Rhodospirillum rubrum TH have been determined at 1.8 A resolution in the absence of dinucleotide and at 1.9 A resolution with NADH bound. Each structure contains two domain I dimers in the asymmetric unit (AB and CD); the dimers are intimately associated and related by noncrystallographic 2-fold axes. NADH binds to subunits A and D, consistent with the half-of-the-sites reactivity of the enzyme. The conformation of NADH in subunits A and D is very similar; the nicotinamide is in the anti conformation, the A-face is exposed to solvent, and both N7 and O7 participate in hydrogen bonds. Comparison of subunits A and D to six independent copies of the subunit without bound NADH reveals multiple conformations for residues and loops surrounding the NADH site, indicating flexibility for binding and release of the substrate (product). The NADH-bound structure is also compared to the structures of R. rubrum domain I with NAD bound (PDB code 1F8G) and with NAD bound in complex with domain III of TH (PDB code 1HZZ). The NADH- vs NAD-bound domain I structures reveal conformational differences in conserved residues in the NAD(H) binding site and in dinucleotide conformation that are correlated with the net charge, i.e., oxidation state, of the nicotinamides. The comparisons illustrate how nicotinamide oxidation state can affect the domain I conformation, which is relevant to the hydride transfer step of the overall reaction.  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