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 ternary complex of NAD-dependent formate dehydrogenase (FDH) from the methylotrophic bacterium Pseudomonas sp. 101 (enzyme-NAD-azide) has been crystallised in the space group P2(1)2(1)2(1) with cell dimensions a = 11.60 nm, b = 11.33 nm, c = 6.34 nm. There is 1 dimeric molecule/asymmetric unit. An electron density map was calculated using phases from multiple isomorphous replacement at 0.30 nm resolution. Four heavy atom derivatives were used. The map was improved by solvent flattening and molecular averaging. The atomic model, including 2 x 393 amino acid residues, was refined by the CORELS and PROLSQ packages using data between 1.0 nm and 0.30 nm excluding structure factors less than 1 sigma. The current R factor is 27.1% and the root mean square deviation from ideal bond lengths is 4.2 pm. The FDH subunit is folded into a globular two-domain (coenzyme and catalytic) structure and the active centre and NAD binding site are situated at the domain interface. The beta sheet in the FDH coenzyme binding domain contains an additional beta strand compared to other dehydrogenases. The difference in quaternary structure between FDH and the other dehydrogenases means that FDH constitutes a new subfamily of NAD-dependent dehydrogenases: namely the P-oriented dimer. The FDH nucleotide binding region of the structure is aligned with the three dimensional structures of four other dehydrogenases and the conserved residues are discussed. The amino acid residues which contribute to the active centre and which make contact with NAD have been identified.  相似文献   

2.
The C-terminal domain (C(t)-FDH) of 10-formyltetrahydrofolate dehydrogenase (FDH, ALDH1L1) is an NADP(+)-dependent oxidoreductase and a structural and functional homolog of aldehyde dehydrogenases. Here we report the crystal structures of several C(t)-FDH mutants in which two essential catalytic residues adjacent to the nicotinamide ring of bound NADP(+), Cys-707 and Glu-673, were replaced separately or simultaneously. The replacement of the glutamate with an alanine causes irreversible binding of the coenzyme without any noticeable conformational changes in the vicinity of the nicotinamide ring. Additional replacement of cysteine 707 with an alanine (E673A/C707A double mutant) did not affect this irreversible binding indicating that the lack of the glutamate is solely responsible for the enhanced interaction between the enzyme and the coenzyme. The substitution of the cysteine with an alanine did not affect binding of NADP(+) but resulted in the enzyme lacking the ability to differentiate between the oxidized and reduced coenzyme: unlike the wild-type C(t)-FDH/NADPH complex, in the C707A mutant the position of NADPH is identical to the position of NADP(+) with the nicotinamide ring well ordered within the catalytic center. Thus, whereas the glutamate restricts the affinity for the coenzyme, the cysteine is the sensor of the coenzyme redox state. These conclusions were confirmed by coenzyme binding experiments. Our study further suggests that the binding of the coenzyme is additionally controlled by a long-range communication between the catalytic center and the coenzyme-binding domain and points toward an α-helix involved in the adenine moiety binding as a participant of this communication.  相似文献   

3.
NAD-dependent formate dehydrogenase (FDH1) was isolated from the alpha-proteobacterium Methylobacterium extorquens AM1 under oxic conditions. The enzyme was found to be a heterodimer of two subunits (alpha1beta1) of 107 and 61 kDa, respectively. The purified enzyme contained per mol enzyme approximately 5 mol nonheme iron and acid-labile sulfur, 0.6 mol noncovalently bound FMN, and approximately 1.8 mol tungsten. The genes encoding the two subunits of FDH1 were identified on the M. extorquens AM1 chromosome next to each other in the order fdh1B, fdh1A. Sequence comparisons revealed that the alpha-subunit harbours putative binding motifs for the molybdopterin cofactor and at least one iron-sulfur cluster. Sequence identity was highest to the catalytic subunits of the tungsten- and selenocysteine-containing formate dehydrogenases characterized from Eubacterium acidaminophilum and Moorella thermoacetica (Clostridium thermoaceticum). The beta-subunit of FDH1 contains putative motifs for binding FMN and NAD, as well as an iron-sulfur cluster binding motif. The beta-subunit appears to be a fusion protein with its N-terminal domain related to NuoE-like subunits and its C-terminal domain related to NuoF-like subunits of known NADH-ubiquinone oxidoreductases.  相似文献   

4.
An homology model of Candida methylica formate dehydrogenase (cmFDH) was constructed based on the Pseudomonas sp. 101 formate dehydrogenase (psFDH) structure. An aspartic acid residue in the model was predicted to interact with the adenine ribose of the NAD cofactor, in common with many NAD-dependent oxoreductases. Replacement of this aspartic acid residue by serine in cmFDH removed the absolute requirement for NAD over NADP shown by the wild type enzyme. Taken with similar results shown by d- and l-lactate dehydrogenases, this suggests that an aspartic acid in this position is a major determinant of coenzyme specificity in NAD/NADP-dependent dehydrogenases.  相似文献   

5.
Sequence alignment shows that residue Arg 284 (according to the numbering of the residues in formate dehydrogenase, FDH, from the methylotrophic bacterium Pseudomonas sp. 101) is conserved in NAD-dependent FDHs and D-specific 2-hydroxyacid dehydrogenases. Mutation of Arg 284 to glutamine and alanine results in a change of the catalytic, thermodynamic and spectral properties of FDH. In comparison to wild-type, the affinity of the mutants for the substrate (K(formate)m) or the transition state analogue (K(azide)i) decreases and correlates with the ability of the side chain of residue 284 to form H-bonds. In contrast, the affinity for the coenzyme (K(NAD)d or K(NAD)m) is either not affected or increases and correlates inversely with the partial positive charge of the side chain. The temperature dependence of circular dichroism (CD) spectra of the wild-type FDH and its Ala mutant has been studied over the 5-90 degrees C temperature range. Both proteins reveal regions of enhanced conformational mobility at the predenaturing temperatures (40-55 degrees C) associated with a change of enzyme kinetic parameters and a co-operative transition around 55-70 degrees C which is followed by the loss of enzyme activity. CD spectra of the wild-type and mutant proteins were deconvoluted and contributions from various types of secondary structure estimated. It is shown that the co-operative transition at 55-70 degrees C in the FDH protein globule is triggered by a loss of alpha-helical secondary structure. The results confirm the conclusion, from the crystal structures, that Arg 284 is directly involved in substrate binding. In addition this residue seems to exert a major structural role by supporting the catalytic conformation of the enzyme active centre.  相似文献   

6.
From the amino acid sequences of the three known fragments of the NAD-specific glutamate dehydrogenase of Neurospora crassa, the secondary structures have been predicted from the rules of Chou and Fasman (Chou, P.Y., and Fasman, G.D. (1979) Biophys. J. 26, 367-384). Comparison of these structures with those calculated for bovine glutamate dehydrogenase has shown that in the regions of homologous sequences containing identified functional regions, there is considerable homology of structure. From these predictions, it has been possible to identify a putative coenzyme-binding domain in the COOH-terminal part of the molecule similar to those of various NAD-specific dehydrogenases. Residues whose modification alters coenzyme binding are located in the putative coenzyme binding domain. The major sites of tryptic cleavage of the native enzyme, described in an accompanying paper (Haberland, M.E., Chen, C.-W., and Smith, E.L. (1980) J. Biol. Chem. 255, 7993-8000), are in regions of random coil structure.  相似文献   

7.
The aldehyde dehydrogenases (ALDHs) are a superfamily of multimeric enzymes which catalyse the oxidation of a broad range of aldehydes into their corresponding carboxylic acids with the reduction of their cofactor, NAD or NADP, into NADH or NADPH. At present, the only known structures concern NAD-dependent ALDHs. Three structures are available in the Protein Data Bank: two are tetrameric and the other is a dimer. We solved by molecular replacement the first structure of an NADP-dependent ALDH isolated from Streptococcus mutans, in its apo form and holo form in complex with NADP, at 1.8 and 2.6 A resolution, respectively. Although the protein sequence shares only approximately 30 % identity with the other solved tetrameric ALDHs, the structures are very similar. However, a large local conformational change in the region surrounding the 2' phosphate group of the adenosine moiety is observed when the enzyme binds NADP, in contrast to the NAD-dependent ALDHs.Structure and sequence analyses reveal several properties. A small number of residues seem to determine the oligomeric state. Likewise, the nature (charge and volume) of the residue at position 180 (Thr in ALDH from S. mutans) determines the cofactor specificity in comparison with the structures of NAD-dependent ALDHs. The presence of a hydrogen bond network around the cofactor not only allows it to bind to the enzyme but also directs the side-chains in a correct orientation for the catalytic reaction to take place. Moreover, a specific part of this network appears to be important in substrate binding. Since the enzyme oxidises the same substrate, glyceraldehyde-3-phosphate (G3P), as NAD-dependent phosphorylating glyceraldehyde-3-phosphate dehydrogenases (GAPDH), the active site of GAPDH was compared with that of the S. mutans ALDH. It was found that Arg103, Arg283 and Asp440 might be key residues for substrate binding.  相似文献   

8.
FDH (10-formyltetrahydrofolate dehydrogenase, Aldh1L1, EC 1.5.1.6) converts 10-formyltetrahydrofolate (10-formyl-THF) to tetrahydrofolate and CO(2) in a NADP(+)-dependent reaction. It is a tetramer of four identical 902 amino acid residue subunits. The protein subunit is a product of a natural fusion of three unrelated genes and consists of three distinct domains. The N-terminal domain of FDH (residues 1-310) carries the folate binding site and shares sequence homology and structural topology with other enzymes utilizing 10-formyl-THF as a substrate. In vitro it functions as 10-formyl-THF hydrolase, and evidence indicate that this activity is a part of the overall FDH mechanism. The C-terminal domain of FDH (residues 400-902) originated from an aldehyde dehydrogenase-related gene and is capable of oxidation of short-chain aldehydes to corresponding acids. Similar to classes 1 and 2 aldehyde dehydrogenases, this domain exists as a tetramer and defines the oligomeric structure of the full-length enzyme. The two catalytic domains are connected by an intermediate linker (residues 311-399), which is a structural and functional homolog of carrier proteins possessing a 4'-phosphopantetheine prosthetic group. In the FDH mechanism, the intermediate linker domain transfers a formyl, covalently attached to the sulfhydryl group of the phosphopantetheine arm, from the N-terminal domain to the C-terminal domain. The overall FDH mechanism is a coupling of two sequential reactions, a hydrolase and a formyl dehydrogenase, bridged by a substrate transfer step. In this mechanism, one domain provides the folate binding site and a hydrolase catalytic center to remove the formyl group from the folate substrate, another provides a transfer vehicle between catalytic centers and the third one contributes the dehydrogenase machinery further oxidizing formyl to CO(2).  相似文献   

9.
Prolyl oligopeptidase, which is involved in memory disorders, is a member of a new family of serine peptidases. In addition to the peptidase domain, the enzyme contains a beta-propeller, which excludes large peptides from the active site. The enzyme is inhibited with thiol reagents, possibly by reacting with Cys-255 located close to the substrate binding site. This assumption was tested with the Cys-255 --> Thr, Cys-255 --> Ala, and Cys-255 --> Ser variants of prolyl oligopeptidase. In contrast to the wild type enzyme, the Cys-255 --> Thr variant was not inhibited with N-ethylmaleimide, indicating that Cys-255, of the 16 free cysteine residues, exclusively accounts for the enzyme inhibition. Unlike the wild type enzyme that showed a doubly bell-shaped pH rate profile, the modified enzyme displayed a single bell-shaped pH dependence with benzyloxycarbonyl-Gly-Pro-naphthylamide. It was the high pH form of the enzyme that virtually disappeared with all three enzyme variants. A substantial reduction was also observed in k(cat)/K(m) for the aminobenzoyl-Ser-Pro-Phe(NO(2))-Ala-OH substrate. The high pK(a) (9.77) of Cys-255 determined by titration with N-ethylmaleimide excluded the possibility that ionization of the thiol group was responsible for generation of the two active enzyme forms. The impaired activity of the enzyme variants could be rationalized in terms of weaker binding, which manifests itself in high K(m) for substrates and high K(i) for inhibitors, like benzyloxycarbonyl-Gly-Pro-OH and aminobenzoyl-Ser-d-Pro-Phe(NO(2))-Ala-OH. It was concluded that, besides selecting substrates by size, the beta-propeller domain containing Cys-255 remarkably contributed to catalysis of the peptidase domain.  相似文献   

10.
Sanghani PC  Bosron WF  Hurley TD 《Biochemistry》2002,41(51):15189-15194
Human glutathione-dependent formaldehyde dehydrogenase plays an important role in the metabolism of glutathione adducts such as S-(hydroxymethyl)glutathione and S-nitrosoglutathione. The role of specific active site residues in binding these physiologically important substrates and the structural changes during the catalytic cycle of glutathione-dependent formaldehyde dehydrogenase was examined by determining the crystal structure of a ternary complex with S-(hydroxymethyl)glutathione and the reduced coenzyme to 2.6 A resolution. The formation of the ternary complex caused the movement of the catalytic domain toward the coenzyme-binding domain. This represents the first observation of domain closure in glutathione-dependent formaldehyde dehydrogenase in response to substrate binding. A water molecule adjacent to the 2'-ribose hydroxyl of NADH suggests that the alcohol proton is relayed to solvent directly from the coenzyme, rather than through the action of the terminal histidine residue as observed in the proton relay system for class I alcohol dehydrogenases. S-(Hydroxymethyl)glutathione is directly coordinated to the active site zinc and forms interactions with the highly conserved residues Arg114, Asp55, Glu57, and Thr46. The active site zinc has a tetrahedral coordination environment with Cys44, His66, and Cys173 as the three protein ligands in addition to S-(hydroxymethyl)glutathione. This is in contrast to zinc coordination in the binary coenzyme complex where all of the ligands were contributed by the enzyme and included Glu67 as the fourth protein ligand. This change in zinc coordination is accomplished by an approximately 2.3 A movement of the catalytic zinc.  相似文献   

11.
Three heavy atom isomorphous derivatives were used for the X-ray analysis of the holo form of NAD-dependent bacterial formate dehydrogenase (ternary complex enzyme-NAD-azide) at 3.0 A resolution. The enzyme subunit contains a catalytic and a coenzyme binding domain, with the active centre and the coenzyme binding site in the cleft between the domains. The polypeptide chain's fold and the position of 393 C alpha-atoms were determined. The secondary structure of the formate dehydrogenase was resolved. The structure of the NAD-binding domain is shown to be similar to that of other NAD-dependent enzymes.  相似文献   

12.
The gene of the NAD-dependent formate dehydrogenase (FDH) from the yeast Candida boidinii was cloned by PCR using genomic DNA as a template. Expression of the gene in Escherichia coli yielded functional FDH with about 20% of the soluble cell protein. To confirm the hypothesis of a thiol-coupled inactivation process, both cysteine residues in the primary structure of the enzyme have been exchanged by site-directed mutagenesis using a homology model based on the 3D structure of FDH from Pseudomonas sp. 101 and from related dehydrogenases. Compared to the wt enzyme, most of the mutants were significantly more stable towards oxidative stress in the presence of Cu(II) ions, whereas the temperature optima and kinetic constants of the enzymatic reaction are not significantly altered by the mutations. Determination of the Tm values revealed that the stability at temperatures above 50 degrees C is optimal for the native and the recombinant wt enzyme (Tm 57 degrees C), whereas the Tm values of the mutant enzymes vary in the range 44-52 degrees C. Best results in initial tests concerning the application of the enzyme for regeneration of NADH in biotransformation of trimethyl pyruvate to Ltert leucine were obtained with two mutants, FDHC23S and FDHC23S/C262A, which are significantly more stable than the wt enzyme.  相似文献   

13.
IL-2R on activated lymphocytes contain the Tac protein. As part of an effort to characterize this molecule, we examined the structure-activity relationship for each of its 12 Cys residues. A preliminary map of intramolecular disulfide bonding was derived by analysis of cystine-linked enzymatic fragments of the Tac protein. The results indicated that disulfide bonds linked Cys-3 with Cys-147, Cys-131 with Cys-163, and Cys-28,30 with Cys-59,61. The contribution of the Cys residues to an active protein conformation was tested by site-specific mutagenesis, followed by expression of the modified molecules in murine L cells. The results indicated that Cys-192 and -225 could be replaced without affecting ligand binding. In contrast, modification of any of the other 10 Cys residues, either singly or in combinations corresponding to the predicted disulfide bonds, greatly reduced the ability of the corresponding protein to bind IL-2 or either of two mAb (anti-Tac and 7G7/B6) which recognize the Tac protein. Each of the latter mutations also interfered with the molecule's post-translational modification and cell-surface expression. Consistent with these findings, transfection of the L cells with vectors containing truncated Tac cDNA inserts resulted in secretion of Tac fragments capable of ligand binding when the polypeptide chains terminated after Cys-163 (the 10th Cys residue in the full length molecule), but resulted in inactive fragments of Tac which were poorly secreted when they terminated before Cys-163. These findings emphasize the remarkable sensitivity of the active conformation of the Tac molecule to each of the postulated intramolecular disulfide bonds.  相似文献   

14.
Short-chain dehydrogenases/reductases form a large, evolutionarily old family of NAD(P)(H)-dependent enzymes with over 60 genes found in the human genome. Despite low levels of sequence identity (often 10-30%), the three-dimensional structures display a highly similar alpha/beta folding pattern. We have analyzed the role of several conserved residues regarding folding, stability, steady-state kinetics, and coenzyme binding using bacterial 3beta/17beta-hydroxysteroid dehydrogenase and selected mutants. Structure determination of the wild-type enzyme at 1.2-A resolution by x-ray crystallography and docking analysis was used to interpret the biochemical data. Enzyme kinetic data from mutagenetic replacements emphasize the critical role of residues Thr-12, Asp-60, Asn-86, Asn-87, and Ala-88 in coenzyme binding and catalysis. The data also demonstrate essential interactions of Asn-111 with active site residues. A general role of its side chain interactions for maintenance of the active site configuration to build up a proton relay system is proposed. This extends the previously recognized catalytic triad of Ser-Tyr-Lys residues to form a tetrad of Asn-Ser-Tyr-Lys in the majority of characterized short-chain dehydrogenases/reductase enzymes.  相似文献   

15.
The gene encoding alanine dehydrogenase (EC 1.4.1.1) from a mesophile, Bacillus sphaericus, was cloned, and its complete DNA sequence was determined. In addition, the same gene from a moderate thermophile, B. stearothermophilus, was analyzed in a similar manner. Large parts of the two translated amino acid sequences were confirmed by automated Edman degradation of tryptic peptide fragments. Each alanine dehydrogenase gene consists of a 1116-bp open reading frame and encodes 372 amino acid residues corresponding to the subunit (Mr = 39,500-40,000) of the hexameric enzyme. The similarity of amino acid sequence between the two alanine dehydrogenases with distinct thermostabilities is very high (greater than 70%). The nonidentical residues are clustered in a few regions with relatively short length, which may correlate with the difference in thermal stability of the enzymes. Homology search of the primary structures of both alanine dehydrogenases with those of other pyridine nucleotide-dependent oxidoreductases revealed significant sequence similarity in the regions containing the coenzyme binding domain. Interestingly, several catalytically important residues in lactate and malate dehydrogenases are conserved in the primary structure of alanine dehydrogenases at matched positions with similar mutual distances.  相似文献   

16.
Direct transfer of NADPH between two NADP-dependent dehydrogenases, isocitrate dehydrogenase and glutamate dehydrogenase, has been investigated. These enzymes have opposite stereospecificity for hydrogen transfer to the coenzyme. In contrast with the general direct-transfer mechanism postulated for NAD-dependent dehydrogenases [Srivastava & Bernhard (1986) Science 234, 1081-1086], no evidence for direct transfer in either direction was found for these NADP-dependent dehydrogenases.  相似文献   

17.
A novel NADH dehydrogenase (NADH-dh) involving FAD as coenzyme, distinct from NADPH dehydrogenase (NADPH-dh, old yellow enzyme, EC 1.6.99.1), was found in the same cytoplasmic fraction of Gluconobacter strains. Conventional artificial electron acceptors were more effective than molecular oxygen in the NADH-dh reaction. NADH-dh did not appear to be identical with any previously described flavoproteins, although the N-terminal amino acid sequence showed 100% similarity with a non-heme chloroperoxidase. The N-terminal amino acid sequence of NADPH-dh matched 100% a putative oxidoreductase containing the old yellow enzyme-like FMN-binding domain. NADH-dh might function to regenerate NAD coupling with NAD-dependent dehydrogenases in the cytoplasm of Gluconobacter strains.  相似文献   

18.
The structure of the complex between Cys-46-carboxymethylated horse liver alcohol dehydrogenase (CM-LADH) and reduced nicotinamide adenine dinucleotide (NADH) has been determined by X-ray analysis. The complex represents NADH binding to the orthorhombic, "open" conformation of the enzyme. Coenzyme binding here induces a local structural change in the peptide loop 293-297, but there is no domain rotation, as observed for the "closed" conformation of the protein. This local movement of a few residues in the loop is sufficient to trap the nicotinamide ring of NADH within the active-site area close to a productive binding position. The carboxymethyl group on the zinc ligand cysteine-46 is oriented between the pyrophosphate bridge of NADH and the guanidinium group of arginine-369 and can occupy this position because the coenzyme binding cleft remains open and unchanged upon coenzyme binding. The zinc coordination sphere is distorted, and the position of the metal atom is shifted 1 A compared to native unliganded LADH. The distance between the zinc ion and the sulfur of the alkylated cysteine residue is of the order of 3 A. Alkylation experiments were performed at 0.15 and 10 mM iodoacetate, and peptide maps were examined. Gentle treatment with reagent yields an enzyme product which is substituted at only one of the two zinc binding sites per subunit of LADH (Cys-46). This enzyme species maintains its structural integrity; it binds coenzyme which induces conformational changes resolved into two steps. Thus, in addition to the orthorhombic complex, a crystalline NADH complex in the closed conformation of CM-LADH was obtained. These crystals showed enzymic activity, and single crystals were analyzed with microspectrophotometric methods. Formation of the stable crystalline abortive complex between CM-LADH-NAD+ and 4-trans-(N,N-dimethylamino)cinnamaldehyde (DACA) could be observed upon addition of excess aldehyde to the closed complex of CM-LADH-NADH. The CM-LADH-NAD+-DACA complex is characterized by an intense absorption band with a lambda max at 456 nm which corresponds to a shift in the spectrum of free DACA of approximately 60 nm. At the higher concentration of iodoacetate, three of the cysteine ligands to the second zinc atom (Cys-100, -103, and -111) are alkylated in addition to Cys-46. This enzyme product rapidly denatures and cannot be crystallized under our conditions. This is an experimental indication that the intact noncatalytic zinc binding site contributes to the structural stability of the protein.  相似文献   

19.
The structure of apo-glyceraldehyde-3-phosphate dehydrogenase (GAPDHase) from Bacillus stearothermophilus has been refined using a restrained least-squares method. The final crystallographic R-factor is 0.177 for all 53,315 reflections between 7.0 and 2.5 A. The resulting model has been analysed with respect to lattice interactions, molecular symmetry, temperature factors and solvent structure showing that, apart from local deviations due to intermolecular contact, the molecule exhibits a very high degree of local 222 symmetry. Analysis of differences between the structure of apo-GAPDHase and the previously refined holo-GAPDHase at 1.8 A resolution reveals details of conformational change in the enzyme induced by cofactor binding. The change, which was previously described as a rigid-body rotation of the coenzyme-binding domain with respect to the catalytic domain, is of more complex nature and involves relative shifts of several structural elements in the coenzyme-binding domain and some small changes in the catalytic domain. A possible mechanism of this conformational change is proposed based on the comparison of the refined structures and model-building studies. According to this mechanism, the adenosine moiety of NAD can initially bind to the protein in the apo-enzyme conformation. Several attractive interactions resulting from the initial binding of the coenzyme trigger conformational changes in the molecule of GAPDHase that: (1) create the productive nicotinamide-moiety binding site; (2) improve enzyme-coenzyme interactions at the adenosine moiety; (3) modify the active site to optimize the positioning of catalytic residues and ion-binding sites. Implications of the proposed mechanism for existing experimental data on binding of NAD analogues to GAPDHase are discussed.  相似文献   

20.
We recently identified the gold compound aurothiomalate (ATM) as a potent inhibitor of the Phox and Bem1p (PB1)-PB1 domain interaction between protein kinase C (PKC) iota and the adaptor molecule Par6. ATM also blocks oncogenic PKCiota signaling and the transformed growth of human lung cancer cells. Here we demonstrate that ATM is a highly selective inhibitor of PB1-PB1 domain interactions between PKCiota and the two adaptors Par6 and p62. ATM has no appreciable inhibitory effect on other PB1-PB1 domain interactions, including p62-p62, p62-NBR1, and MEKK3-MEK5 interactions. ATM can form thio-gold adducts with cysteine residues on target proteins. Interestingly, PKCiota (and PKCzeta) contains a unique cysteine residue, Cys-69, within its PB1 domain that is not present in other PB1 domain containing proteins. Cys-69 resides within the OPR, PC, and AID motif of PKCiota at the binding interface between PKCiota and Par6 where it interacts with Arg-28 on Par6. Molecular modeling predicts formation of a cysteinyl-aurothiomalate adduct at Cys-69 that protrudes into the binding cleft normally occupied by Par6, providing a plausible structural explanation for ATM inhibition. Mutation of Cys-69 of PKCiota to isoleucine or valine, residues frequently found at this position in other PB1 domains, has little or no effect on the affinity of PKCiota for Par6 but confers resistance to ATM-mediated inhibition of Par6 binding. Expression of the PKCiota C69I mutant in human non-small cell lung cancer cells confers resistance to the inhibitory effects of ATM on transformed growth. We conclude that ATM inhibits cellular transformation by selectively targeting Cys-69 within the PB1 domain of PKCiota.  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