首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Payne VA  Chang YT  Loew GH 《Proteins》1999,37(2):176-190
The CYP2C subfamily of human liver P450 isozymes is of major importance in drug metabolism. The most abundant 2C isozyme, CYP2C9, regioselectively hydroxylates a wide variety of substrates. A major obstacle to understanding this specificity in human CYP2C9 is the absence of a 3D structure. A 3D model of CYP2C9 was built, assessed, and used to characterize explicit enzyme-substrate complexes using methods previously developed in our laboratory. The 3D model was assessed by determining its stability to unconstrained molecular dynamics and by comparison of specific properties with those of known protein structures. The CYP2C9 model was then used to characterize explicit enzyme complexes with three structurally and chemically diverse substrates: (S)-naproxen, phenytoin, and progesterone. Each substrate was found to bind to the enzyme with a favorable interaction energy and to remain in the binding site during unconstrained molecular dynamics. Moreover, the mode of binding of each substrate led to calculated preferred hydroxylation sites consistent with experiment. Binding-site residues identified for the models included Arg 105 and Arg97 as key cationic residues, as well as Asn 202, Asp 293, Pro 101, Leu 102, Gly 296, and Phe 476. Site-specific mutations are proposed for further integrated computational and experimental study.  相似文献   

2.
The silverleaf whitefly (Bemisia argentifolii, Bellows and Perring) accumulates sorbitol as a thermoprotectant in response to elevated temperature. Sorbitol synthesis in this insect is catalyzed by an unconventional ketose reductase (KR) that uses NADPH to reduce fructose. A cDNA encoding the NADPH-KR from adult B. argentifolii was cloned and sequenced to determine the primary structure of this enzyme. The cDNA encoded a protein of 352 amino acids with a calculated molecular mass of 38.2 kDa. The deduced amino acid sequence of the cDNA shared 60% identity with sheep NAD(+)-dependent sorbitol dehydrogenase (SDH). Residues in SDH involved in substrate binding were conserved in the whitefly NADPH-KR. An important structural difference between the whitefly NADPH-KR and NAD(+)-SDHs occurred in the nucleotide-binding site. The Asp residue that coordinates the adenosyl ribose hydroxyls in NAD(+)-dependent dehydrogenases (including NAD(+)-SDH), was replaced by an Ala in the whitefly NADPH-KR. The whitefly NADPH-KR also contained two neutral to Arg substitutions within four residues of the Asp to Ala substitution. Molecular modeling indicated that addition of the Arg residues and loss of the Asp decreased the electric potential of the adenosine ribose-binding pocket, creating an environment favorable for NADPH-binding. Because of the ability to use NADPH, the whitefly NADPH-KR synthesizes sorbitol under physiological conditions, unlike NAD(+)-SDHs, which function in sorbitol catabolism.  相似文献   

3.
Ribitol dehydrogenase from Zymomonas mobilis (ZmRDH) catalyzes the conversion of ribitol to d-ribulose and concomitantly reduces NAD(P)(+) to NAD(P)H. A systematic approach involving an initial sequence alignment-based residue screening, followed by a homology model-based screening and site-directed mutagenesis of the screened residues, was used to study the molecular determinants of the cofactor specificity of ZmRDH. A homologous conserved amino acid, Ser156, in the substrate-binding pocket of the wild-type ZmRDH was identified as an important residue affecting the cofactor specificity of ZmRDH. Further insights into the function of the Ser156 residue were obtained by substituting it with other hydrophobic nonpolar or polar amino acids. Substituting Ser156 with the negatively charged amino acids (Asp and Glu) altered the cofactor specificity of ZmRDH toward NAD(+) (S156D, [k(cat)/K(m)(,NAD)]/[k(cat)/K(m)(,NADP)] = 10.9, where K(m)(,NAD) is the K(m) for NAD(+) and K(m)(,NADP) is the K(m) for NADP(+)). In contrast, the mutants containing positively charged amino acids (His, Lys, or Arg) at position 156 showed a higher efficiency with NADP(+) as the cofactor (S156H, [k(cat)/K(m)(,NAD)]/[k(cat)/K(m)(,NADP)] = 0.11). These data, in addition to those of molecular dynamics and isothermal titration calorimetry studies, suggest that the cofactor specificity of ZmRDH can be modulated by manipulating the amino acid residue at position 156.  相似文献   

4.
Pletnev VZ  Weeks CM  Duax WL 《Proteins》2004,57(2):294-301
The dominant role of long-range electrostatic interatomic interactions in nicotinamide adenine dinucleotide/nicotinamide adenine dinucleotide phosphate (NAD/NADP) cofactor recognition has been shown for enzymes of the short-chain oxidoreductase (SCOR) family. An estimation of cofactor preference based only on the contribution of the electrostatic energy term to the total energy of enzyme-cofactor interaction has been tested for approximately 40 known three-dimensional (3D) crystal complexes and approximately 330 SCOR enzymes, with cofactor preference predicted by the presence of Asp or Arg recognition residues at specific 3D positions in the beta2alpha3 loop (Duax et al., Proteins 2003;53:931-943). The results obtained were found to be consistent with approximately 90% reliable cofactor assignments for those subsets. The procedure was then applied to approximately 170 SCOR enzymes with completely uncertain NAD/NADP dependence, due to the lack of Asp and Arg marker residues. The proposed 3D electrostatic approach for cofactor assignment ("3D_DeltaE(el)") has been implemented in an automatic screening procedure, and together with the use of marker residues proposed earlier (Duax et al., Proteins 2003;53:931-943), increases the level of reliable predictions for the putative SCORs from approximately 70% to approximately 90%. It is expected to be applicable for any NAD/NADP-dependent enzyme subset having at least 25-30% sequence identity, with at least one enzyme of known 3D crystal structure.  相似文献   

5.
The crystal structure of the human cystatin C (hCC) dimer revealed that a stable twofold-symmetric dimer was formed via 3D domain swapping. Domain swapping with the need for near-complete unfolding has been proposed as a possible route for amyloid fibril initiation. Thus, the interesting interactions that occur between the two molecules may be important for the further aggregation of the protein. In this work, we performed steered molecular dynamics (SMD) simulations to investigate the dissociation of the β2 and β3 strands in the hCC dimer. The energy changes observed during the SMD simulations showed that electrostatic interactions were the dominant interactions involved in stabilizing the two parts of the dimer during the early stages of SMD simulation, whereas van der Waals (VDW) interactions and electrostatic interactions were equally matched during the latter stages. Furthermore, our data indicated that the two parts of the dimer are stabilized by intermolecular hydrogen bonds among the residues Arg51 (β2), Gln48 (β2), Asp65 (β3), and Glu67 (β3), salt bridges among the residues Arg53 (β2), Arg51 (β2), and Asp65 (β3), and VDW interactions among the residues Gln48 (β2), Arg51 (β2), Glu67 (β3), Asp65 (β3), Phe63 (β3), and Asn61 (β3). The residues Gln48 (β2), Arg51 (β2), Asp65 (β3) and Glu67 (β3) appear to be crucial, as they play important roles in both electrostatic and VDW interactions. Thus, the present study determined the key residues involved in the stabilization of the domain-swapped dimer structure, and also provided molecular-level insights into the dissociation process of the hCC dimer.  相似文献   

6.
We purified a novel ADP-ribosyltransferase produced by a Clostridium limosum strain isolated from a lung abscess and compared the exoenzyme with Clostridium botulinum ADP-ribosyltransferase C3. The C. limosum exoenzyme has a molecular weight of about 25,000 and a pI of 10.3. The specific activity of the ADP-ribosyltransferase is 3.1 nmol/mg/min with a Km for NAD of 0.3 microM. Partial amino acid sequence analysis of the tryptic peptides revealed about 70% homology with C3. The novel exoenzyme modifies selectively the small GTP-binding proteins of the rho family in human platelet membranes presumably at the same amino acid (asparagine 41) as known for C3. Recombinant rhoA and rhoB serve as substrates for C3 and the C. limosum exoenzyme. Whereas recombinant rac1 protein is only marginally ADP-ribosylated by C3 or by the C. limosum exoenzyme in the absence of detergent, in the presence of 0.01% sodium dodecyl sulfate rac1 is modified by C3 but not by the C. limosum exoenzyme. Recombinant CDC42Hs protein is a poor substrate for C. limosum exoenzyme and is even less modified by C3. The C. limosum exoenzyme is auto-ADP-ribosylated in the presence of 0.01% sodium dodecyl sulfate by forming an ADP-ribose protein bond highly stable toward hydroxylamine. The data indicate that ADP-ribosylation of small GTP-binding proteins of the rho family is not unique to C. botulinum C3 ADP-ribosyltransferase but is also catalyzed by a C3-related exoenzyme from C. limosum.  相似文献   

7.
Clostridium botulinum C3 is the prototype of the family of the C3-like transferases that ADP-ribosylate exclusively RhoA, -B and -C. The ADP-ribose at Asn-41 results in functional inactivation of Rho reflected by disaggregation of the actin cytoskeleton. We report on a new C3-like transferase produced by a pathogenic Staphylococcus aureus strain. The transferase designated C3(Stau) was cloned from the genomic DNA. At the amino acid level, C3(Stau) revealed an identity of 35% to C3 from C. botulinum and Clostridium limosum exoenzyme, respectively, and of 78% to EDIN from S. aureus. In addition to RhoA, which is the target of the other C3-like transferases, C3(Stau) modified RhoE and Rnd3. RhoE was ADP-ribosylated at Asn-44, which is equivalent to Asn-41 of RhoA. RhoE and Rnd3 are members of the Rho subfamily, which are deficient in intrinsic GTPase activity and possess a RhoA antagonistic cell function. The protein substrate specificity found with recombinant Rho proteins was corroborated by expression of RhoE in Xenopus laevis oocytes showing that RhoE was also modified in vivo by C3(Stau) but not by C3 from C. botulinum. The poor cell accessibility of C3(Stau) was overcome by generation of a chimeric toxin recruiting the cell entry machinery of C. botulinum C2 toxin. The chimeric C3(Stau) caused the same morphological and cytoskeletal changes as the chimeric C. botulinum C3. C3(Stau) is a new member of the family of the C3-like transferases but is also the prototype of a subfamily of RhoE/Rnd modifying transferases.  相似文献   

8.
Certain pathogenic trypanosomatids are highly dependent on glycolysis for ATP production, and hence their glycolytic enzymes, including glycerol-3-phosphate dehydrogenase (GPDH), are considered attractive drug targets. The ternary complex structure of Leishmania mexicana GPDH (LmGPDH) with dihydroxyacetone phosphate (DHAP) and NAD(+) was determined to 1.9A resolution as a further step towards understanding this enzyme's mode of action. When compared with the apo and binary complex structures, the ternary complex structure shows an 11 degrees hinge-bending motion of the C-terminal domain with respect to the N-terminal domain. In addition, residues in the C-terminal domain involved in catalysis or substrates binding show significant movements and a previously invisible five-residue loop region becomes well ordered and participates in NAD(+) binding. Unexpectedly, DHAP and NAD(+) appear to form a covalent bond, producing an adduct in the active site of LmGPDH. Modeling a ternary complex glycerol 3-phosphate (G3P) and NAD(+) with LmGPDH identified ten active site residues that are highly conserved among all GPDHs. Two lysine residues, Lys125 and Lys210, that are presumed to be critical in catalysis, were mutated resulting in greatly reduced catalytic activity. Comparison with other structurally related enzymes found by the program DALI suggested Lys210 as a key catalytic residue, which is located on a structurally conserved alpha-helix. From the results of site-directed mutagenesis, molecular modeling and comparison with related dehydrogenases, a catalytic mechanism of LmGPDH and a possible evolutionary scenario of this group of dehydrogenases are proposed.  相似文献   

9.
Water dynamics clue to key residues in protein folding   总被引:1,自引:0,他引:1  
A computational method independent of experimental protein structure information is proposed to recognize key residues in protein folding, from the study of hydration water dynamics. Based on all-atom molecular dynamics simulation, two key residues are recognized with distinct water dynamical behavior in a folding process of the Trp-cage protein. The identified key residues are shown to play an essential role in both 3D structure and hydrophobic-induced collapse. With observations on hydration water dynamics around key residues, a dynamical pathway of folding can be interpreted.  相似文献   

10.
Yun M  Park CG  Kim JY  Park HW 《Biochemistry》2000,39(35):10702-10710
The crystal structures of gyceraldehyde 3-phosphate dehydrogenase (GAPDH) from Escherichia coli have been determined in three different enzymatic states, NAD(+)-free, NAD(+)-bound, and hemiacetal intermediate. The NAD(+)-free structure reported here has been determined from monoclinic and tetragonal crystal forms. The conformational changes in GAPDH induced by cofactor binding are limited to the residues that bind the adenine moiety of NAD(+). Glyceraldehyde 3-phosphate (GAP), the substrate of GAPDH, binds to the enzyme with its C3 phosphate in a hydrophilic pocket, called the "new P(i)" site, which is different from the originally proposed binding site for inorganic phosphate. This observed location of the C3 phosphate is consistent with the flip-flop model proposed for the enzyme mechanism [Skarzynski, T., Moody, P. C., and Wonacott, A. J. (1987) J. Mol. Biol. 193, 171-187]. Via incorporation of the new P(i) site in this model, it is now proposed that the C3 phosphate of GAP initially binds at the new P(i) site and then flips to the P(s) site before hydride transfer. A superposition of NAD(+)-bound and hemiacetal intermediate structures reveals an interaction between the hydroxyl oxygen at the hemiacetal C1 of GAP and the nicotinamide ring. This finding suggests that the cofactor NAD(+) may stabilize the transition state oxyanion of the hemiacetal intermediate in support of the flip-flop model for GAP binding.  相似文献   

11.
Cui M  Shen J  Briggs JM  Luo X  Tan X  Jiang H  Chen K  Ji R 《Biophysical journal》2001,80(4):1659-1669
The association of the scorpion toxin Lq2 and a potassium ion (K(+)) channel has been studied using the Brownian dynamics (BD) simulation method. All of the 22 available structures of Lq2 in the Brookhaven Protein Data Bank (PDB) determined by NMR were considered during the simulation, which indicated that the conformation of Lq2 affects the binding between the two proteins significantly. Among the 22 structures of Lq2, only 4 structures dock in the binding site of the K(+) channel with a high probability and favorable electrostatic interactions. From the 4 candidates of the Lq2-K(+) channel binding models, we identified a good three-dimensional model of Lq2-K(+) channel complex through triplet contact analysis, electrostatic interaction energy estimation by BD simulation and structural refinement by molecular mechanics. Lq2 locates around the extracellular mouth of the K(+) channel and contacts the K(+) channel using its beta-sheet rather than its alpha-helix. Lys27, a conserved amino acid in the scorpion toxins, plugs the pore of the K(+) channel and forms three hydrogen bonds with the conserved residues Tyr78(A-C) and two hydrophobic contacts with Gly79 of the K(+) channel. In addition, eight hydrogen-bonds are formed between residues Arg25, Cys28, Lys31, Arg34 and Tyr36 of Lq2 and residues Pro55, Tyr78, Gly79, Asp80, and Tyr82 of K(+) channel. Many of them are formed by side chains of residues of Lq2 and backbone atoms of the K(+) channel. Thirteen hydrophobic contacts exist between residues Met29, Asn30, Lys31 and Tyr36 of Lq2 and residues Pro55, Ala58, Gly79, Asp80 and Tyr82 of the K(+) channel. These favorable interactions stabilize the association between the two proteins. These observations are in good agreement with the experimental results and can explain the binding phenomena between scorpion toxins and K(+) channels at the level of molecular structure. The consistency between the BD simulation and the experimental data indicates that our three-dimensional model of Lq2-K(+) channel complex is reasonable and can be used in further biological studies such as rational design of blocking agents of K(+) channels and mutagenesis in both toxins and K(+) channels.  相似文献   

12.
Zarrabi M  Naderi-Manesh H 《Proteins》2008,71(3):1441-1449
Kappa-Hefutoxin1 is a K(+) channel-blocking toxin from the scorpion Heterometrus fluvipes. It is a 22-residue protein that adapts a novel fold of two parallel helices linked by two disulfide bridges without beta-sheets. Recognition of interactions of kappa-Hefutoxin1 with the voltage-gated potassium channels, Kv1.1, Kv1.2, and Kv1.3, was studied by 3D-Dock software package. All structures of kappa-Hefutoxin1 were considered during the simulations, which indicated that even small changes in the structure of kappa-Hefutoxin1 considerably affected both the recognition and the binding between kappa-Hefutoxin1 and the Kv1 channels. kappa-Hefutoxin1 is located around the extracellular part of the Kv1 channels, making contacts with its helices. Lys 19, Tyr 5, Arg 6, Trp 9, or Arg 10 in the toxin and residues Asp 402, His 404, Thr 407,Gly 401, and Asp 386 in each subunit of the Kv potassium channel are the key residues for the toxin-channel recognition. Moreover, the simulation result demonstrates that the hydrophobic interactions are important in interaction of negatively charged toxins with potassium channels. The results of our docking/molecular dynamics simulations indicate that our 3D model structure of the kappa-Hefutoxin1-complex is both reasonable and acceptable and could be helpful for smarter drug design and the blocking agents of Kv1 channels.  相似文献   

13.
With homology modeling techniques, molecular mechanics, and molecular dynamics methods, a 3D structure model of Ndx1 is created and refined. This model is further assessed by Profile-3D and ProStat, which confirm that the refined model is reliable. With this model, a flexible docking study is performed and the result indicates that Glu46, Arg88, and Glu90 are three important determinant residues in binding, as they have strong hydrogen bonding interactions and electrostatic interactions with Ap6A. In addition, we further find that three residues, Ser38, Leu39 and Glu46, coordinate enzyme-bound Mg2+ ions in complex N-A. The Glu46 is consistent with the experimental results by Iwai et al., and the other four residues mentioned above may also play vital roles in catalysis of Ndx1.  相似文献   

14.
15.
L A LeBrun  B V Plapp 《Biochemistry》1999,38(38):12387-12393
The rate of association of NAD(+) with wild-type horse liver alcohol dehydrogenase (ADH) is maximal at pH values between pK values of about 7 and 9, and the rate of NADH association is maximal at a pH below a pK of 9. The catalytic zinc-bound water, His-51 (which interacts with the 2'- and 3'-hydroxyl groups of the nicotinamide ribose of the coenzyme in the proton relay system), and Lys-228 (which interacts with the adenosine 3'-hydroxyl group and the pyrophosphate of the coenzyme) may be responsible for the observed pK values. In this study, the Lys228Arg, His51Gln, and Lys228Arg/His51Gln (to isolate the effect of the catalytic zinc-bound water) mutations were used to test the roles of the residues in coenzyme binding. The steady state kinetic constants at pH 8 for the His51Gln enzyme are similar to those for wild-type ADH. The Lys228Arg and Lys228Arg/His51Gln substitutions decrease the affinity for the coenzymes up to 16-fold, probably due to altered interactions with the arginine at position 228. As determined by transient kinetics, the rate constant for association of NAD(+) with the mutated enzymes no longer decreases at high pH. The pH profile for the Lys228Arg enzyme retains the pK value near 7. The His51Gln and Lys228Arg/His51Gln substitutions significantly decrease the rate constants for NAD(+) association, and the pH dependencies show that these enzymes bind NAD(+) most rapidly at a pH above pK values of 8. 0 and 9.0, respectively. It appears that the pK of 7 in the wild-type enzyme is shifted up by the H51Q substitutions, and the resulting pH dependence is due to the deprotonation of the catalytic zinc-bound water. Kinetic simulations suggest that isomerization of the enzyme-NAD(+) complex is substantially altered by the mutations. In contrast, the pH dependencies for NADH association with His51Gln, Lys228Arg, and Lys228Arg/His51Gln enzymes were the same as for wild-type ADH, suggesting that the binding of NAD(+) and the binding of NADH are controlled differently.  相似文献   

16.
Homology modeling of the catalase, CatC cloned and sequenced from rice (Oryza sativa L., cv Ratna an Indica cultivar) has been performed based on the crystal structure of the catalase CatF (PDB code 1m7s) by using the software MODELLER. With the aid of molecular mechanics and molecular dynamics methods, the final model is obtained and is further assessed by PROCHECK and VERIFY - 3D graph, which show that the final refined model is reliable. With this model, a flexible docking study with the hydrogen peroxide, the substrate for catalase, is performed and the results indicate that Arg310, Asp343 and Arg346 in catalase are three important determinant residues in binding as they have strong hydrogen bonding contacts with the substrate. These hydrogen-bonding interactions play an important role for the stability of the complex. Our results may be helpful for further experimental investigations.  相似文献   

17.
Clostridium botulinum C3 exoenzyme inactivates the small GTP-binding protein family Rho by ADP-ribosylating asparagine 41, which depolymerizes the actin cytoskeleton. C3 thus represents a major family of the bacterial toxins that transfer the ADP-ribose moiety of NAD to specific amino acids in acceptor proteins to modify key biological activities in eukaryotic cells, including protein synthesis, differentiation, transformation, and intracellular signaling. The 1.7 A resolution C3 exoenzyme structure establishes the conserved features of the core NAD-binding beta-sandwich fold with other ADP-ribosylating toxins despite little sequence conservation. Importantly, the central core of the C3 exoenzyme structure is distinguished by the absence of an active site loop observed in many other ADP-ribosylating toxins. Unlike the ADP-ribosylating toxins that possess the active site loop near the central core, the C3 exoenzyme replaces the active site loop with an alpha-helix, alpha3. Moreover, structural and sequence similarities with the catalytic domain of vegetative insecticidal protein 2 (VIP2), an actin ADP-ribosyltransferase, unexpectedly implicates two adjacent, protruding turns, which join beta5 and beta6 of the toxin core fold, as a novel recognition specificity motif for this newly defined toxin family. Turn 1 evidently positions the solvent-exposed, aromatic side-chain of Phe209 to interact with the hydrophobic region of Rho adjacent to its GTP-binding site. Turn 2 evidently both places the Gln212 side-chain for hydrogen bonding to recognize Rho Asn41 for nucleophilic attack on the anomeric carbon of NAD ribose and holds the key Glu214 catalytic side-chain in the adjacent catalytic pocket. This proposed bipartite ADP-ribosylating toxin turn-turn (ARTT) motif places the VIP2 and C3 toxin classes into a single ARTT family characterized by analogous target protein recognition via turn 1 aromatic and turn 2 hydrogen-bonding side-chain moieties. Turn 2 centrally anchors the catalytic Glu214 within the ARTT motif, and furthermore distinguishes the C3 toxin class by a conserved turn 2 Gln and the VIP2 binary toxin class by a conserved turn 2 Glu for appropriate target side-chain hydrogen-bonding recognition. Taken together, these structural results provide a molecular basis for understanding the coupled activity and recognition specificity for C3 and for the newly defined ARTT toxin family, which acts in the depolymerization of the actin cytoskeleton. This beta5 to beta6 region of the toxin fold represents an experimentally testable and potentially general recognition motif region for other ADP-ribosylating toxins that have a similar beta-structure framework.  相似文献   

18.
A site-directed mutagenesis, D244E, of S-adenosylhomocysteine hydrolase (AdoHcyase) changes drastically the nature of the protein, especially the NAD(+) binding affinity. The mutant enzyme contained NADH rather than NAD(+) (Gomi, T., Takata, Y., Date, T., Fujioka, M., Aksamit, R. R., Backlund, P. S., and Cantoni, G. L. (1990) J. Biol. Chem. 265, 16102-16107). In contrast to the site-directed mutagenesis study, the crystal structures of human and rat AdoHcyase recently determined have shown that the carboxyl group of Asp-244 points in a direction opposite to the bound NAD molecule and does not participate in any hydrogen bonds with the NAD molecule. To explain the discrepancy between the mutagenesis study and the x-ray studies, we have determined the crystal structure of the recombinant rat-liver D244E mutant enzyme to 2.8-A resolution. The D244E mutation changes the enzyme structure from the open to the closed conformation by means of a approximately 17 degrees rotation of the individual catalytic domains around the molecular hinge sections. The D244E mutation shifts the catalytic reaction from a reversible to an irreversible fashion. The large affinity difference between NAD(+) and NADH is mainly due to the enzyme conformation, but not to the binding-site geometry; an NAD(+) in the open conformation is readily released from the enzyme, whereas an NADH in the closed conformation is trapped and cannot leave the enzyme. A catalytic mechanism of AdoHcyase has been proposed on the basis of the crystal structures of the wild-type and D244E enzymes.  相似文献   

19.
Long-[Arg(3)]insulin-like growth factor-I (IGF-I) is a potent analog of insulin-like growth factor-I that has been modified by a Glu(3) --> Arg mutation and a 13-amino acid extension appended to the N terminus. We have determined the solution structure of (15)N-labeled Long-[Arg(3)]-IGF-I using high resolution NMR and restrained molecular dynamics techniques to a precision of 0.82 +/- 0.28 A root mean square deviation for the backbone heavy atoms in the three alpha-helices and 3.5 +/- 0.9 A root mean square deviation for all backbone heavy atoms excluding the 8 N-terminal residues and the 8 C-terminal eight residues. Overall, the structure of the IGF-I domain is consistent with earlier studies of IGF-I with some minor changes remote from the N terminus. The major variations in the structure, compared with IGF-I, occur at the N terminus with a substantial reorientation of the N-terminal three residues of the IGF-I domain. These results are interpreted in terms of the lower binding affinity for insulin-like growth factor-binding proteins. The backbone dynamics of Long-[Arg(3)]IGF-I were investigated using (15)N nuclear spin relaxation and the heteronuclear nuclear Overhauser enhancement (NOE). There is a considerable degree of flexibility in Long-[Arg(3)]IGF-I, even in the alpha-helices, as indicated by an average ((1)H)(15)N NOE of 0.55 for the regions. The largest heteronuclear NOEs are observed in the helical regions, lower heteronuclear NOEs are observed in the C-domain loop separating helix 1 from helix 2, and negative heteronuclear NOEs are observed in the N-terminal extension and at the C terminus. Despite these data indicating conformational flexibility for the N-terminal extension, slow amide proton exchange was observed for some residues in this region, suggesting some transitory structure does exist, possibly a molten helix. A certain degree of flexibility may be necessary in all insulin-like growth factors to enable association with various receptors and binding proteins.  相似文献   

20.
DNA methyltransferases (DNMTs) are involved in epigenetic regulation of the genome and are promising targets for therapeutic intervention in cancer and other diseases. Until now, very limited information is available concerning the molecular dynamics of DNMTs. The natural product nanaomycin A is the first selective inhibitor of DNMT3B that induce genomic demethylation. Herein we report long (>100 ns) molecular dynamics simulations for human DNMT3B bound to nanaomycin A with and without the presence of the cofactor S-adenosyl-l-methionine (SAM). We concluded that SAM favors the binding of nanaomycin A to DNMT3B. Key interactions of nanaomycin A with DNMT3B involve long lasting interactions with Arg731, Arg733, Arg832, and the catalytic Cys651. Results further support the previous hypothesis that nanaomycin A has key interactions with amino acid residues involved in the mechanism of methylation. This work represents one of the first molecular dynamics studies of DNMT3B. Results of this work shed light on the structure and binding recognition process of a key epigenetic enzyme with a small molecule inhibitor.  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