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 thermodynamic stability of RNA hairpin loops has been a subject of considerable interest in the recent past (Wimberly et al., 1991). There have been experimental reports indicating that the hairpins with a C(UUCG)G loop sequence are thermodynamically very stable (Wimberly et al., 1991). We used the solution structure of GGAC(UUCG)GUCC (Cheong et al., 1990; Varani et al., 1991) as the starting conformation in our attempt to understand its thermodynamic stability. We carried out molecular dynamics/free energy simulations to understand the basis for the destabilization of the C(UUCG)G loop by mutating cytosine (C7)-->uracil. Because of the limited length of simulation and the presence of kinetic barriers (solvent intervention) to the uracil-->cytosine mutation, all of our computed free energy differences are based on multiple forward simulations. Based on these calculations we find that the cytosine-->uracil mutation in the loop destabilizes it by approximately 1.5kcal/mol relative to that of the reference state, an A-form RNA but with cytosine (C7) looped out. This is the same sign and magnitude as that observed in the thermodynamic studies carried out by Varani et al.(1991). We have carried out free energy component analysis to understand the effect of mutating the cytosine residue to uracil on the thermodynamic stability of the C(UUCG)G hairpin loops. Our calculations show that the most significant contribution to the stability is from the phosphate group linking U5 and U6, which favors the cytosine residue over uracil by about 6.0 kcal/mol. The residues U5, U6, and G8 in the loop region also contribute significantly to the stability. The contributions from the salt and solvent compensate each other, indicating the dynamic nature of interactions of the environment with the nucleic acid system and the coupling between these two components.  相似文献   

2.
Association rate constants for antigen/antibody associations have been computed by Brownian Dynamics simulations of D. L. Ermak and J. A. McCammon, J. Chem. Phys. 69:1352-1360, 1978. The model of monoclonal antibody (mAb) D44.1 is based on crystallographic data (B. C. Braden et al., J. Mol. Biol. 243:767-781, 1994). Electrostatic forces that steer the antigen to the antibody-combining site are computed by solving the linearized Poisson-Boltzmann equation. D44. 1-HEL complex displays very similar association motifs to a related anti-lysozyme antibody, HyHEL-5-HEL system. The computed association rate constants are comparable in the two systems, although the experimental affinity constants differ by three orders of magnitude (D. Tello et al., Biochem. Soc. Trans. 21:943-946, 1993; K. A. Hibbits et al., Biochemistry. 33:3584-3590, 1994). Simulations suggest that the origin of the differences in the affinity come from dissociation rate constants. We have also carried out simulation experiments on a number of mutant antibody fragment-HEL associations to address the role of electrostatics and, to a limited extent, the orientational aspects of association.  相似文献   

3.
We have investigated the effect of switching ribose to deoxyribose at the closing base-pair of an extra-stable RNA hairpin. Specifically, we studied the sequence 5'-GGAC(UUCG)GUCC, a dodecanucleotide that folds into a well-characterized, "extra stable" RNA hairpin structure. Recently, we showed that hairpins containing a 2',5'-linked (UUCG) loop instead of the native 3',5'-linked loop also exhibit extra-stability (Hannoush and Damha, J. Am. Chem. Soc., 2001, 123, 12368-12374). In this article, we show that the ribose units located at the loop-closing positions (i.e., rC4 and rG9) contribute significantly to the stabilization of RNA hairpins, particularly those containing the 3',5'-UUCG loop. Interestingly, the requirement of rC4 and rG9 is more relaxed for DNA hairpins containing the 2',5'-UUCC loop and, in fact, they may be replaced altogether (ribose--> deoxyribose) without affecting stability. The results broaden our understanding of the behavior of highly stable (UUCG) hairpin loops and how they respond to structural perturbation of the loop-closing base pairs.  相似文献   

4.
Experimental and computational studies of the G[UUCG]C RNA tetraloop   总被引:7,自引:0,他引:7  
In prokaryotic ribosomal RNAs, most UUCG tetraloops are closed by a C-G base-pair. However, this preference is greatly reduced in eukaryotic rRNA species where many UUCG tetraloops are closed by G-C base-pairs. Here, biophysical properties of the C[UUCG]G and G[UUCG]C tetraloops are compared, using experimental and computational methods. Thermal denaturation experiments are used to derive thermodynamic parameters for the wild-type G[UUCG]C tetraloop and variants containing single deoxy substitutions in the loop. A comparison with analogous experiments on the C[UUCG]G motif shows that the two RNA species exhibit similar patterns in response to the substitutions, suggesting that their loop structures are similar. This conclusion is supported by NMR data that suggest that the essential UUCG loop structure is maintained in both tetraloops. However, NMR results show that the G[UUCG]C loop structure is disrupted prior to melting of the stem; this behavior is in contrast to the two-state behavior of the C[UUCG]G molecule. Stochastic dynamics simulations using the GB/SA continuum solvation model, run as a function of temperature, show rare conformational transitions in several G[UUCG]C simulations. These results lead to the conclusion that substitution of a G-C for a C-G closing base-pair increases the intrinsic flexibility of the UUCG loop.  相似文献   

5.
Olson MA 《Proteins》2004,57(4):645-650
The treatment of hydration effects in protein dynamics simulations varies in model complexity and spans the range from the computationally intensive microscopic evaluation to simple dielectric screening of charge-charge interactions. This paper compares different solvent models applied to the problem of estimating the free-energy difference between two loop conformations in acetylcholinesterase. Molecular dynamics (MD) simulations were used to sample potential energy surfaces of the two basins with solvent treated by means of explicit and implicit methods. Implicit solvent methods studied include the generalized Born (GB) model, atomic solvation potential (ASP), and the distance-dependent dieletric constant. By using the linear response approximation (LRA), the explicit solvent calculations determined a free-energy difference that is in excellent agreement with the experimental estimate, while rescoring the protein conformations with GB or the Poisson equation showed inconsistent and inferior results. While the approach of rescoring conformations from explicit water simulations with implicit solvent models is popular among many applications, it perturbs the energy landscape by changing the solvent contribution to microstates without conformational relaxation, thus leading to non-optimal solvation free energies. Calculations applying MD with a GB solvent model produced results of comparable accuracy as observed with LRA, yet the electrostatic free-energy terms were significantly different due to optimization on a potential energy surface favored by an implicit solvent reaction field. The simpler methods of ASP and the distance-dependent scaling of the dielectric constant both produced considerable distortions in the protein internal free-energy terms and are consequently unreliable.  相似文献   

6.
Calimet N  Schaefer M  Simonson T 《Proteins》2001,45(2):144-158
Implicit solvent models are increasingly important for the study of proteins in aqueous solution. Here, the generalized Born (GB) solvent polarization model as implemented in the analytical ACE potential [Schaefer and Karplus (1996) J Phys Chem 100:1578] is used to perform molecular dynamics simulations of two small, homologous proteins: the immunoglobulin-binding domain of streptococcal protein G and the Ras binding domain of Raf. Several model parameterizations are compared through more than 60 ns of simulation. Results are compared with two simpler solvent models-an accessible surface area model and a distant-dependent dielectric model, with finite-difference Poisson calculations, with existing explicit solvent simulations, and with experimental data. The simpler models yield stable but distorted structures. The best GB/ACE implementation uses a set of atomic Voronoi volumes reported recently, obtained by averaging over a large database of crystallographic protein structures. A 20% reduction is applied to the volumes, compensating in an average sense for an excessive de-screening of individual charges inherent in the ACE self-energy and for an undersolvation of dipolar groups inherent in the GB screening function. This GB/ACE parameterization yields stable trajectories on the 0.5-1-ns time scale that deviate moderately (approximately 1.5-2.5 A) from the X-ray structure, reproduce approximately the surface distribution of charged, polar, and hydrophobic groups, and reproduce accurately backbone flexibility as measured by amide NMR-order parameters. Over longer time scales (1.5-3 ns), some of the protein G runs escape from the native energy basin and deviate strongly (3 A) from the native structure. The conformations sampled during the transition out of the native energy basin are overstabilized by the GB/ACE solvation model, as compared with a numerical treatment of the full dielectric continuum model.  相似文献   

7.
van Dijk AD  Fushman D  Bonvin AM 《Proteins》2005,60(3):367-381
When classical, Nuclear Overhauser Effect (NOE)-based approaches fail, it is possible, given high-resolution structures of the free molecules, to model the structure of a complex in solution based solely on chemical shift perturbation (CSP) data in combination with orientational restraints from residual dipolar couplings (RDCs) when available. RDCs can be incorporated into the docking following various strategies: as direct restraints and/or as intermolecular intervector projection angle restraints (Meiler et al., J Biomol NMR 2000;16:245-252). The advantage of the latter for docking is that they directly define the relative orientation of the molecules. A combined protocol in which RDCs are first introduced as intervector projection angle restraints and at a later stage as direct restraints is shown here to give the best performance. This approach, implemented in our information-driven docking approach HADDOCK (Dominguez et al., J Am Chem Soc 2003;125:1731-1737), is used to determine the solution structure of the Lys48-linked di-ubiquitin, for which chemical shift mapping, RDCs, and (15)N-relaxation data have been previously obtained (Varadan et al., J Mol Biol 2002;324:637-647). The resulting structures, derived from CSP and RDC data, are cross-validated using (15)N-relaxation data. The solution structure differs from the crystal structure by a 20 degrees rotation of the two ubiquitin units relative to each other.  相似文献   

8.
It has previously been shown that synthetic peptides corresponding to calcium-binding sites III (SCIII) and IV (SCIV) from troponin-C can undergo a calcium-induced dimerization to form the respective homodimers (Shaw GS, Hodges RS, Sykes BD, 1990, Science 249:280-283; Shaw GS et al., 1992a, J Am Chem Soc 114:6258-6259). In addition, an equimolar mixture of SCIII and SCIV has been shown to form preferentially the SCIII/SCIV heterodimer (Shaw GS et al., 1992a, J Am Chem Soc 114:6258-6259). The stabilities of these dimers have been investigated by using 1H-NMR and circular dichroism spectroscopies to follow temperature- and guanidine hydrochloride (GuHCl)-induced denaturations. It has been found that the most stable species, the SCIII/SCIV heterodimer (delta GuH2O = -64.8 kJ/mol), is about 13 kJ/mol more stable than the least stable species, the SCIV homodimer, while the SCIII homodimer is of intermediate stability. This trend of free energies agrees well with the trend of delta G0 values derived from the products of the dissociation constants for calcium binding and peptide association determined from earlier calcium-titration studies. These observations provide evidence that calcium affinity and the association of 2-calcium binding sites are tightly linked. However, it was noted that in all cases delta G0 was considerably more negative than delta GuH2O determined from GuHCl experiments. This difference increased as the stability of the peptide complex increased, providing evidence that linear extrapolation of GuHCl data for very stable proteins may significantly underestimate the value for delta G0.  相似文献   

9.
We have used a combination of simulated annealing (SA), molecular dynamics (MD) and locally enhanced sampling (LES) methods in order to predict the favourable topologies and loop conformations of dimeric DNA quadruplexes with T2 or T3 loops. This follows on from our previous MD simulation studies on the influence of loop lengths on the topology of intramolecular quadruplex structures [P. Hazel et al. (2004) J. Am. Chem. Soc., 126, 16405–16415], which provided results consistent with biophysical data. The recent crystal structures of d(G4T3G4)2 and d(G4BrUT2G4) (P. Hazel et al. (2006) J. Am. Chem. Soc., in press) and the NMR-determined topology of d(TG4T2G4T)2 [A.T. Phan et al. (2004) J. Mol. Biol., 338, 93–102] have been used in the present study for comparison with simulation results. These together with MM-PBSA free-energy calculations indicate that lateral T3 loops are favoured over diagonal loops, in accordance with the experimental structures; however, distinct loop conformations have been predicted to be favoured compared to those found experimentally. Several lateral and diagonal loop conformations have been found to be similar in energy. The simulations suggest an explanation for the distinct patterns of observed dimer topology for sequences with T3 and T2 loops, which depend on the loop lengths, rather than only on G-quartet stability.  相似文献   

10.
We have reported previously that the apparent rate of peroxynitrite (ONOO(-) ) decay, as followed from its absorbance at 302 nm, decreases in the presence of hydrogen peroxide, mannitol and ethanol (Alvarez et al., 1995, Chem. Res. Toxicol. 8:859-864; Alvarez et al., 1998, Free Radic. Biol. Med. 24:1331-1337). Recently, two papers confirmed the observation and proposed that this slowing effect was due to the formation of absorbing peroxynitrate (O(2) NOO(-) ) as intermediate (Goldstein and Czapski, 1998, J. Am. Chem. Soc. 120:3458-3463; Hodges and Ingold, 1999, J. Am. Chem. Soc. 121:10695-10701). Peroxynitrate would be formed from the reaction of peroxynitrite-derived nitrogen dioxide with superoxide. Superoxide, in turn, would arise from the one-electron oxidation of hydrogen peroxide, or from the reaction of reductive radicals derived from mannitol and ethanol with dioxygen. In agreement with this concept, we show herein that under the conditions of our previous work, the slowing effect is prevented by superoxide dismutase and, in the case of mannitol and ethanol, by reducing the dioxygen concentration of the reaction solutions. Thus, superoxide formation is necessary for the decrease in the rate of absorbance decay. In addition, by simulations using known rate constants and absorption coefficients, we show that the slowing effect can be quantitatively accounted for by the formation of peroxynitrate.  相似文献   

11.

We have investigated the effect of switching ribose to deoxyribose at the closing base-pair of an extra-stable RNA hairpin. Specifically, we studied the sequence 5′-GGAC(UUCG)GUCC, a dodecanucleotide that folds into a well-characterized, “extra stable” RNA hairpin structure. Recently, we showed that hairpins containing a 2′,5′-linked (UUCG) loop instead of the native 3′,5′-linked loop also exhibit extra-stability (Hannoush and Damha, J. Am. Chem. Soc., 2001, 123, 12368–12374). In this article, we show that the ribose units located at the loop-closing positions (i.e., rC 4 and rG 9 ) contribute significantly to the stabilization of RNA hairpins, particularly those containing the 3′,5′-UUCG loop. Interestingly, the requirement of rC4 and rG9 is more relaxed for DNA hairpins containing the 2′,5′-UUCG loop and, in fact, they may be replaced altogether (ribose → deoxyribose) without affecting stability. The results broaden our understanding of the behavior of highly stable (UUCG) hairpin loops and how they respond to structural perturbation of the loop-closing base pairs.  相似文献   

12.
We present a general approach to the design, docking, and virtual screening of multiple combinatorial libraries against a family of proteins. The method consists of three main stages: docking the scaffold, selecting the best substituents at each site of diversity, and comparing the resultant molecules within and between the libraries. The core "divide-and-conquer" algorithm for side-chain selection, developed from an earlier version (Sun et al., J Comp Aided Mol Design 1998;12:597-604), provides a way to explore large lists of substituents with linear rather than combinatorial time dependence. We have applied our method to three combinatorial libraries and three serine proteases: trypsin, chymotrypsin, and elastase. We show that the scaffold docking procedure, in conjunction with a novel vector-based orientation filter, reproduces crystallographic binding modes. In addition, the free-energy-based scoring procedure (Zou et al., J Am Chem Soc 1999;121:8033-8043) is able to reproduce experimental binding data for P1 mutants of macromolecular protease inhibitors. Finally, we show that our method discriminates between a peptide library and virtual libraries built on benzodiazepine and tetrahydroisoquinolinone scaffolds. Implications of the docking results for library design are explored.  相似文献   

13.
Jee J  Byeon IJ  Louis JM  Gronenborn AM 《Proteins》2008,71(3):1420-1431
The immunoglobulin-binding domain B1 of streptococcal protein G (GB1), a very stable, small, single-domain protein, is one of the most extensively used models in the area of protein folding and design. Variants derived from a library of randomized hydrophobic core residues previously revealed alternative folds, namely a completely intertwined tetramer (Frank et al., Nat Struct Biol 2002;9:877-885) and a domain-swapped dimer (Byeon et al., J Mol Biol 2003;333:141-152). Here, we report the NMR structure of the single amino acid mutant Ala-34-Phe which exists as side-by-side dimer. The dimer dissociation constant is 27 +/- 4 microM. The dimer interface comprises two structural elements: First, the beta-sheets of the two monomers pair in an antiparallel arrangement, thereby forming an eight-stranded beta-sheet. Second, the alpha-helix is shortened, ending in a loop that engages in intermolecular contacts. The largest difference between the monomer unit in the A34F dimer and the monomeric wild-type GB1 is the dissolution of the C-terminal half of the alpha-helix associated with a pronounced slow conformational motion of the interface loop. This involves a large movement of the Tyr-33 side chain that swings out from the monomer to engage in dimer contacts.  相似文献   

14.
TL Joseph  DP Lane  CS Verma 《PloS one》2012,7(8):e43985
Atomistic simulations of a set of stapled alpha helical peptides derived from the BH3 helix of MCL-1 (Stewart et al. (2010) Nat Chem Biol 6: 595-601) complexed to a fragment (residues 172-320) of MCL-1 revealed that the highest affinity is achieved when the staples engage the surface of MCL-1 as has also been demonstrated for p53-MDM2 (Joseph et al. (2010) Cell Cycle 9: 4560-4568; Baek et al. (2012) J Am Chem Soc 134: 103-106). Affinity is also modulated by the ability of the staples to pre-organize the peptides as helices. Molecular dynamics simulations of these stapled BH3 peptides were carried out followed by determination of the energies of interactions using MM/GBSA methods. These show that the location of the staple is a key determinant of a good binding stapled peptide from a bad binder. The good binder derives binding affinity from interactions between the hydrophobic staple and a hydrophobic patch on MCL-1. The position of the staple was varied, guiding the design of new stapled peptides with higher affinities.  相似文献   

15.
The linear optical spectra (absorbance, linear dichroism, circular dichroism, fluorescence) of the CP43 (PsbC) antenna of the photosystem II core complex (PSIIcc) pertaining to the S(0)?→?S(1) (Q(Y)) transitions of the chlorophyll (Chl) a pigments are simulated by applying a combined quantum chemical/electrostatic method to obtain excitonic couplings and local transition energies (site energies) on the basis of the 2.9?? resolution crystal structure (Guskov et al., Nat Struct Mol Biol 16:334-342, 2009). The electrostatic calculations identify three Chls with low site energies (Chls 35, 37, and 45 in the nomenclature of Loll et al. (Nature 438:1040-1044, 2005). A refined simulation of experimental spectra of isolated CP43 suggests a modified set of site energies within 143?cm(-1) of the directly calculated values (root mean square deviation: 80?cm(-1)). In the refined set, energy sinks are at Chls 37, 43, and 45 in agreement with earlier fitting results (Raszewski and Renger, J Am Chem Soc 130:4431-4446, 2008). The present structure-based simulations reveal that a large part of the redshift of Chl 37 is due to a digalactosyldiacylglycerol lipid. This finding suggests a new role for lipids in PSIIcc, namely the tuning of optical spectra and the creation of an excitation energy funnel towards the reaction center. The analysis of electrostatic pigment-protein interactions is used to identify amino acid residues that are of potential interest for an experimental approach to an assignment of site energies and energy sinks by site-directed mutagenesis.  相似文献   

16.
The 3-dimensional structures of mirror-image forms of a Leu-5 variant of the trypsin inhibitor Ecballium elaterium (EETI-II) have been determined by 1H NMR spectroscopy and simulated annealing calculations incorporating NOE-derived distance constraints. Spectra were assigned using 2-dimensional NMR methods at 400 MHz, and internuclear distances were determined from NOESY experiments. Three-bond spin-spin couplings between C alpha H and amide protons, amide exchange rates, and the temperature dependence of amide chemical shifts were also measured. The structure consists largely of loops and turns, with a short region of beta-sheet. The Leu-5 substitution produces a substantial reduction in affinity for trypsin relative to native EETI-II, which contains an Ile at this position. The global structure of the Leu-5 analogue studied here is similar to that reported for native EETI-II (Heitz A, Chiche L, Le-Nguyen D, Castro B, 1989, Biochemistry 28:2392-2398) and to X-ray and NMR structures of the related proteinase inhibitor CMTI-I (Bode W et al., 1989, FEBS Lett 242:285-292; Holak TA et al., 1989a, J Mol Biol 210:649-654; Holak TA, Gondol D, Otlewski J, Wilusz T, 1989b, J Mol Biol 210:635-648; Holak TA, Habazettl J, Oschkinat H, Otlewski J, 1991, J Am Chem Soc 113:3196-3198). The region near the scissile bond is the most disordered part of the structure, based on geometric superimposition of 40 calculated structures. This disorder most likely reflects additional motion being present in this region relative to the rest of the protein. This motional disorder is increased in the Leu-5 analogue relative to the native form and may be responsible for its reduced trypsin binding. A second form of the protein synthesized with all (D) amino acids was also studied by NMR and found to have a spectrum identical with that of the (L) form. This is consistent with the (D) form being a mirror image of the (L) form and not distinguishable by NMR in an achiral solvent (i.e., H2O). The (D) form has no activity against trypsin, as would be expected for a mirror-image form.  相似文献   

17.
The extent of insertion of beta-strand s4A into sheet A in intact serpin alpha 1-proteinase inhibitor (alpha 1PI has been probed by peptide annealing experiments [Schulze et al. (1990) Eur. J. Biochem. 194, 51-56]. Twelve synthetic peptides of systematically varied length corresponding in sequence to the unprimed (N-terminal) side of the active site loop were complexed with alpha 1PI. The complexes were then characterized by circular dichroism spectroscopy and tested for inhibitory activity. Four peptides formed complexes which retained inhibitory activity, one of which was nearly as effective as the native protein. Comparison with the three dimensional structures of cleaved alpha 1PI [L?bermann et al. (1984) J. Mol. Biol. 177, 531-556] and plakalbumin [Wright et al. (1990) J. Mol. Biol. 213, 513-528] supports a model in which alpha 1PI requires the insertion of a single residue, Thr345, into sheet A for activity.  相似文献   

18.
Pitera JW  Kollman PA 《Proteins》2000,41(3):385-397
We have extended and applied a multicoordinate free energy method, chemical Monte Carlo/Molecular Dynamics (CMC/MD), to calculate the relative free energies of different amino acid side-chains. CMC/MD allows the calculation of the relative free energies for many chemical species from a single free energy calculation. We have previously shown its utility in host:guest chemistry (Pitera and Kollman, J Am Chem Soc 1998;120:7557-7567)1 and ligand design (Eriksson et al., J Med Chem 1999;42:868-881)2, and here demonstrate its utility in calculations of amino acid properties and protein stability. We first study the relative solvation free energies of N-methylated and acetylated alanine, valine, and serine amino acids. With careful inclusion of rotameric states, internal energies, and both the solution and vacuum states of the calculation, we calculate relative solvation free energies in good agreement with thermodynamic integration (TI) calculations. Interestingly, we find that a significant amount of the unfavorable solvation of valine seen in prior work (Sun et al., J Am Chem Soc 1992;114:6798-6801)3 is caused by restraining the backbone in an extended conformation. In contrast, the solvation free energy of serine is calculated to be less favorable than expected from experiment, due to the formation of a favorable intramolecular hydrogen bond in the vacuum state. These monomer calculations emphasize the need to accurately consider all significant conformations of flexible molecules in free energy calculations. This development of the CMC/MD method paves the way for computations of protein stability analogous to the biochemical technique of "exhaustive mutagenesis." We have carried out just such a calculation at position 133 of T4 lysozyme, where we use CMC/MD to calculate the relative stability of eight different side-chain mutants in a single free energy calculation. Our T4 calculations show good agreement with the prior free energy calculations of Veenstra et al. (Prot Eng 1997;10:789-807)4 and excellent agreement with the experiments of Mendel et al. (Science 1992;256:1798-1802).  相似文献   

19.
M D Wirt  I Sagi    M R Chance 《Biophysical journal》1992,63(2):412-417
X-ray edge and extended x-ray absorption fine structure (EXAFS) techniques provide powerful tools for analysis of local molecular structure of complexes in solution. We present EXAFS results for Co(I) B12 that demonstrate a four-coordinate (distorted) square-planar configuration. Comparison of EXAFS solutions for Co(I) and Co(II) B12 (collected previously; Sagi et al. 1990. J. Am. Chem. Soc. 112:8639-8644) suggest that modulation of the Co-N bond to the axial 5,6-dimethylbenzimidazole (DMB), in the absence of changes in Co-N (equatorial) bond distances, may be a key mechanism in promoting homolytic versus heterolytic cleavage. As Co-C bond homolysis occurs, the Co-N (DMB) bond becomes stronger. However, for heterolytic cleavage to occur, earlier electrochemical studies (D. Lexa and J. M. Saveant. 1976. J. Am. Chem. Soc. 98:2652-2658) and recent studies of methylcobalamin-dependent Clostridium thermoaceticum (Ragsdale et al. 1987. J. Biol. Chem. 262:14289-14297) suggest that removal of the DMB ligand (before Co-C bond cleavage) favors formation of the four-coordinate square-planar Co(I) species while inhibiting formation of the five-coordinate Co(II) B12 complex. This paper presents the first direct evidence that formation of the Co(I) B12 intermediate must involve breaking of the Co-N (DMB) bond.  相似文献   

20.
Computer simulations of phospholipid membranes have been carried out by using a combined approach of molecular and stochastic dynamics and a mean field based on the Marcelja model. First, the single-chain mean field simulations of Pastor et al. [(1988) J. Chem. Phys. 89, 1112-1127] were extended to a complete dipalmitoylphosphatidylcholine molecule; a 102-ns Langevin dynamics simulation is presented and compared with experiment. Subsequently, a hexagonally packed seven-lipid array was simulated with Langevin dynamics and a mean field at the boundary and with molecular dynamics (and no mean field) in the center. This hybrid method, mean field stochastic boundary molecular dynamics, reduces bias introduced by the mean field and eliminates the need for periodic boundary conditions. As a result, simulations extending to tens of nanoseconds may be carried out by using a relatively small number of molecules to model the membrane environment. Preliminary results of a 20-ns simulation are reported here. A wide range of motions, including overall reorientation with a nanosecond decay time, is observed in both simulations, and good agreement with NMR, IR, and neutron diffraction data is found.  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