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.
DFT optimization studies of 90 syn α-maltotetraose (DP-4) amylose fragments have been carried out at the B3LYP/6-311++G∗∗ level of theory. The DP-4 fragments studied include V-helix, tightly bent conformations, a boat, and a 1C4 conformer. The standard hydroxymethyl rotamers (gg, gt, tg) were examined at different locations in the residue sequence, and their influence on the bridge conformations ?/ψ values and conformer energy is described. Hydroxyl groups were considered to be homodromic, that is, they are either in the all clockwise, ‘c’, or all counterclockwise, ‘r’. Energy differences between conformations are examined in order to assess the stability of the different conformations and to identify the sources of energy that dictate amylose polymer formation. A small nearly cyclic compact structure is of low energy as one would expect when these flexible molecules are studied in vacuo. Many conformations in which the only differences are a single hydroxymethyl variation in the residue sequence show similar energies and bridge conformations, with trends being a result of the hydroxymethyl as well as hydroxyl orientation. In general the ‘c’ structures are of lower energy than the ‘r’ structures, although this is only true for the in vacuo state. The solvent dependence on conformational preference of several low-energy DP-4 structures was investigated via the continuum solvation method COSMO. These results suggest that the ‘r’ structures may be favored for fully solvated molecules.  相似文献   

2.
The σ-charges on various atoms of methyl glycosides have been computed by using the MO-LCAO method of Del Re. The potential and free energies of methyl aldohexopyranosides and methyl aldopentopyranosides in their C1(d) and 1C(d) conformations have been calculated. Minimization of the energies of these conformations has been studied by suitably tilting the axial C-C and C-O bonds. Considerable release of strain is achieved when tilts of 4.5 and 3° are given to the axial hydroxymethyl and hydroxyl groups, respectively, that are involved in Hassel-Ottar effect. A tilt of 3° is also found necessary for the axial OMe group involved in the Hassel-Ottar effect. The calculated free-energy values are in accord with experimental ones, after adding a value of 0.8 kcal.mole?1 for the anomeric effect of -OMe group. These studies predict that all of the methyl aldohexopyranosides, except methyl α-d- and methyl β-d-idopyranosides, favour the C1 conformation. On the other hand, the energy calculations also predict that, of the eight methyl aldopentopyranosides studied, only methyl α-d- and methyl β-d-xylopyranosides and methyl α-d -ribopyranoside favour the C1(d) conformation; for the other pentopyranosides, considerable amounts of both C1(d) and 1C(d) conformations are present in the equilibrium mixture. The calculated values of the percentage of α-anomer present in the equilibrium mixture agree fairly well with those obtained experimentally.  相似文献   

3.
The net charges on various atoms of aldohexopyranose pentaacetates were computed by using the MO-LCAO method of Del Re for σ-charges and the Hückel MO method for π-charges. The potential and free energies of sixteen aldohexopyranose pentaacetates in the C1(D) and 1C(D) conformations were estimated. Minimization of the energies of these conformations was studied by suitably tilting the axial CC and CO bonds. As with the free sugars, considerable release of strain is achieved when tilts of 4.5 and 2° are given to the axial CH2OAc and the axial OAc groups, respectively, involved in the Hassel—Ottar effect in the 1C(D) conformations. In the case of C1(D) conformations, the ideal models have the minimum energy even when the acetate groups are involved in syn-axial interactions, indicating that strain induced by axial acetate groups is less than that of axial hydroxyl groups. The calculated free-energies agree well with the experimental values after adding a value of 0.9 kcal.mole-1 for the anomeric effect of the acetoxyl group. The free-energy calculations also predict that α-D-idohexopyranose pentaacetate and α-D-altrose pentaacetate favour the C1(D) conformation and β-D-idose pentaacetate a C1?1C equilibrium in solution, in agreement with n.m.r. studies.  相似文献   

4.
Previous DFT in vacuo studies on the conformational preferences for cellobiose showed that upon optimization the φH-anti conformations were of lower energy than the syn forms. Upon optimization using an implicit solvation method, COSMO, the syn or observed form was still not predicted to be of lower energy than the φH-anti form, even though optimization after addition of several explicit water molecules did show a relative energy difference favoring the syn form. In order to examine the predictive ability of COSMO on this carbohydrate, constant energy dynamics, DFTMD, simulations were carried out on low energy syn and φH-anti conformations with and without COSMO included during the dynamics. The resulting analysis confirmed that when COSMO is included in the dynamics, the syn conformations become energetically favored over the φH-anti forms suggesting that both solvent and entropy play roles in dictating the solution conformation of cellobiose. Analysis of the dynamic runs includes distributions of selected dihedral angles versus time, conformational transitions, and populations of some quasi-planar, boat, skew forms during the simulations.  相似文献   

5.
Five water molecules were placed in 37 different configurations around alpha- and beta-D-glucopyranose in the gt, gg, and tg conformational states, and the glucose-water complexes were geometry optimized using density functionals at the B3LYP/6-311++G** level of theory. The five water molecules were organized in space and energy minimized using an empirical potential, AMB02C, and then further geometry optimized using DFT algorithms to minimum energy positions. Electronic energy, zero point vibrational energy, enthalpy, entropy, stress energy on glucose and the water cluster, hydrogen-bond energy, and relative free energy were obtained for each configuration using thermodynamic procedures and an analytical Hessian program. The lowest energy complex was that of a clustering of water molecules around the 1- and 6-hydroxyl positions of the beta-gt anomer. Configurations in which the water molecules created a favorable network completely around and under glucose were found to have low energy for both alpha and beta anomers. Calculation of the alpha/beta anomeric ratio using the zero point corrected energy gave, approximately 32/68%, highly favoring the beta anomer in agreement with the experimental approximately 36/64% value. This ratio is better than the approximately 50/50% ratio found in our previous monohydrate study. An approximate hydroxymethyl population was obtained by noting average relative energies among the three conformational states, gt, gg, and tg. In the beta anomer complexes the gt conformation was favored over the gg state, while in the alpha anomer complexes the gg state was favored over the gt conformation, with the tg conformations all being of higher energy making little or no contribution to the rotamer population. Some geometry variances, found between glucose in vacuo and glucose after interaction with water molecules, are described and account for some observed C-5-C-6 bond length anomalies reported by us previously for the vacuum glucose structures.  相似文献   

6.
Thirty-five conformations of alpha- and beta-d-mannopyranose, the C-2 substituted epimer of glucopyranose, were geometry optimized using the density functional (B3LYP), and basis set (6-311++G**). Full geometry optimization was performed on the hydroxymethyl rotamers (gg/gt/tg) and an analytical hessian program was used to calculate the harmonic vibrational frequencies, zero point energy, enthalpy, and entropy. The lowest energy conformation investigated is the beta-tg in the (4)C(1) chair conformation. The in vacuo calculations showed little energetic preference for either the alpha or beta anomer for mannopyranose in the (4)C(1) chair conformation. Results are compared to similar glucopyranose calculations in vacuo where the alpha anomer is approximately 1kcal/mol lower in electronic energy than the beta anomer. In the case of the generally higher energy (1)C(4) chair conformations, one low-energy, low-entropy beta-gg-(1)C(4) chair conformation was identified that is within approximately 1.4kcal/mol of the lowest energy (4)C(1) conformation of mannopyranose. Other (1)C(4) chair conformations in our investigation are approximately 2.9-7.9kcal/mol higher in overall energy. Many of the (3,O)B, B(3,O), (1,4)B, and B(1,4) boat forms passed through transitions without barriers to (1)S(3), (5)S(1), (1)S(5) skew forms with energies between approximately 3.6 and 8.9kcal/mol higher in energy than the lowest energy conformation of mannopyranose. Boat forms were found that remained stable upon gradient optimization. As with glucopyranose, the orientation and interaction of the hydroxy groups make a significant contribution to the conformation/energy relationship in vacuo.  相似文献   

7.
Forty-one conformations of alpha- and beta-d-galactopyranose were geometry optimized using the B3LYP density functional and 6-311++G** basis set. Full geometry optimization was performed on different ring geometries and different hydroxymethyl rotamers (gg/gt/tg). Analytically derived Hessians were used to calculate zero point energy, enthalpy, and entropy. The lowest energy and free-energy conformation found is the alpha-gg-(4)C(1)-c chair conformation, which is of lower electronic and free energy than the lowest energy alpha-d-glucopyranose conformer because of favorable hydrogen-bonding interactions. The in vacuo calculations showed considerable ( approximately 2.2kcal/mol) energetic preference for the alpha over the beta anomer for galactopyranose in both the (4)C(1) and (1)C(4) chair conformations. Results are compared to glucopyranose and mannopyranose calculations in vacuo. Boat and skew-boat forms were found that remained stable upon gradient optimization, although many starting conformations moved to other boat forms upon optimization. As with glucopyranose and mannopyranose, the orientation and interaction of the hydroxyl groups make the most significant contributions to the conformation-energy relationship in vacuo.  相似文献   

8.
The first single-crystal X-ray diffraction study of tetrafluorinated monosaccharide derivatives is presented. Both α- and β-methyl 2,3-dideoxy-2,2,3,3-tetrafluoro-d-galactopyranoside anomers adopt the 4C1 conformation. The values for the C1–O1 and C1–O5 bond lengths and the O5–C1–O1–CH3 dihedral angles are in line with what can be expected from the anomeric and exo-anomeric effects. The chair conformations are slightly distorted, presumably due to repulsion between 1,3-diaxial C–O and C–F bonds. The asymmetric unit of both compounds contains up to three independent molecules, which differ in the conformation of the hydroxymethyl group (including in one case a ‘forbidden’ gg rotamer). The molecular packing of the β-anomer shows a clear segregation between fluorinated and hydrophilic domains, while for the α-anomer the regions of fluorine segregation are broken by interleafing of OMe groups. There is one close OH?F contact, which is likely to arise from the crystal packing. NMR studies show that the two anomers also adopt a 4C1 conformation in solution (D2O, CDCl3).  相似文献   

9.
Renfrew PD  Butterfoss GL  Kuhlman B 《Proteins》2008,71(4):1637-1646
Amino acid side chains adopt a discrete set of favorable conformations typically referred to as rotamers. The relative energies of rotamers partially determine which side chain conformations are more often observed in protein structures and accurate estimates of these energies are important for predicting protein structure and designing new proteins. Protein modelers typically calculate side chain rotamer energies by using molecular mechanics (MM) potentials or by converting rotamer probabilities from the protein database (PDB) into relative free energies. One limitation of the knowledge‐based energies is that rotamer preferences observed in the PDB can reflect internal side chain energies as well as longer‐range interactions with the rest of the protein. Here, we test an alternative approach for calculating rotamer energies. We use three different quantum mechanics (QM) methods (second order Møller‐Plesset (MP2), density functional theory (DFT) energy calculation using the B3LYP functional, and Hartree‐Fock) to calculate the energy of amino acid rotamers in a dipeptide model system, and then use these pre‐calculated values in side chain placement simulations. Energies were calculated for over 36,000 different conformations of leucine, isoleucine, and valine dipeptides with backbone torsion angles from the helical and strand regions of the Ramachandran plot. In a subset of cases these energies differ significantly from those calculated with standard molecular mechanics potentials or those derived from PDB statistics. We find that in these cases the energies from the QM methods result in more accurate placement of amino acid side chains in structure prediction tests. Proteins 2008. © 2007 Wiley‐Liss, Inc.  相似文献   

10.
CheY is a response regulator protein involved in bacterial chemotaxis. Much is known about its active and inactive conformations, but little is known about the mechanisms underlying long-range interactions or correlated motions. To investigate these events, molecular dynamics simulations were performed on the unphosphorylated, inactive structure from Salmonella typhimurium and the CheY-BeF(3)(-) active mimic structure (with BeF(3)(-) removed) from Escherichia coli. Simulations utilized both sequences in each conformation to discriminate sequence- and structure-specific behavior. The previously identified conformational differences between the inactive and active conformations of the strand-4-helix-4 loop, which are present in these simulations, arise from the structural, and not the sequence, differences. The simulations identify previously unreported structure-specific flexibility features in this loop and sequence-specific flexibility features in other regions of the protein. Both structure- and sequence-specific long-range interactions are observed in the active and inactive ensembles. In the inactive ensemble, two distinct mechanisms based on Thr-87 or Ile-95 rotameric forms, are observed for the previously identified g+ and g- rotamer sampling by Tyr-106. These molecular dynamics simulations have thus identified both sequence- and structure-specific differences in flexibility, long-range interactions, and rotameric form of key residues. Potential biological consequences of differential flexibility and long-range correlated motion are discussed.  相似文献   

11.
The ribose-binding protein (RBP) is a sugar-binding bacterial periplasmic protein whose function is associated with a large allosteric conformational change from an open to a closed conformation upon binding to ribose. The crystal structures of RBP in open and closed conformations have been solved. It has been hypothesized that the open and closed conformations exist in a dynamic equilibrium in solution, and that sugar binding shifts the population from open conformations to closed conformations. Here, we study by computer simulations the thermodynamic changes that accompany this conformational change, and model the structural changes that accompany the allosteric transition, using umbrella sampling molecular dynamics and the weighted histogram analysis method. The open state is comprised of a diverse ensemble of conformations; the open ribose-free X-ray crystal conformations being representative of this ensemble. The unligated open form of RBP is stabilized by conformational entropy. The simulations predict detectable populations of closed ribose-free conformations in solution. Additional interdomain hydrogen bonds stabilize this state. The predicted shift in equilibrium from the open to the closed state on binding to ribose is in agreement with experiments. This is driven by the energetic stabilization of the closed conformation due to ribose-protein interactions. We also observe a significant population of a hitherto unobserved ribose-bound partially open state. We believe that this state is the one that has been suggested to play a role in the transfer of ribose to the membrane-bound permease complex.  相似文献   

12.
Liang Xu  Yonggang Chen  Xiaojuan Wang 《Proteins》2014,82(12):3286-3297
Although the N‐terminal region of Amyloid β (Aβ) peptides plays dual roles as metal‐coordinating sites and conformational modulator, few studies have been performed to explore the effects of mutations at this region on the overall conformational ensemble of Aβ and the binding propensity of metal ions. In this work, we focus on how three familial Alzheimer's disease mutations (D7H, D7N, and H6R) alter the structural characteristics and thermodynamic stabilities of Aβ42 using molecular dynamics simulations. We observe that each mutation displays increased β‐sheet structures in both N and C termini. In particular, both the N terminus and central hydrophobic region of D7H can form stable β‐hairpin structures with its C terminus. The conserved turn structure at Val24–Lys28 in all peptides and Zn2+‐bound Aβ42 is confirmed as the common structural motif to nucleate folding of Aβ. Each mutant can significantly increase the solvation free energy and thus enhance the aggregation of Aβ monomers. The correlation dynamics between Aβ(1–16) and Aβ(17–42) fragments are elucidated by linking the domain motions with the corresponding structured conformations. We characterize the different populations of correlated domain motions for each mutant from a more macroscopic perspective, and unexpectedly find that Zn2+‐bound Aβ42 ensemble shares the same populations as Aβ42, indicating that the binding of Zn2+ to Aβ follows the conformational selection mechanism, and thus is independent of domain motions, even though the structures of Aβ have been modified at a residue level. Proteins 2014; 82:3286–3297. © 2014 Wiley Periodicals, Inc.  相似文献   

13.
Side-chain modeling has a widespread application in many current methods for protein tertiary structure determination, prediction, and design. Of the existing side-chain modeling methods, rotamer-based methods are the fastest and most efficient. Classically, a rotamer is conceived as a single, rigid conformation of an amino acid sidechain. Here, we present a flexible rotamer model in which a rotamer is a continuous ensemble of conformations that cluster around the classic rigid rotamer. We have developed a thermodynamically based method for calculating effective energies for the flexible rotamer. These energies have a one-to-one correspondence with the potential energies of the rigid rotamer. Therefore, the flexible rotamer model is completely general and may be used with any rotamer-based method in substitution of the rigid rotamer model. We have compared the performance of the flexible and rigid rotamer models with one side-chain modeling method in particular (the self-consistent mean field theory method) on a set of 20 high quality crystallographic protein structures. For the flexible rotamer model, we obtained average predictions of 85.8% for chi1, 76.5% for chi1+2 and 1.34 A for root-mean-square deviation (RMSD); the corresponding values for core residues were 93.0%, 87.7% and 0.70 A, respectively. These values represent improvements of 7.3% for chi1, 8.1% for chi1+2 and 0.23 A for RMSD over the predictions obtained with the rigid rotamer model under otherwise identical conditions; the corresponding improvements for core residues were 6.9%, 10.5% and 0.43 A, respectively. We found that the predictions obtained with the flexible rotamer model were also significantly better than those obtained for the same set of proteins with another state-of-the-art side-chain placement method in the literature, especially for core residues. The flexible rotamer model represents a considerable improvement over the classic rigid rotamer model. It can, therefore, be used with considerable advantage in all rotamer-based methods commonly applied to protein tertiary structure determination, prediction, and design and also in predictions of free energies in mutational studies.  相似文献   

14.
Site specific incorporation of molecular probes such as fluorescent- and nitroxide spin-labels into biomolecules, and subsequent analysis by F?rster resonance energy transfer (FRET) and double electron-electron resonance (DEER) can elucidate the distance and distance-changes between the probes. However, the probes have an intrinsic conformational flexibility due to the linker by which they are conjugated to the biomolecule. This property minimizes the influence of the label side chain on the structure of the target molecule, but complicates the direct correlation of the experimental inter-label distances with the macromolecular structure or changes thereof. Simulation methods that account for the conformational flexibility and orientation of the probe(s) can be helpful in overcoming this problem. We performed distance measurements using FRET and DEER and explored different simulation techniques to predict inter-label distances using the Rpo4/7 stalk module of the M. jannaschii RNA polymerase. This is a suitable model system because it is rigid and a high-resolution X-ray structure is available. The conformations of the fluorescent labels and nitroxide spin labels on Rpo4/7 were modeled using in vacuo molecular dynamics simulations (MD) and a stochastic Monte Carlo sampling approach. For the nitroxide probes we also performed MD simulations with explicit water and carried out a rotamer library analysis. Our results show that the Monte Carlo simulations are in better agreement with experiments than the MD simulations and the rotamer library approach results in plausible distance predictions. Because the latter is the least computationally demanding of the methods we have explored, and is readily available to many researchers, it prevails as the method of choice for the interpretation of DEER distance distributions.  相似文献   

15.
A branched α-cyclodextrin is a derivative of an α-cyclodextrin with a branch consisting of an extra glucose unit. Its water solubility is considerably higher than that of the unbranched one. We have studied the high solubility of the molecule in aqueous solution by molecular dynamics simulations. Trajectories of the molecule at 293 K were calculated using GROMOS programs in three different environments, i.e., in vacuo, in the crystalline state, and in aqueous solution. A simulation in vacuo was carried out to explore stable conformations of the molecule in the isolated system. The quality of the simulations were examined by comparing the X-ray and the simulated crystal structures.The results of the simulations show three remarkable structural features of the molecule: self-inclusion with its branched portion, twist-boat conformation of a glucose ring, and wobbling of its macrocycle. Among these, the last feature is closely related to the water solubility of the molecule. The solubility of cyclodextrin appears to be mainly governed by its intramolecular interglucose hydrogen bonds, which inhibit hydration by solvent water molecules. The results of our simulations indicate that the capability to form hydrogen bonds in branched α-cyclodextrin decreased as the macrocycle of the molecule lost its regular circular shape. Such wobbling of the macrocycle was observed on a relatively short time scale (several picoseconds). An extra glucose unit introduced to α-cyclodextrin may cause the improved water solubility of the molecule through the greater wobbling motion of its macrocycle.  相似文献   

16.
The structural and thermodynamics characters of α-syn12 (residues 1-12 of the human α-synuclein protein) peptide in aqueous solution were investigated through temperature replica-exchange molecular dynamics (T-REMD) simulations with the GROMOS 43A1 force field. The two independent T-REMD simulations were completed starting from an initial conformational α-helix and an irregular structure, respectively. Each replica was run for 300 ns. The structural and thermodynamics characters were studied based on parameters such as distributions of backbone dihedral angles, free energy surface, stability of folded β-hairpin structure, and favorite conformations. The results showed that the isolated α-syn12 peptide in water adopted four different conformational states: the first state was a β-hairpin ensemble with Turn(9-6) and four hydrogen bonds, the second state was a β-hairpin ensemble with two turns (Turn(9-6) and Turn(5-2)) and three hydrogen bonds, the third state was a disordered structure with both Turn(8-5) and Turn(5-2), and the last state was a π-helix ensemble. Meanwhile, we studied the free energy change of α-syn12 peptide from the unfolded state to the β-hairpin state, which was in good agreement with the experiments and molecular dynamics simulations for some other peptides. We also analyzed the driving force of the peptide transition. The results indicated that the driving forces were high solvent exposure of hydrophobic Leu8 and hydrophobic residues in secondary structure. To our knowledge, this was the first report to study the isolated α-syn12 peptide in water by T-REMD.  相似文献   

17.
One hundred and two conformations of alpha- and beta-D-allopyranose, the C-3 substituted epimer of glucopyranose, were geometry optimized using the density functional, B3LYP, and the basis set, 6-311++G **. Full geometry optimization was performed on different ring geometries and on the hydroxymethyl rotamers (gg/gt/tg). Analytically derived Hessians were used to calculate zero point energy, enthalpy, and entropy. The lowest energy and free energy conformation found is the alpha-tg(g-)-4C1-c conformation, which is only slightly higher in electronic (approximately 0.2 kcal/mol) and free energy than the lowest energy alpha-D-glucopyranose. The in vacuo calculations showed a small (approximately 0.3 kcal/mol) energetic preference for the alpha- over the beta-anomer for allopyranose in the 4C1 conformation, whereas in the 1C4 conformation a considerable (approximately 1.6 kcal/mol) energetic preference for the beta- over the alpha-anomer for allopyranose was encountered. The results are compared to previous aldohexose calculations in vacuo. Boat and skew forms were found that remained stable upon gradient optimization although many starting boat conformations moved to other skew forms upon optimization. As found for glucose, mannose, and galactose the orientation and interaction of the hydroxyl groups make the most significant contributions to the conformation/energy relationship in vacuo. A comparison of different basis sets and density functionals is made in the Discussion section, confirming the appropriateness of the level of theory used here.  相似文献   

18.
Recognition of Ras by its downstream target Raf is mediated by a Ras-recognition region in the Ras-binding domain (RBD) of Raf. Residues 78–89 in this region occupy two different conformations in the ensemble of NMR solution structures of the RBD: a fully α-helical one, and one where 87–90 form a type IV β-turn. Molecular dynamics simulations of the RBD in solution were performed to explore the stability of these and other possible conformations of both the wild-type RBD and the R89K mutant, which does not bind Ras. The simulations sample a fully helical conformation for residues 78–89 similar to the NMR helical structures, a conformation where 85–89 form a 310-helical turn, and a conformation where 87–90 form a type I |iB-turn, whose free energies are all within 0.3 kcal/mol of each other. NOE patterns and Hα chemical shifts from the simulations are in reasonable agreement with experiment. The NMR turn structure is calculated to be 3 kcal/mol higher than the three above conformations. In a simulation with the same implicit solvent model used in the NMR structure generation, the turn conformation relaxes into the fully helical conformation, illustrating possible structural artifacts introduced by the implicit solvent model. With the Raf R89K mutant, simulations sample a fully helical and a turn conformation, the turn being 0.9 kcal/mol more stable. Thus, the mutation affects the population of RBD conformations, and this is expected to affect Ras binding. For example, if the fully helical conformation of residues 78–89 is required for binding, its free energy increase in R89K will increase the binding free energy by about 0.6 kcal/mol. Proteins 31:186–200, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

19.
Wu Y  Voth GA 《Biophysical journal》2005,89(4):2402-2411
In this study, four possible conformations of the His-37 and Trp-41 residues for the closed state of the influenza M2 ion channel were identified by a conformation scan based on a solid-state NMR restraint. In the four conformations, the His-37 residue can be of either the t-160 or t60 rotamer, whereas Trp-41 can be of either the t-105 or t90 rotamer. These conformations were further analyzed by density functional theory calculations and molecular dynamics simulations, and the data indicate that the His-37 residue most likely adopts the t60 rotamer and should be monoprotonated at the delta-nitrogen site, whereas Trp-41 adopts the t90 rotamer. This result is consistent with published experimental data and points to a simple gating mechanism: in the closed state, the His-37 and Trp-41 residues adopt the (t60, t90) conformation, which nearly occludes the pore, preventing nonproton ions from passing through due to the steric and desolvation effects. Moreover, the His-37 tetrad interrupts the otherwise continuous hydrogen-bonding network of the pore water by forcing the water molecules above and below it to adopt opposite orientations, thus adding to the blockage of proton shuttling. The channel can be easily opened by rotating the His-37 chi2 angle from 60 to 0 degrees . This open structure allows pore water to penetrate the constrictive region and to form a continuous water wire for protons to shuttle through, while being still narrow enough to exclude other ions.  相似文献   

20.
Four 7a-methyl octa(or hexa)hydrocyclopenta[d][1,3]oxazines, five 8a-methyl octa(or hexa)hydro[3,1]benzoxazines, two 6-phenyl hexahydro[3,1]benzoxazinones, and 8a-methyl hexahydro[1,3]benzoxazinone, all cis-fused, were prepared and their stereostructures studied by various one- and two-dimensional (1)H, (13)C, and (15)N NMR spectroscopic methods. In solution, the cyclopentane-fused 2-oxo derivatives and the 1,3-benzoxazinone were found to attain exclusively the N-in/O-in conformation, whereas the 6-phenyl 2-oxo/thioxo derivatives were found to be present predominantly in the N-out conformation. The C-2 unsubstituted and the 2-oxo/thioxo 7a/8a-methyl derivatives were all present in solution as a rapidly interconverting equilibrium of the N-in and N-out conformations. The C-2 methyl derivatives were each found to be interconvertable mixtures of epimers (at C-2) with the N-in conformer predominating for one epimer and the N-out conformer predominating for the other, with both predominating conformers having the C-2 methyl group equatorially orientated. The substituent on the nitrogen (H or Me) was found to be always predominantly equatorial with respect to the heteroring, except for the epimeric 2-methyl derivatives with N-out conformations where steric constraints and the generalized anomeric effect resulted in the axial orientation of the C-2 methyl being favored.  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