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 62 毫秒
1.
Two molecular orbitals: MO7 (29a) and MO13 (23a) have been identified using dual space analysis (DSA) as the signatures of adenine non-planarity (C1 point group symmetry). The non-planarity of adenine has been demonstrated to be from the attachment of the amino group (NH2) to purine rings as well as the non-rigid deformability of the purine ring of adenine. Orbital 29a (3a″ in the planar case), a π-like orbital, is the direct result of the attachment of the amino group to the purine ring. Orbital 23a (23a′ in the planar case) is the result of the deformability of the purine ring in non-planar adenine (NP) and will be experimentally challenging to resolve.  相似文献   

2.
The conformational characteristics of the peptide sequence X-l-Pro, where X  Gly or l-Ala and the peptide bond joining X and l-Pro is cis, are evaluated. Semi-empirical potential functions are used to estimate the contributions to the conformational energy made by the non-bonded van der Waals' and electrostatic interactions and the intrinsic torsional potentials about the NCa and CaC′ bonds. Rotations φ1 and ψ1 about the NCa and CaC′ bonds in residue X and rotation ψ2 about the CaC′ bond in l-Pro are permitted, while the angle of rotation φ2 about the NCa bond in l-Pro is fixed at 120 ° by the pyrrolidine ring. The presence of the cis peptide bond connecting X and l-Pro renders the backbone rotations φ1, ψ1 in X dependent upon the rotation ψ2 about the CaC′ bond in l-Pro. (Interdependence of rotations in neighboring residues joined by a cis peptide bond was previously observed in l-alanine oligomers.) The number of energetically allowed conformations for the Gly and l-Ala residues preceding a cis peptide bond l-Pro residue are found to be substantially reduced from those permitted when the peptide bond is trans or when l-Pro is replaced by an amino acid residue. On the other hand, ψ2 = 100 to 160 ° (cis′) and 300 to 0 ° (trans′) are found to be the lowest energy conformations of the l-Pro residue irrespective of the cis or trans conformation of the X-l-Pro peptide bond.  相似文献   

3.
A systematic analysis of the conformational space of the basic structure unit of peptoids in comparison to the corresponding peptide unit was performed based on ab initio MO theory and complemented by molecular mechanics (MM) and molecular dynamics (MD) calculations both in the gas phase and in aqueous solution.The calculations show three minimum conformations denoted as C, aD and a that do not correspond to conformers on the gas phase peptide potential energy hypersurface. The influence of aqueous solvation was estimated by means of continuum models. The MD simulations indicate the aD form as the preferred conformation in solution both in cis and trans peptide bond orientations.  相似文献   

4.
5.
The virtual bond scheme set forth in preceding papers for treating the average properties of polyriboadenylic acid (poly rA) is here applied to the calculation of the unperturbed mean-square end-to-end distance of polydeoxyriboadenylic acid (poly dA). The modifications in structure and in charge distribution resulting from the replacement of the hydroxyl group at C2′ in the ribose residue by hydrogen in deoxyribose produce only minor modifications in the conformational energies associated with the poly dA chain as compared to those found for poly rA. The main difference is manifested in the energy associated with rotations about the C3′–O3′ bond of the deoxyribose residue in the C2′-endo conformation; accessible rotations are confined to the range between 0° and 30° relative to the trans conformation, whereas in the ribose unit the accessible regions comprise two ranges centered at approximately 35° and 85°. The characteristic ratio 〈r2〉0/nl2 calculated on the basis of the conformational energy estimates is ≈9 for the poly dA chain with all deoxyribose residues in the C3′-endo conformation and ≈21 with all residues in the C2′-endo form. Satisfactory agreement is achieved between the theoretical values and experimental results on apurinic acid by treating the poly dA chain as a random copolymer of C3′-endo and C2′-endo conformational isomers present in a ratio of ~1 to 9.  相似文献   

6.
In preceding papers the energies associated with the internal rotations in the sugar–phosphate–sugar complex were described with an analytical potential consisting of a Lennard-Jones 6–12 term and an intrinsic torsional term and representing the best fit to a large number of energies computed with a quantum mechanical ab initio technique. The complex considered there (of 37 atoms and with the chemical formula C10H18O8P) is repesentative of deoxyribonucleic acids. In this paper we apply our potential to evaluating the intramolecular energies of the 39-atom complex C10H18O10P, representative of the ribonucleic acids. The potential energies for the internal rotations (considered independent from one another) and the energy maps for rotations about consecutive bonds of the backbone chain are critically compared, both with those obtained for the deoxy system and with those obtained from different theoretical approaches as available from literature. It is shown that, at least for certain combinations of the internal rotation angles, the choice of the starting geometry for the sugarphosphate–sugar molecule (bond lengths and valence angles) strongly affects the value of the computed energy. If a proper geometry is used, very low energies are predicted by our potential in correspondence of the sets of torsional angles found in various RNAs by x-ray crystallography.  相似文献   

7.
Intramolecularly hydrogen bonded conformations of (Aib-Pro)n sequences have been analysed theoretically. Both 4→1 (C10 and 3→1 (C7 hydrogen bonded regular structures are shown to be stereochemically feasible. Conformational energies for the helical structures have been estimated using classical potential energy methods. Both C10 and C7 conformations have very similar energies. Pyrrolidine ring puckering has a pronounced effect on the energies, and only -endo puckered Pro residues can be accommodated. The theoretical calculations using spectroscopic data suggest that the recently proposed novel 310 helical conformation for benzyloxycarbonyl(Aib-Pro)4-methyl ester is in solution, is indeed energetically and stereochemically favourable.  相似文献   

8.
This paper reports the conformation energy (CE) calculations on three forms of prostaglandins (PGs) PGA1, PGB1 and PGE1 on the basis of the empirical potential energy functions, for the simultaneous rotations around C7–C8 (θ), C12–C13 (β) and C14–C15 (β) bonds [Fig. 1(a)]. The isoenergy contours plotted for θβ rotations for the different β values show the existence of two low energy regions for thg equal to about 90° and 240° in all the three cases. The absolute minimum was obtained for thg = 240° and almost coincided with the crystallographic conformation for PGE1 and PGA1. In the case of PGB1 series of low energy conformations were obtained with the thg values equal to about 90° and 270°, but none of them coincided with the observed crystallographic conformation. The paper discusses the comparison of the different low energy conformations in these three molecules, their biological relevance and the cause of disagreement in the case of PGB1 with the crystallographic data.  相似文献   

9.
Abstract

The Hel UV photoelectron spectrum of trimethyl phosphate (TMP) has been measured and interpreted with the aid of SCF molecular orbital calculations carried out with STO-3G, STO-3G* and 4–31G basis functions. The photoelectron spectrum of TMP is more accurately reproduced by results from 4–31G calculations than by results from STO-3G or STO-3G* calculations. However, all three basis sets yield results which predict the same assignment of the photoelectron spectrum. Results at the 4–31G level indicate that whether calculations are based on crystallographic bond angles and bond lengths or on STO-3G optimized geometries has little effect on the energetic ordering of the upper occupied orbitals. The energetic ordering of orbitals is also found to be only weakly dependent upon the torsional angle φ, describing rotation of ester groups about P-O bonds and upon the torsional angle ψ, describing rotation of methyl groups about C-O bonds. For trimethyl phosphate, with C3 symmetry, the vertical ionization potentials of the upper occupied orbitals are 10.81 eV (8e), 11.4 eV (9a), 11.93 eV (7e), 12.6–12.9 eV (8a and 6e), 14.4 eV (7a) and 15.0–16.0 eV(5e and 6a). Calculations at the 4–31G level indicate that many of the highest occupied orbitals in neutral dimethyl phosphate and methyl phosphate have energies and electron distributions similar to orbitals in TMP.

For TMP, a search for optimized values of φ and ψ has been carried out at the STO-3G* level. In agreement with previous NMR studies and with classical potential calculations, the STO- 3G* results indicate that both the gauche φ= 53.1 °) and anticlinal (φ = 141.9°) conformations are thermally accessible. Also in agreement with the classical potential calculations, the STO-3G* results predict that in the all gauche conformation energy is minimized when the methyl groups assume a staggered geometry (ψ= 60° to 80°) and that an energy maximum occurs for an eclipsed geometry (ψ = 0° to 20°). A study of the dependence of optimized values of O-P-O ester bond angles on the torsional angles, φ, was carried out at the STO-3G, STO-3G* and 4–31G levels. The results demonstrate that for C3 symmetry, the coupling of O-P-O angles to φ is influenced by repulsive steric interactions.  相似文献   

10.
Conformational energy calculations were performed on monosaccharide and oligosaccharide inhibitors and substrates of lysozyme to examine the preferred conformations of these molecules. A grid-search method was used to locate all of the low-energy conformational regions for N-acetyl-β-D -glycosamine (NAG), and energy minimization was then carried out in each of these regions. Three stable positions for the N-acetyl group have ben located, in two of which the plane of the amide unit is normal to the mean plane of the pyranosyl ring. Nine local energy minima were located for the —CH2OH group. The positions of the two vicinal cis —OH groups are determined predominantly by interactions with either the —CH2OH or the N-acetyl group. The most stable conformations of β-N-acetylmuramic acid (NAM) were determined from the study of the low-energy conformations of NAG. In the two stable orientations for the D -lactic acid side chain, the O—C—C′ plane (C′ being the carbon atom of the terminal carboxyl group) was found to be normal to the mean plane of the pyranosyl ring. The low-energy positions for the COOH group of NAM are determined mainly by interactions with neighboring groups. The conformational preferences of the α-anomers of NAG and NAM were also explored. The calculated conformation of the N-acetyl group for α-NAG was quite close to that determined by X-ray analysis. Two of the three lowest energy conformations of α-NAM are similar to the corresponding conformations of the β-anomer. A third low-energy structure, which has a hydrogen bond from the NH of the N-acetyl group to the C?O of the lactic acid group, corresponds very closely to the X-ray structure of this molecule. The preferred conformations of the disaccharides NAG–NAG, NAM–NAG and NAG–NAM were also investigated. Two preferred orientations of the reducing pyranosyl ring relative to the nonreducing ring were found for all of these disaccharides, both of which are close to the extended conformation. In one of these conformations, a hydrogen bond can form between the OH group attached to C3 of the reducing sugar and the ring oxygen of the preceding residue. Each conformation can be stabilized further by a hydrogen bond between the CH2OH (donor) of residue i + 1 and the C?O of residue i (acceptor). The interactions that determine conformations for all oligosaccharides containing both NAG and NAM are shown to be exclusively intraresidue and nearest neighbor interactions, so that it is possible to predict all stable conformations of oligosaccharides containing NAG and NAM in any sequence.  相似文献   

11.
Some general rules governing hydrogen bonding at the ring oxygens of furanosides, pyranosides, and bridge oxygens of glycosides have been formulated from existing data on crystal structures of carbohydrates. Ring oxygens of the majority of the glycopyranosides in the hemiacetal or acetal form are involved in hydrogen bonding such that the hydrogen bond direction is usually equatorial to the ring plane and not axial. In contrast, there are no known examples of ring oxygens of glycofuranosides and methyl-glycopyranosides displaying hydrogen bonding in the crystal. Also, the bridge oxygens of glycosides are not involved in hydrogen bonding. The observed shortening in the exocyclic and endocyclic anomeric C(1)? O bonds and the geminal C? O bonds indicate that compounds with two oxygen atoms attached to the same saturated carbon atom may participate in double-bond-no-bond resonance interaction in the same manner as difluoromethane. It is also possible that under these circumstances the carbon atom exhibits greater than tetracovalency. The “anomeric effect” may also be related to (a) the differences in the “double bonding” or bond shortening in the anomeric C? O bonds of the anomeric glycopyranosides, (b) the shorter intramolecular O(1)…?O(5) non-bonded interaction, and (c) the smaller O(1)C(1)O(5) valence angle in the equatorial anomer compared to the axial anomer. An analysis has been made of the energetically preferred conformations about the glycosyl and glycosidic bonds of 1,4- and 1,3-polysuc-charides. In the 1a, 4e-glycopyranosides the projected angle ?1 [O(5)C(1)OR, where R = C or H] is positive, while it is negative in the 1e, 4e-glycopyranosides. Angle ?2 [C(1)OC(4′)C(3′)] is positive in both the 1,4-anomeric polyglycosides. 1e, 4e- and 1a, 4e -polysaccharides are stabilized by intramolecular O(5)…?H? O(3′) and O(2′)…?O(3′) hydrogen bonding, respectively, and generate linear and helical (cyclic) structures, respectively. 1e, 3e- and 1a, 3e-polysaccharides may be stablized by one of two possible intramolecular hydrogen-bonding schemes such that the 1a, 3e -polysaccharides generate helical structures while the 1a, 3e-polysaccharides generate nonhelical structures. The conformation about the C(5)? C(6) bond in the pyranosides falls into two groups where the angle ?00 [O(5)C(5)C(6)O(6)] is either positive, ~+60 ± 30°, or negative, ~–60 ± 30°, the former conformation being found more frequently. In the furanosides the latter conformation is preferred.  相似文献   

12.
N. V. Joshi  V. S. R. Rao 《Biopolymers》1979,18(12):2993-3004
Conformational energies of α- and β-D -glucopyranoses were computed by varying all the ring bond angles and torsional angles using semiempirical potential functions. Solvent accessibility calculations were also performed to obtain a measure of solvent interaction. The results indicate that the 4C1 (D ) chair is the most favored conformation, both by potential energy and solvent accessibility criteria. The 4C1 (D ) chair conformation is also found to be somewhat flexible, being able to accommodate variations up to 10° in the ring torsional angles without appreciable change in energy. Observed solid-state conformations of these sugars and their derivatives lie in the minimum-energy region, suggesting that the substituents and crystal field forces play a minor role in influencing the pyranose ring conformation. Theory also predicts the variations in the ring torsional angles, i.e., CCCC < CCCO < CCOC, in agreement with the experimental results. The boat and twist-boat conformations are found to be at least 5 kcal mol?1 higher in energy compared to the 4C1 (D ) chair, suggesting that these forms are unlikely to be present in a polysaccharide chain. The 1C4 (D ) chair has energy intermediate between that of the 4C1 (D ) chair and that of the twist-boat conformation. The calculated energy barrier between 4C1 (D ) and 1C4 (D ) conformations is high—about 11 kcal mol?1.  相似文献   

13.
The electronic structural impact on intramolecular proton transfer in the cis- and trans-imino N7 and N9 tautomers of adenine (A) has been studied quantum mechanically, using density functional theory (B3LYP/TZVP, SAOP/TZ2P, LB94/TZ2P) and Green function (OVGF/TZVP) models. It is found that proton transfer does not significantly change isotropic properties but has profound impact on electron distributions of the species through anisotropic properties. The relative energies with respect to the canonical A tautomer (amino-9H), ΔE, for imino 7Hcis, imino 7Htrans, imino 9Hcis and imino 9Htrans are calculated as 16.15, 16.43, 18.46 and 13.80 kcal mol? 1 (B3LYP/TZVP model) and some minor changes in perimeters of the purine ring is also observed. The Hirshfeld atomic charges indicate that whether a proton attached to N(7) or N(9) causes a significant local charge redistribution. However, these charges are insensitive to cistrans proton transfer. Condensed Fukui function reveals N(10) and C(8) as the most electrophilic reactive site among N- and C-atom sites, respectively. We also found that proton transfer significantly alters in-plane σ orbitals, rather than out of plane π orbitals including the frontier orbital 6a″. Moreover, orbital based responses to various proton transfers are presented: the orbital 29a′ (HOMO-1) is a signature orbital differentiating all the four tautomers. Orbital 27a′ is a site (N(7) and N(9)) sensitive orbital, whereas orbital 22a′ is only sensitive to proton orientation on the imino group = N–H.  相似文献   

14.
The molecular conformation of the monoclinic crystalline polymorph of prostaglandin A1 has been determined by X-ray diffraction techniques. The space group is P21 with a = 13.637 (2), b = 7.567 (1), I c = 10.576 (2) Å, β = 107.37 (3)°; Dc = 1.073 g·cm−3 for Z = 2. The molecular conformation is characterized by the nearly parallel arrangement of the C1–C7 and C13–C20 side chains, with a general flattening of the overall structure when compared with the orthorhombic polymorph. The cyclopentenone moiety assumes a C8 envelope conformation with C8 and O9 displaced +0.29 Å and −0.18 Å from the C9–C10=C11–C12 plane respectively. Concerted, small variations of the torsion angles, primarily about the C8–C12, C14–C15 and C16–C17 bonds, bring the monoclinic and orthorhombic conformations into coincidence.  相似文献   

15.
The crystal and molecular structure of rotenone, a naturally occurring insecticide with mitochondrial and mitotic spindle inhibitory properties, was determined by direct methods. The crystals were orthorhombic, space group, P2I2I2I with two molecules in the asymmetric unit; a = 8.413 (1) Å, b = 19.840(1), c = 23.581(1), V = 3936 Å3, Z = 8. The structure was refined by least-squares methods to a final R = 0.067. The two molecules in the asymmetric unit have different conformations about the junction between the nonaromatic rings B and C. Ring B is in a sofa conformation in both molecules, with a slight distortion toward a half-chair in I, but with C8 and C8′ on opposite sides of the planar part of the rings. This difference in conformation results in I having an extended (linear) shape while II is V-shaped. The more elongated conformation of the molecule (I) has not been reported in previous studies. Ring C also has opposite conformations in the two molecules. The angle between the planes formed by rings A and D in molecule I is 64.3°, while in molecule II it is 88.3°. Molecular mechanics techniques were used to determine the energy of the two conformations. These calculations, at room temperature, predict molecule II to be the more stable conformer. The highly flexible site in the B/C ring junction is also chemically very reactive. This flexibility and reactivity are further discussed in terms of rotenone's inhibitory activities.  相似文献   

16.
The sugar–phosphate–sugar complex C10H18O8P, a unit of the polynucleotide chains, was analyzed, making use of 100 conformational energies computed in the Hartree-Fock approximation with a small basis set of Gaussian type orbitals. The geometry of the conformations [which corresponds to the C(2′)-endo deoxy system], the basis set, and the computed total energies are reported in this work. In addition, a number of attempts are presented where we searched for a computationally very simple analytical expression apt to fit, with reasonable accuracy, the computed energies. Lennard-Jones type potential seems to offer an appropriate form capable of reproducing the positions of the maxima and the minima resulting from ab initio computations, but neither the 6-12 nor other similar forms seem to be able to correctly reproduce the intensity of the barriers. Form a details analysis of the barriers to rotation about the bonds O(5′)—C(5′) and C(5′)—C(4′) in terms of nonboned interactions, we found that a substantial improvement in the fit of analytical to ab initio energies may be obtained by distinguishing between atoms characterized by the same atomic number but having different chemical characteristics, like the oxygen atoms of the phosphate group, on one hand, and the oxygen atoms of the sugar rings and the hydroxyl groups, on the other.  相似文献   

17.
Abstract

The molecular and crystal structure of 2′(R)-mercapto-2′-deoxyneplanocin A, C11H13N5O2S M.W.=279.32, has been determined by X-ray analysis. The space group is P212121 with a=10.322(1), b=22.870(2), c=5.273(1)Å and z=4. The structure was solved by direct method, and least-squares refinement using 1806 reflections with |Fo| > 30(F) led to the final R value of 0.045. The sugar C(2′) atom is displaced by 0.35Å opposite to the base N(9), i.e., C(2′)-exo conformation and the torsion angle about the N(9)-C(1) bond is 26.3(4)° (anti conformation).  相似文献   

18.
The elucidation of the mutual influence between peptide bond geometry and local conformation has important implications for protein structure refinement, validation, and prediction. To gain insights into the structural determinants and the energetic contributions associated with protein/peptide backbone plasticity, we here report an extensive analysis of the variability of the peptide bond angles by combining statistical analyses of protein structures and quantum mechanics calculations on small model peptide systems. Our analyses demonstrate that all the backbone bond angles strongly depend on the peptide conformation and unveil the existence of regular trends as function of ψ and/or φ. The excellent agreement of the quantum mechanics calculations with the statistical surveys of protein structures validates the computational scheme here employed and demonstrates that the valence geometry of protein/peptide backbone is primarily dictated by local interactions. Notably, for the first time we show that the position of the Hα hydrogen atom, which is an important parameter in NMR structural studies, is also dependent on the local conformation. Most of the trends observed may be satisfactorily explained by invoking steric repulsive interactions; in some specific cases the valence bond variability is also influenced by hydrogen‐bond like interactions. Moreover, we can provide a reliable estimate of the energies involved in the interplay between geometry and conformations. Proteins 2015; 83:1973–1986. © 2015 Wiley Periodicals, Inc.  相似文献   

19.
In the course of investigation of 8-alkylamino substituted adenosines, the title compounds were synthesized as potential partial agonists for adenosine receptors. The structure determination of these compounds was carried out with the X-ray crystallography study. Crystals of 8-(2-hydroxyethylamino)adenosine are monoclinic, space group P 21; a = 7.0422(2), b = 11.2635(3), c = 8.9215(2) Å, β = 92.261(1)°, V = 707.10(3) Å3, Z = 2; R-factor is 0.0339. The nucleoside is characterized by the anti conformation; the ribose ring has the C(2′)-endo conformation and gauchegauche form across C(4′)–C(5′) bond. The molecular structure is stabilized by intramolecular hydrogen bond of N–H·O type. Crystals of 8-(pyrrolidin-1-yl)adenosine are monoclinic, space group C 2; a = 19.271(1), b = 7.3572(4), c = 11.0465(7) Å, β = 103.254(2)°, V = 1524.4(2) Å3, Z = 4; R-factor is 0.0498. In this compound, there is syn conformation of the nucleoside; the ribose has the C(2′)-endo conformation and gauchegauche form across C(4′)–C(5′) bond. The molecular structure is stabilized by intramolecular hydrogen bond of O–H·N type. For both compounds, the branching net of intermolecular hydrogen bonds occur in the crystal structures.  相似文献   

20.
In the present paper we describe the synthesis, purification, single crystal x-ray analysis, and nmr solution characterization, combined with restrained molecular dynamic simulations, of the cyclic hexapeptide cyclo-(L -Pro-L -Phe-β-Ala)2. The peptide was synthesized by classical solution methods and the cyclization of the free hexapeptide was accomplished in good yields in diluted methylene chloride solution using N,N-dicyclohexyl-carbodiimide. The compound crystallizes in the monoclinic space group P21 from methanol-dichloro-methane solution. The two identical halves of the molecule adopt in the solid state two different conformations. One β-Ala-L -Pro peptide bond is trans, while the second is cis. The molecule is present in dimethylsulfoxide d6 solutions as a mixture of conformational families. One of these corresponds to a C2 symmetrical molecule with both β-Ala-Pro cis peptide bonds, while the second major conformation is very similar to that observed in the solid state. All Pro-Phe segments, both in the solid state and the symmetrical and unsym-metrical solution conformations, display ?,ψ angles close to that of position i + 1 and i + 2 of type II β-turns. In addition, the segments preceeded by a trans β-Ala-Pro peptide bond are characterized by a typical ii + 3 hydrogen bond, which is absent in the conformer containing a cis β-Ala-Pro peptide bond. The latter conformation corresponds to a new structural domain we define as the “pseudo type II β-turn.” © 1994 John Wiley & Sons, Inc.  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