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 562 毫秒
1.
The interactions between oxaloacetic (OAA) and phosphoenolpyruvic carboxykinase (PEPCK) binding pocket in the presence and absence of hydrazine were carried out using quantum chemical calculations, based on the two-layered ONIOM (ONIOM2) approach. The complexes were partially optimized by ONIOM2 (B3LYP/6-31G(d):PM6) method while the interaction energies between OAA and individual residues surrounding the pocket were performed at the MP2/6-31G(d,p) level of theory. The calculated interaction energies (INT) indicated that Arg87, Gly237, Ser286, and Arg405 are key residues for binding to OAA with the INT values of ?1.93, ?2.06, ?2.47, and ?3.16 kcal mol?1, respectively. The interactions are mainly due to the formation of hydrogen bonding interactions with OAA. Moreover, using ONIOM2 (B3LYP/6-31G(d):PM6) applied on the PEPCKHS complex, two proton transfers were observed; first, the proton was transferred from the carboxylic group of OAA to hydrazine while the second one was from Asp311 to Lys244. Such reactions cause the generation of binding strength of OAA to the pocket via electrostatic interaction. The orientations of Lys243, Lys244, His264, Asp311, Phe333, and Arg405 were greatly deviated after hydrazine incorporation. These indicate that hydrazine plays an important role in terms of not only changing the conformation of the binding pocket, but is also tightly bound to OAA resulting in its conformation change in the pocket. The understanding of such interaction can be useful for the design of hydrazine-based inhibitor for antichachexia agents.
Figure
Backbone superimposition of PEPCK-OAA complexes with (color by element) and without (yellow) inclusion of hydrazine demonstrated the change of binding pocket could inhibit the chemical reaction. The key interactions are based on Arg87 and Lys244 through hydrogen bonding formation with OAA  相似文献   

2.
The reaction force and the electronic flux, first proposed by Toro-Labbé et al. (J Phys Chem A 103:4398, 1999) have been expressed by the existing conceptual DFT apparatus. The critical points (extremes) of the chemical potential, global hardness and softness have been identified by means of the existing and computable energy derivatives: the Hellman-Feynman force, nuclear reactivity and nuclear stiffness. Specific role of atoms at the reaction center has been unveiled by indicating an alternative method of calculation of the reaction force and the reaction electronic flux. The electron dipole polarizability on the IRC has been analyzed for the model reaction HF + CO→HCOF. The electron polarizability determined on the IRC α e (ξ) was found to be reasonably parallel to the global softness curve S(ξ). The softest state on the IRC (not TS) coincides with zero electronic flux.
Figure
Variation of the electronic dipole polarizability  相似文献   

3.
A high-throughput screen against Aurora A kinase revealed several promising submicromolar pyrimidine-aniline leads. The bioactive conformation found by docking these leads into the Aurora A ATP-binding site had a semicircular shape. Macrocycle formation was proposed to achieve novelty and selectivity via ring-closing metathesis of a diene precursor. The nature of the optimal linker and its size was directed by docking. In a kinase panel screen, selected macrocycles were active on other kinase targets, mainly FLT3, JAK2, and CDKs. These compounds then became leads in a CDK/FLT3/JAK2 inhibitor project. Macrocycles with a basic nitrogen in the linker form a salt bridge with Asp86 in CDK2 and Asp698 in FLT3. Interaction with this residue explains the observed selectivity. The Asp86 residue is conserved in most CDKs, resulting in potent pan-CDK inhibition by these compounds. Optimized macrocycles generally have good DMPK properties, and are efficacious in mouse models of cancer. Compound 5 (SB1317/TG02), a pan-CDK/FLT3/JAK2 inhibitor, was selected for preclinical development, and is now in phase 1 clinical trials.
Figure
Structure of SB1317 (left). SB1317 docked into CDK2 (right)  相似文献   

4.
A synchronous, concerted chemical process is rigorously divided by the reaction force F(R), the negative gradient of V(R), into “reactant” and “product” regions which are dominated by structural changes and an intervening “transition” region which is electronically intensive. The reaction force constant κ(R), the second derivative of V(R), is negative throughout the transition region, not just at the nominal transition state, at which κ(R) has a minimum. This is consistent with experimental evidence that there is a transition region, not simply a specific point. We show graphically that significant nonsynchronicity in the process is associated with the development of a maximum of κ(R) in the transition region, which increases as the process becomes more nonsynchronous. (We speculate that for a nonconcerted process this maximum is actually positive.) Thus, κ(R) can serve as an indicator of the level of nonsynchronicity.
Figure
Profiles of potential energy V(R), reaction force F(R), and reaction force constant κ(R) along the intrinsic reaction coordinate R for a nonsynchronous concerted chemical reaction.  相似文献   

5.
6.
The dynamics of Na+ transportation in a transmembrane cyclic peptide nanotube of 8?×?(WL)4/POPE has been simulated. The curve of PMF (potential of mean force) for Na+ moving through the tube, based on ABF (adaptive biasing force) method, indicates that Na+ possesses lower free energy in an α-plane region than in a mid-plane one. It was found that Na+ would desorb one or two water molecules in the first solvation shell when entering the tube and later maintain in a solvation state. The average numbers of water molecules around Na+ are 4.50, 4.09 in the first solvation shell, and 3.10, 4.08 in the second one for Na+ locating in an α-plane zone and a mid-plane region, respectively. However, water molecules far away from Na+ location still nearly arrange in a form of 1-2-1-2 file. The dipole orientations of water molecules in the regions of gaps 1 and 7 display “D-defects”, resulted from the simultaneous electrostatic potentials generated by Na+ and the bare carbonyls at the tube mouths. Such “D-defects” accommodate the energetically favorable water orientations thereby.
Figure
The PMF profile of Na+ transportation in an octa-CPNT and D-defects emerging in the tube.  相似文献   

7.
The energetics of the Menshutkin-like reaction between four mesylate derivatives and ammonia have been computed using B3LYP functional with the 6-31+G** basis set. Additionally, MPW1K/6-31+G** level calculations were carried out to estimate activation barrier heights in the gas phase. Solvent effect corrections were computed using PCM/B3LYP/6-31+G** level. The conversion of the reactant complexes into ion pairs is accompanied by a strong energy decrease in the gas phase and in all solvents. The ion pairs are stabilized with two strong hydrogen bonds in the gas phase. The bifurcation at C2 causes a significant activation barrier increase. Also, bifurcation at C5 leads to noticeable barrier height differentiation. Both B3LYP/6-31+G** and MPW1K/6-31+G** activation barriers suggest the reaction 2 (2a?+?NH3) to be the fastest in the gas phase. The reaction 4 is the slowest one in all environments.
Figure
Ammonium salt formation in a Menshutkin-like reaction between ammonia and (S)-1,4-andydro-2,3-dideoxy-5-O-mesylpentitol (2a)  相似文献   

8.
Gibberellin (GA) is an essential plant hormone and plays a significant role during the growth and development of the higher plants. The molecular recognition mode between GA and receptor Arabidopsis thaliana GIBBERELLIN INSENSITIVE DWARF1 A (AtGID1A) was investigated by molecular docking and dynamics simulations to clarify the selective perceived mechanism of different bioactive GA molecules to AtGID1A. The 6-COOH group of GA, especially its β configuration, was found to be an indispensable pharmacophore group for GA recognition and binding to AtGID1A. Not only does a strong salt bridge interaction between the 6β-COOH group of GA and Arg244 of AtGID1A play a very important role in the GA recognition of the receptor, but also an indirect water bridge interaction between the pharmacophore group 6β-COOH of GA and the residue Tyr322 of AtGID1A is essential for the GA binding to the receptor. The site-directed residues mutant modeling study on the receptor-binding pocket confirmed that the mutations of Arg244 and Tyr322 decreased the GA binding activity due to the disappearances of the salt bridge and the hydrogen bond interaction. The 3β-OH group of GA was well known to be necessary for the GA bioactivity due to its forming a unique hydrogen bond with Tyr127 of AtGID1A. In addition, the hydrophobic interaction between GA and AtGID1A was considered a necessary factor to lock the GA active conformation and stabilize the GA-GID1A complex structure. The novel molecular recognition mode will be beneficial in elucidating the GA regulation function on the growth and development of the higher plants.
Figure
A novel molecular recognition mechanism was determined that the signaling molecule GAs interacted with receptor AtGID1A by not only an important salt bridge and multiple hydrogen bond interactions, but also the hydrophobic interaction as a necessary factor for the stability of the complex GAs-GID1A  相似文献   

9.
Dipole moments (μ), charge distributions, and static electronic first-order hyperpolarizabilities (β μ ) of the two lowest-energy keto tautomers of guanine (7H and 9H) were determined in the gas phase using Hartree–Fock, Møller–Plesset perturbation theory (MP2 and MP4), and DFT (PBE1PBE, B97-1, B3LYP, CAM-B3LYP) methods with Dunning’s correlation-consistent aug-cc-pVDZ and d-aug-cc-pVDZ basis sets. The most stable isomer 7H exhibits a μ value smaller than that of the 9H form by a factor of ca. 3.5. The β μ value of the 9H tautomer is strongly dependent on the computational method employed, as it dramatically influences the β μ (9H)/β μ (7H) ratio, which at the highest correlated MP4/aug-cc-pVDZ level is predicted to be ca. 5. The Coulomb-attenuating hybrid exchange-correlation CAM-B3LYP method is superior to the conventional PBE1PBE, B3LYP, and B97-1 functionals in predicting the β μ values. Differences between the largest diagonal hyperpolarizability components were clarified through hyperpolarizability density analyses. Dipole moment and first-order hyperpolarizability are molecular properties that are potentially useful for distinguishing the 7H from the 9H tautomer.
Figure
Hyperpolarizability density analysis of the most stable guanine tautomer  相似文献   

10.
The interaction between oxazepam and C60 fullerene was explored using first-principles vdW-DF calculations. It was found that oxazepam binds weakly to the fullerene cage via its carbonyl group. The binding of oxazepam to C60 is affected drastically by nonlocal dispersion interactions, while vdW forces affect the corresponding geometries only a little. Furthermore, aqueous solution affects the geometries of the oxazepam approaching to fullerene slightly, while oxazepam binds slightly farther away from the nanocage. The results presented provide evidence for the applicability of the vdW-DF method and serve as a practical benchmark for the investigation of host–guest interactions in biological systems.
Figure
ab initio vdW-DF study on the possibility of formation of oxazepam/C60 complex at aqueous solution  相似文献   

11.
The binding geometry of fluorouracil/cucurbit[n]urils (CB[n]s) complexes with n?=?5–8 is investigated using the first-principles van der Waals density functional (vdW-DF) method, involving full geometry optimization. Such host-guest complexes are typically calculated using conventional DFT method, which significantly underestimates non-local dispersion forces (or vdW contributions) and therefore affects interactions between respected entities. We address here the role of vdW forces for the fluorouracil and CB[n]s molecules which can form directional hydrogen bonds with each other. It was found that the inclusion of dispersion interactions significantly affects the host-guest binding properties and the hydrogen bonding between the molecules provides the main binding mechanism, while results in the same geometries for the considered complexes. The 0.84 eV binding energy, for the thermodynamically favorable state, reveals that the interaction of fluorouracil with CB[n]s is an exothermic interaction and typical for strong hydrogen bonding. Furthermore, we have investigated the binding nature of these host-guest systems in aqueous solution with ab initio MD simulations adopting vdW-DF method. These findings afford evidence for the applicability of the vdW-DF approach and provide a realistic benchmark for the investigation of the host-guest complexes.
Figure
The binding geometry of fluorouracil/CB[n]s complexes is investigated using the first-principles vdW-DF method, involving full geometry optimization.  相似文献   

12.
A comprehensive molecular dynamics simulation study of n-alkanes using the optimized potential for liquid simulation with all-atoms (OPLS-AA) force field at ambient condition has been performed. Our results indicate that while simulations with the OPLS-AA force field accurately predict the liquid state mass density for n-alkanes with carbon number equal or less than 10, for n-alkanes with carbon number equal or exceeding 12, the OPLS-AA force field with the standard scaling factor for the 1–4 intramolecular Van der Waals and electrostatic interaction gives rise to a quasi-crystalline structure. We found that accurate predictions of the liquid state properties are obtained by successively reducing the aforementioned scaling factor for each increase of the carbon number beyond n-dodecane. To better understand the effects of reducing the scaling factor, its influence on the torsion potential profile, and the corresponding gauche-trans conformer distribution, heat of vaporization, melting point, and self-diffusion coefficient for n-dodecane were investigated. This relatively simple procedure enables more accurate predictions of the thermo-physical properties of longer n-alkanes.
Figure
  相似文献   

13.
Density functional theory (DFT) calculations at B3LYP/6-31 G (d,p) and B3LYP/6-311?+?G(d,p) levels for the substituted pyridine-catalyzed isomerization of monomethyl maleate revealed that isomerization proceeds via four steps, with the rate-limiting step being proton transfer from the substituted pyridinium ion to the C=C double bond in INT1. In addition, it was found that the isomerization rate (maleate to fumarate) is solvent dependent. Polar solvents, such as water, tend to accelerate the isomerization rate, whereas apolar solvents, such as chloroform, act to slow down the reaction. A linear correlation was obtained between the isomerization activation energy and the dielectric constant of the solvent. Furthermore, linearity was achieved when the activation energy was plotted against the pK a value of the catalyst. Substituted-pyridine derivatives with high pK a values were able to catalyze isomerization more efficiently than those with low pK a values. The calculated relative rates for prodrugs 16 were: 1 (406.7), 2 (7.6?×?106), 3 (1.0), 4 (20.7), 5 (13.5) and 6 (2.2?×?103). This result indicates that isomerizations of prodrugs 1 and 35 are expected to be slow and that of prodrugs 2 and 6 are expected to be relatively fast. Hence, prodrugs 2 and 35 have the potential to be utilized as prodrugs for the slow release of monomethylfumarate in the treatment of psoriasis and multiple sclerosis.
Figure
Substituted pyridine-catalyzed isomerization of monomethylmaleate (prodrug, cis-isomer) to monomethylfumerate (parental drug, trans-isomer)  相似文献   

14.
We have analyzed the variation of the reaction force F(ξ) and the reaction force constant κ(ξ) along the intrinsic reaction coordinates ξ of the water-assisted proton transfer reactions of HX-N = Y (X,Y = O,S). The profile of the force constant of the vibration associated with the reactive mode, k ξ (ξ), was also determined. We compare our results to the corresponding intramolecular proton transfers in the absence of a water molecule. The presence of water promotes the proton transfers, decreasing the energy barriers by about 12 – 15 kcal mol-1. This is due in part to much smaller bond angle changes being needed than when water is absent. The κ(ξ) profiles along the intrinsic reaction coordinates for the water-assisted processes show striking and intriguing differences in the transition regions. For the HS-N = S and HO-N = S systems, two κ(ξ) minima are obtained, whereas for HO-N = O only one minimum is found. The k ξ (ξ) show similar behavior in the transition regions. We propose that this fine structure reflects the degree of synchronicity of the two proton migrations in each case.
Figure
Fine Structure in the Transition Region  相似文献   

15.
The structural and thermodynamic properties of an anthraquinone derivative were studied by means of quantum-chemical calculations. Conformational analysis using ab initio and density functional theory methods revealed 14 low-energy conformers. In order to discuss similarities and differences in entropy of the conformers, the rotational and vibrational contributions to entropy were correlated with changes in conformer structure. The component of the moment of inertia perpendicular to the molecular plane gives significant input to ΔS rot , whereas the largest contributions to the ΔS vib have vibrations associated with the τ S1C20 coordinate.
Figure
Optimized B3LYP/6-311++G(d,p) geometry of 1-[(2-mercaptoethyl)amino]-9,10-anthraquinone (MEAA) and vibrational contributions to entropy (ΔSvib, in J mol?1?K?1) relative to the most stable conformer  相似文献   

16.
The preliminary cytotoxic effect of 4-ethoxycarbonylmethyl-1-(piperidin-4-ylcarbonyl)-thiosemicarbazide hydrochloride (1)—a potent topoisomerase II inhibitor—was measured using a MTT assay. It was found that the compound decreased the number of viable cells in both estrogen receptor-positive MCF-7 and estrogen receptor-negative MDA-MB-231breast cancer cells, with IC50 values of 146?±?2 and 132?±?2 μM, respectively. To clarify the molecular basis of the inhibitory action of 1, molecular docking studies were carried out. The results suggest that 1 targets the ATP binding pocket.
Figure
4-ethoxycarbonylmethyl-1-(piperidin-4-ylcarbonyl)-thiosemicarbazide hydrochloride  相似文献   

17.
There are only 15 diterpenoids known sharing the rare cubitane skeleton (1,3-diisopropyl-6,10-dimethylcyclododecane, Fig. 1) named after the natural product (+)-cubitene ((+)-1) from the termite Cubitermes umbratus. Cubitane-type diterpenoids (“cubitanoids”) have since then been isolated from gorgonian corals and there are reports that cubitene also occurs in plants. The twelve-membered ring of the cubitanoids is an interesting feature for which we have developed a novel synthesis involving bicyclic precursors. In this review, we discuss the state-of-research regarding isolation, structure elucidation, biological activity, biosynthesis, and total synthesis of cubitane-type diterpenoids.
Fig. 1
Cubitane skeleton and cubitanoids from termites  相似文献   

18.
The solvation and the solvatochromic behavior of the 5-(methylthio)-5′-nitro-2,2′-bithiophene 1 in diethyl ether, dichloromethane, acetonitrile, methanol and formamide was theoretically investigated with an iterative molecular and quantum mechanics (QM/MM) approach. Calculated longest-wavelength solvatochromic absorption band of 1, obtained as averages of statistically uncorrelated configurations, including the solute and explicit solvent molecules of the first and second solvation layer, were in excellent agreement with the experimental results.
Figure
Study of the solvation and the solvatochromism of a donor-acceptor bithiophene in a wide range of solvent polarities  相似文献   

19.
In our quest to explore molecules with chemically significant regions where the Fukui function is negative, we explored reactions where the frontier orbital that indicates the sites for electrophilic attack is not the highest occupied molecular orbital. The highest occupied molecular orbital (HOMO) controls the location of the regions where the Fukui function is negative, supporting the postulate that negative values of the Fukui function are associated with orbital relaxation effects and nodal surfaces of the frontier orbitals. Significant negative values for the condensed Fukui function, however, were not observed.
Figure
The ?10?5isosurface of $ {f^{-}}\left( \mathbf{r} \right) $ (opaque silver surface) traces the nodal regions of the HOMO (translucent colored lobes, with different colors for different phases) of the phenoxide anion  相似文献   

20.
Dimerization of 2-naphthalenecarbonitrile (2-NpCN) mediated by cucurbit[8]uril (CB[8]) has been investigated employing the density functional theory. Different structures of 2-NpCN dimers were generated by combining monomers in anti-head-to-head (A), anti-head-to-tail (B) and syn-head-to-tail (C) fashion. All these dimeric structures possess rigid cube-like architecture. On confinement within the CB[8] dimer A turns out to be the lowest energy structure. Calculated 1H NMR spectra revealed that the 2-NpCN dimer exhibits large shielding for aromatic protons consistent with the experiment. The protons attached to cubane moiety on the other hand, led to down-field signals. Dimerization mediated with CB[8] cavitand is further accompanied by the frequency up-shift (blue shift) of methylene stretching vibration in its infrared spectra.
Figure
2-naphthalenecarbonitrile  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