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.
A series of optically active P‐chiral oligophosphines (S,R,R,S)‐ 2 , (S,R,S,S,R,S)‐ 3 , (S,R,S,R,R,S,R,S)‐ 4 , and (S,R,S,R,S,R,R,S,R,S,R,S)‐ 5 with four, six, eight, and 12 chiral phosphorus atoms, respectively, were successfully synthesized by a step‐by‐step oxidative‐coupling reaction from (S,S)‐ 1 . The corresponding optically inactive oligophosphines 1′ – 5′ were also prepared. Their properties were characterized by DSC, XRD, and optical‐rotation analyses. While optically active bisphosphine (S,S)‐ 1 and tetraphosphine (S,R,R,S)‐ 2 behaved as small molecules, octaphosphine (S,R,S,R,R,S,R,S)‐ 4 and dodecaphosphine (S,R,S,R,S,R,R,S,R,S,R,S)‐ 5 exhibited the features of a polymer. Furthermore, DSC and XRD analyses showed that hexaphosphine (S,R,S,S,R,S)‐ 3 is an intermediate between a small molecule and a polymer. Comparison of optically active oligophosphines 1 – 5 with the corresponding optically inactive oligophosphines 1′ – 5′ revealed that the optically active phosphines have higher crystallinity than the optically inactive counterparts. It is considered that the properties of oligophosphines depend on the enantiomeric purity as well as the oligomer chain length.  相似文献   

2.
The Pseudomonas species lipase inhibition shows enantioselectivity for R‐enantiomer over S‐enantiomer of exo‐2‐norbornyl‐Nn‐butylcarbamates. R‐, S‐, and racemic‐exo‐2‐norbornyl‐Nn‐butylcarbamates are all characterized as pseudo substrate inhibitors of the enzyme. Thus, the mechanism for Pseudomonas species lipase‐catalyzed hydrolysis of the inhibitor is formation of the first enzyme‐inhibitor Michaelis complex via nucleophilic attack of the active site serine to the inhibitor (Ki step) then formation of the butylcarbamyl enzyme intermediate from this complex (k2 step). Comparison of bimolecular rate constants (ki = k2 / Ki) of the inhibitors indicates that R‐enantiomer is 1.8 times more potent than S‐enantiomer. Thus, Pseudomonas species lipase shows enantioselectivity of 1.8 for Rexo‐2‐norbornyl‐Nn‐butyl‐carbamate over Sexo‐2‐norbornyl‐Nn‐butylcarbamate. Protein‐ligand interaction studies on both enantiomers of exo‐2‐norbornyl‐Nn‐butylcarbamate as inhibitors of Pseudomonas species lipase using AutoDock suggest that R‐enantiomer binds more tightly into the active site of the enzyme than S‐enantiomer. The norbornyl ring of Sexo‐2‐norbornyl‐Nn‐butylcarbamate is repulsive to Ser 82 and His 251 of the catalytic triad as well as to Met 16 of the oxyanion hole. These repulsions may create few unfavorable interactions between Sexo‐2‐norbornyl‐Nn‐butylcarbamate and the enzyme and make this inhibitor a less potent one.  相似文献   

3.
The C3‐symmetric propeller‐chiral compounds (P,P,P)‐ 1 and (M,M,M)‐ 1 with planar π‐cores perpendicular to the C3‐axis were synthesized in optically pure states. (P,P,P)‐ 1 possesses two distinguishable propeller‐chiral π‐faces with rims of different heights named the (P/L)‐face and (P/H)‐face. Each face is configurationally stable because of the rigid structure of the helicenes contained in the π‐core. (P,P,P)‐ 1 formed dimeric aggregates in organic solutions as indicated by the results of 1H NMR, CD, and UV/Vis spectroscopy and vapor pressure osmometry analyses. The (P/L)/(P/L) interactions were observed in the solid state by single‐crystal X‐ray analysis, and they were also predominant over the (P/H)/(P/H) and (P/L)/(P/H) interactions in solution, as indicated by the results of 1H and 2D NMR spectroscopy analyses. The dimerization constant was obtained for a racemic mixture, which showed that the heterochiral (P,P,P)‐ 1 /(M,M,M)‐ 1 interactions were much weaker than the homochiral (P,P,P)‐ 1 /(P,P,P)‐ 1 interactions. The results indicated that the propeller‐chiral (P/L)‐face interacts with the (P/L)‐face more strongly than with the (P/H)‐face, (M/L)‐face, and (M/H)‐face. The study showed the π‐face‐selective aggregation and π‐face chiral recognition of the configurationally stable propeller‐chiral molecules.  相似文献   

4.
The spread s(G) of a graph G is defined as s(G) = max i,j i − λ j |, where the maximum is taken over all pairs of eigenvalues of G. Let U(n,k) denote the set of all unicyclic graphs on n vertices with a maximum matching of cardinality k, and U *(n,k) the set of triangle-free graphs in U(n,k). In this paper, we determine the graphs with the largest and second largest spectral radius in U *(n,k), and the graph with the largest spread in U(n,k).   相似文献   

5.
Suppose G is a chemical graph with vertex set V(G). Define D(G) = {{u, v} ⊆ V (G) | d G (u, v) = 3}, where d G (u, v) denotes the length of the shortest path between u and v. The Wiener polarity index of G, W p (G), is defined as the size of D(G). In this article, an ordering of chemical unicyclic graphs of order n with respect to the Wiener polarity index is given.  相似文献   

6.
Summary In this paper two approximate formulae have been developed for calculation of the integral òT0Tmexp(-E/RT)dT by using integration-by-parts approaches. They are in the following forms: I(m,T) = (RTm+2)/(E+(m+2)RT)exp(-E/RT) I(m,T) = (RTm+2)/(E+(m+2)(0.00099441E+0.93695599RT)exp(-E/RT) The validity of the two formulae has been confirmed and their accuracies have been tested with data from numerical calculating. In contrast to existing other integral methods, both the present approaches are simply used, accurate, and can be used for arbitrary values of m.  相似文献   

7.
We call a subgroup H of a finite group G c-supplemented in G if there exists a subgroup K of G such that G = HK and HK ≤ core(H). In this paper it is proved that a finite group G is p-nilpotent if G is S 4-free and every minimal subgroup of PG N is c-supplemented in N G(P), and when p = 2 P is quaternion-free, where p is the smallest prime number dividing the order of G, P a Sylow p-subgroup of G. As some applications of this result, some known results are generalized.  相似文献   

8.
Observation of Newton black film (NBF) in foam film is possible only with a certain probability W which depends on the concentration C of surfactant in the solution and on the time ta during which adsorption of surfactant at the solution/air interface has taken place. In the paper, the W(C,ta) dependence is derived and used to analyze the effect of ta on the critical surfactant concentration Cc below which NBF in foam film practically cannot be observed. An expression for the Cc(ta) function is obtained which reveals that Cc decreases substantially with increasing ta. This expression is found to describe well experimental Cc(ta) data for foam films obtained from aqueous solution of the therapeutic surfactant INFASURF.  相似文献   

9.
The oscillator strengthsf forE1 transitions along an isoelectronic sequence can be written asf=aK 2+bK+c whereK is a gauge parameter representing the gauge condition of the electromagnetic field. The coefficientsa,b, andc are functions of length (f l) and velocity (f v) values of the oscillator strengths at the Hartree-Fock level. We have shown by making a perturbation expansion of oscillator strengthsf,f l andf v that the gauge parameterK is independent of the nuclear charge. This property has been exploited to extrapolatef values along the isoelectronic sequence of Boron for some representativeE1 transitions within then=2 complex. We obtain good agreement between the extrapolated results with the configuration interaction results.  相似文献   

10.
Two trans stereoisomers of 3‐methylcyclopentadecanol (=muscol), (1R,3R)‐ 2 and (1S,3S)‐ 2 , were efficiently synthesized from (3RS)‐3‐methylcyclopentadecanone (=muscone; (3RS)‐ 1 ) by a highly stereoselective reduction (Scheme). L‐Selectride® (=lithium tri(sec‐butyl)borohydride) was used, followed by the enantiomer resolution by lipase QLG (Alcaligenes sp.). The cis stereoisomers of muscol, (1S,3R)‐ 2 and (1R,3S)‐ 2 , were obtained by the Mitsunobu inversion of (1R,3R)‐ 2 and (1S,3S)‐ 2 , respectively (Scheme). The absolute configuration of (1R,3R)‐ 2 was determined by X‐ray crystal‐structure analysis of its 3‐nitrophthalic acid monoester, 2‐[(1R,3R)‐3‐methylcyclopentadecyl hydrogen benzene‐1,2‐dicarboxylate ((1R,3R)‐ 3b ), and by oxidation of (1R,3R)‐ 2 to (3R)‐muscone.  相似文献   

11.
STUDY ON LIGHTLY SULFONATED SYNDIOTACTIC POLYSTYRENE IONOMERS   总被引:4,自引:0,他引:4  
Sulfonated syndiotactic polystyrene ionomers (SsPS) with 1.8 mol% degree of sulfonation have been studied.WAXD shows that the crystallinity of SsPS ionomers was decreased with increasing diameter size of the counter ions andsPS>SsPS-H>SsPS-K>SsPS-Zn. Moreover, SsPS ionomers only have a α cystal form, while original sPS has two crystalforms: α and β crystal form. TGA shows that the thermal stability of SsPS ionomers is higher than that of the original sPSand SsPS-Zn>SsPS-K>SsPS-H. DSC shows that all the glass transition temperatures (T_g) of SsPS ionomers are higherthan that of the neat sPS and SsPS-Zn>SsPS-Na>SsPS-K>SsPS-H. However, the melting temperature (T_m) andcrystallization peak temperature (T_p) of SsPS ionomers are lower and SsPS-H>SsPS-Zn>SsPS-K>SsPS-Na, while thecrystallinity (X_c) of SsPS-Zn is the lowest. Nonisothermal crystallization kinetics shows that the Avrami index of sPS andSsPS-H are both about 4, suggesting the nucleation growth of SsPS-H with lower degree of sulfonation still keeps its three-dimension form. FTIR spectra of SsPS ionomers show a splitting absorption band for asymmetric stretching vibration ofsulfonation group. The CH in-plane bending vibration of benzene ring shifted to higher wavenumber and the symmetricstretching vibration of sulfonation group changed slightly with different counter ion neutralized SsPS ionomers.  相似文献   

12.
On the bases of the topological structures of the three big classes of icosahedral fullerenes: (1) Cn(Ih, n=60h2; h=1, 2,…), (2) Cn(Ih, n=20h2; h=1, 2,…), and (3) Cn(I, n=20(h2+hk+k2), h>k; h, k=1, 2,…), we derived formulas for the decomposition of their nuclear motions into irreducible representations. Hence, we obtained the infrared and Raman active modes for all of the icosahedral (Ih and I) fullerenes theoretically. © 1998 John Wiley & Sons, Inc. Int J Quant Chem 66 : 113–117, 1998  相似文献   

13.
The spherically averaged electron-pair intracule (relative motion) h(u) and extracule (center-of-mass motion) d(R) densities are a couple of densities which characterize the motion of electron pairs in atomic systems. We study a generalized electron-pair density (q; a, b) that represents the probability density function for the magnitude of two-electron vector a r j +b r k of any pair of electrons j and k to be q, where a and b are nonzero real numbers. In particular, h(u)=g(u;1, −1) and d(R) = . It is shown that the scaling property of the Dirac delta function and the inversion symmetry of orbitals in atoms due to the central force field generate several isomorphic relations in the electron-pair density (q; a, b) with respect to the two parameters a and b. The approximate isomorphism d(R)≅8h(2R) known in the literature between the intracule and extracule densities is a special case of the present results. Received: 24 May 2000 / Accepted: 18 July 2000 / Published online: 27 September 2000  相似文献   

14.
Topological indices are numerical parameters of a molecular graph, which characterize its topology and are usually graph invariant. In quantitative structure–activity relationship/quantitative structure–property relationship study, physico‐chemical properties and topological indices such as Randić, atom–bond connectivity (ABC), and geometric–arithmetic (GA) index are used to predict the bioactivity of chemical compounds. Graph theory has found a considerable use in this area of research. In this paper, we study hex‐derived networks HDN1(n) and HDN2(n), which are generated by hexagonal network of dimension n and derive analytical closed results of general Randić index Rα(G) for different values of α, for these networks of dimension n. We also compute the general first Zagreb, ABC, GA, ABC4, and GA5 indices for these hex‐derived networks for the first time and give closed formulae of these degree‐based indices for hex‐derived networks. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

15.
Electron spectra are generally presented in arbitrary units. The experimental elastic peak intensity Iespec(E) is determined by the elastic backscattering probability Ie(E) of electrons backscattered elastically within the solid angle of the spectrometer. The experimental elastic peak Iespec(E) is converted to Ie(E) backscattering probability using our new procedure based on the Goto ie(E) elastic backscattering current database. The elastic backscattering probability Ic(E) was calculated applying the EPESWIN software of Jablonski. Ie(E) < Ic(E) due to the surface losses of electrons, characterized by the surface excitation parameter Pse (SEP). Pse(E) was determined experimentally using the Goto database and the relationship of Tanuma. Our new procedure is applied to angular‐resolved (AREPES) spectra of Jablonski and Zemek presented in arbitrary units. In their AREPES experiments, the experimental elastic peak intensity Iespec = Ie(E, αd, ΔΩ) was measured at αd angle of detection (35–74°) with a small HSA, with ΔΩ solid angle. The experimental value at 42° $I_{e}(E, {\it{42}}\deg{\hbox{}}, {\Delta}\Omega)$ was converted to probability with the Goto database. It was corrected with a SEP parameter Pse, determined by trial and error method for Si, Ni, Cu and Ag for E = 0.5 and 1 keV primary energies. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

16.
The regio-selective four step synthesis of (1S,2R,3S,4R)-4,7,7-trimethyl-3-(neopentyloxy)bicyclo[2.2.1]heptan-2-ol, as recognized efficient chiral auxiliary, is presented. The strategy based on opening of the key acetal 15 (=(2S,3aR,4S,7R,7aS)-2-tert-butyl-4,8,8-trimethylhexahydro-2H-4,7-methano-1,3-benzodioxole) thus circumvents the poor reactivity of the neopentyl electrophile under alkylation conditions. Following the same strategy, but using the unreported acetal 22 (=(2R,3aS,4S,7R,7aR)-2-tert-butyl-4,8,8-trimethylhexahydro-2H-4,7-methano-1,3-benzodioxole), the corresponding unreported bis-endo alcohol 23 (=(1R,2R,3S,4S)-3-(2,2-dimethylpropoxy)-4,7,7-trimethylbicyclo[2.2.1]heptan-2-ol) could be isolated only in poor yield. An alternative regioselective synthesis, including an ultimate endo-reduction remains to be found. Several erroneous chiroptical properties from the literature are corrected.  相似文献   

17.
Summary: Biodegradation of film specimens from polyhydroxyalkanoates (PHAs) of two types – poly-3-hydroxybutyrate (PHB) and poly-3-hydroxybutyrate-co-3-hydroxyvalerate (PHBV) – was analysed in different environments: tropical sea waters of the South China Sea (Nha Trang, Vietnam) and soils in the environs of Hanoi (Vietnam), Nha Trang (Vietnam) and Krasnoyarsk (Siberia, Russia). In seawater, the mass loss of the specimens of both types was almost equal. However, in tropical soils, PHB degraded quicker than PHBV. In the Siberian soil, the degradation rate of the PHBV was generally higher than that of PHBV. Analysis of molecular mass of PHA specimens showed its decreasing during biodegradation. In the tropical sea conditions, PHA degrading microorganisms were represented by bacteria of Enterobacter, Bacillus and Gracilibacillus genera. Among PHA degrading bacteria, Burkholderia, Alcaligenes, Bacillus, Mycobacterium and Streptomyces genera were identified in Vietnamese soils, and Variovorax, Stenotrophomonas, Acinetobacter, Pseudomonas, Bacillus and Xanthomonas genera in Siberian soils. Micromycetes of Gongronella, Paecilomyces, Penicillium and Trichoderma genera exhibited PHA degrading activity in Vietnamese soils, and Paecilomyces, Penicillium, Acremonium, Verticillium and Zygosporium genera – in Siberian soils.  相似文献   

18.
The electric dipole moment p ( r ) was computed as the integral of the permanent dipole moment of the solvent molecule μ( r ) weighted by the orientational probability distribution Ω( r ; O ) over all orientations, where O is the orientation of the solvent molecule at r . The relationship between Ω( r ; O ) and the potential of the mean torque was derived; p ( r ) is proportional to the electric field E ( r ) under the following assumptions: (1) the van der Waals (vdW) interaction is independent of the orientation of the solvent molecule at r ; (2) the solvent molecule and its electrical effect are modeled as a point dipole moment; (3) the solvent molecule at r is in a region far from the solute; and (4) μE( r ) ? kBT, where kB is Boltzmann's constant and T is absolute temperature. The errors caused by calculating near‐solute Ω( r ) and p ( r ) from E ( r ) are unclear. The results show that Ω( r ) is inconsistent with the value calculated from E ( r ) for water molecules in the first and second shells of solute with charge state Q = ±1 e, and a large variation in solvent molecular polarizability γmol(r), which appeared in the first valley of 4πr2E(r) for |Q| < 1 e. Nonetheless, p (r) is consistent with the values calculated from E (r) for |Q| ≤ 1 e. The implication is that the assumptions for calculating p ( r ) can be ignored in the calculation of the solvation free energy of biomolecules, as they pertain to protein folding and protein–protein/ligand interactions. © 2011 Wiley Periodicals, Inc. J Comput Chem, 2011  相似文献   

19.
Quantum mechanical (QM) calculations were carried out in order to study the host-guest inclusion complexes of procaine hydrochloride (Pro-H) and butacaine hydrochloride (But-H) with α- and β-cyclodextrins (α- and β-CDs) by PM3 and AM1 methods. The systems were studied by a 1:1 (α-CD/Pro-H, α-CD/But-H, β-CD/Pro-H, and β-CD/But-H) stoichiometric ratio. In this work we calculated the energy of complex formation in vacuo, and this investigation was carried out on the basis of the host-guest approach. The stabilization energy results for the 1:1 host-guest inclusion complexes indicate that the β-CD/Pro-H complex is more stable than the α-CD/Pro-H complex. Furthermore, stabilization energy for the 1:1 inclusion complex of α-CD with But-H is lower than that for the 1:1 inclusion complex of β-CD with But-H. The calculation results show that all complexation processes for the four complexes are exothermic. Enthalpy changes for the α-CD/But-H and β-CD/Pro-H host-guest inclusion complexes are more negative than those for the other ones. ΔG o values for both the β-CD/Pro-H and α-CD/But-H complexes are negative. Correspondence: S. M. Hashemianzadeh, Department of Physical Chemistry, College of Chemistry, Iran University of Science and Technology, Tehran, Iran.  相似文献   

20.
For a connected graph G we denote by d(G,k) the number of vertex pairs at distance k. The Hosoya polynomial of G is H(G,x) = ∑k≥0 d(G,k)xk. In this paper, analytical formulae for calculating the polynomials of armchair open‐ended nanotubes are given. Furthermore, the Wiener index, derived from the first derivative of the Hosoya polynomial in x = 1, and the hyper‐Wiener index, from one‐half of the second derivative of the Hosoya polynomial multiplied by x in x = 1, can be calculated. © 2006 Wiley Periodicals, Inc. Int J Quantum Chem, 2007  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