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 46 毫秒
1.
The spin transition of high-spin (sT2?)low-spin (1A1) of (dithiocyanato) (N-phenyl-2-pyridinaldimine) iron (II) complexes can be altered by substituents on the phenyl ring. Mössbauer spectra at 78K for the 4-substituted derivatives (with the exception of the 4-OH-substituted derivative) indicate that the fraction of low-spin states increases with decreasing substituent electron-withdrawing ability, as measured by the Hammet σ constant (4-OCH3<4-CH3CONH 4-C6H5<4-CH3<4-H<4-Cl<4-NO2). In addition, the effect of methyl-substitution at the ortho-, meta- or paraposition of the phenyl ring on the spin transition was examined. Mössbauer spectra of these methyl-substituted complexes reveal quite different spin equilibria.  相似文献   

2.
Measurements at half-wave potentials of betainylhydrazones and semicarbazones of acyl compounds of the type CH3-CO-X enabled determinations to be made of values for Taft polar substituent constants σx/*. The acceptor effects of these substituents can be compared by the increasing value of the substituent constant σx/*: COO? (+0.78)<COOH (+1.10)<CO-CH3 (+1.30)<CH=N-NH-CO-(+1.50)<COOR (+1.58)<C≡N and COSR (σ* not measured).  相似文献   

3.
The electrostatic properties of halogen atoms are studied theoretically in relation to their ability of halogen bonding, which is an attractive intermolecular interaction of a covalently bonded halogen atom with a negatively charged atom of a neighboring molecule. The electric quadrupole (of electronic origin) with a positive zz component Θzz of a covalently bonded halogen atom, where the z axis is taken along the covalent bond involving the halogen atom, is mainly responsible for the attractive electrostatic interaction with a negatively charged atom. This positive Θzz is an intrinsic property of halogen atoms with the px2py2pz configuration of the valence electronic shell, as shown by ab initio molecular orbital calculations for isolated halogen atoms with this electronic configuration, and increases in the order of F < Cl < Br < I, in parallel with the known general sequence of the strength of halogen bonding. For halogen‐containing aromatic compounds, the substituent effects on the electrostatic properties are also studied. It is shown that the magnitude of Θzz and the electric field originating from it are rather insensitive to the substituent effect, whereas the electric field originating from atomic partial charges has a large substituent effect. The latter electric field tends to partially cancel the former. The extent of this partial cancellation is reduced in the order of Cl < Br < I and is also reducible by proper substitution on or within the six‐membered ring of halobenzene. Perspectives on the development of potential function parameters applicable to halogen‐bonding systems are also briefly discussed. © 2009 Wiley Periodicals, Inc. J Comput Chem 2010  相似文献   

4.
Summary A study of the isotope exchange of the ethyl esters of -brommercuryarylacetic acids with203Hg Br2 and 70% aqueous dioxane at 60°, made it possible to arrange the substituents in the following series in order of decreasing reaction rate: p-I, Br, Cl > p-F >H > o-CH3 > p-CH3> p-typet. C4H9.  相似文献   

5.
Preparation and N.M.R.-spectroscopie Investigation of a Hexacoordinated p-Tolylotantalum(V) Complex: [(p-CH3C6H4)5Ta(p-CH3C6H4)Li · O(C2H5)2] Reaction of tantalum(V) bromide with an etheral solution of p-tolyllithium in the molar ratio 1:5 leads to precipitation of a yellow complex of the brutto formula LiTaTol6 · Ae[Tol = p-CH3C6H4?; Ae = (C2H5)2O]. The 1H and 13C n.m.r. spectra are discussed.  相似文献   

6.
A series of ruthenium(II) complexes with electron-donor or electron-acceptor groups in intercalative ligands, [Ru(phen)2(o-MOP)]2+ (1), [Ru(phen)2(o-MP)]2+ (2), [Ru(phen)2(o-CP)]2+ (3) and [Ru(phen)2(o-NP)]2+ (4), have been synthesized and characterized by elementary analysis, ES-MS, 1H NMR, electronic absorption and emission spectra. The binding properties of these complexes to CT-DNA have been investigated by spectroscopy and viscosity experiments. The results showed that these complexes bind to DNA in intercalation mode and their intrinsic binding constants (Kb) are 1.1, 0.35, 0.53 and 1.7 × 105 M−1, respectively. The subtle but detectable differences occurred in the DNA-binding properties of these complexes are mainly ascribed to the electron-withdrawing abilities of substituents (–OCH3 < –CH3 < –Cl < –NO2) on the intercalative ligands as well as the intramolecular H-bond (for substituent –OCH3) which increase the planarity area of the intercalative ligand to some extent. The density functional theory (DFT) calculations were also performed and used to further discuss the trend in the DNA-binding affinities of these complexes.  相似文献   

7.
The mass spectrometry of substituted benzenesulfonylhydrazides (X.C6H4.SO2NHNH2) has been studied, with X = p-CH3, H, p-CH3O and p-Br. The intensities of [X? C6H4SO2]+ and [X? C6H4]+ follow the Hammett relationship [In(Z/Z0) =ρδp+] with ρ of 0.375 and 2.37, respectively.  相似文献   

8.
Rate constants for the gas-phase reactions of O3 with a series of monoterpenes and related compounds have been determined at 296 ± 2 K and 740 torr total pressure of air or O2 using a combination of absolute and relative rate techniques. Good agreement between the absolute and relative rate data was observed, and the rate constants obtained (in units of 10?17 cm3 molecule?1 s?1) were: α-pinene, 8.7; β-pinene, 1.5; Δ3-carene, 3.8; 2-carene, 24; sabinene, 8.8; d-limonene, 21; γ-terpinene, 14; terpinolene, 140; α-phellandrene, 190; α-terpinene, 870; myrcene, 49; trans-ocimene, 56; p-cymene, <0.005; and 1,8-cineole, <0.015. While these rate constants for α- and β-pinene and sabinene are in good agreement with recent absolute and relative rate determinations, those for the other monoterpenes are generally lower than the literature data by factors of ca. 2–10. The measured rate constants for the monoterpenes are reasonably consistent with predictions based upon the number and positions of the substituent groups around the 〉C?C〈 bond(s).  相似文献   

9.
Kevin W. Cormier  Michael Lewis   《Polyhedron》2009,28(14):3120-3128
The Li+ and Na+ binding of substituted cyclopentadienyl (Cp) anions were investigated using computational techniques. The ring centroid-metal distances and the binding energies of the Cp-metal complexes correlate very well with the ∑σm of the substituted Cp ring. These properties also correlate well with the Cp Θzz values. The trend in the correlations is the more electron-rich the Cp (negative ∑σm and Θzz values values), the shorter the Cp-metal bond and the stronger the binding energy. The NBO metal charges correlate, though not very well in either case, with the Cp Θzz and ∑σm values. However, there is a substantial increase in correlation when the sum of the absolute value of the Hammett σm (∑|σm|) is employed. The significantly improved correlation when the ∑|σm| values are employed leads us to propose a model for substituted Cp charge transfer upon Li+ or Na+ complexation, and it also informs us that the Hammett substituent constant σm contains information about substituent polarizabilities, at least in the case of Li+- and Na+-substituted Cp anions.  相似文献   

10.
Two types of imidazoliophosphane with additional electron‐withdrawing substituents, such as alkoxy or imidazolio groups, are experimentally described and theoretically studied. Diethyl N,N′‐2,4,6‐methyl(phenyl)imidazoliophosphonite is shown to retain a P‐coordinating ability toward a {RhCl(cod)} (cod=cycloocta‐1,5‐diene) center, thus competing with the cleavage of the labile C? P bond. Derivatives of N,N′‐phenylene‐bridged diimidazolylphenylphosphane were isolated in good yield. Whereas the dicationic phosphane proved to be inert in the presence of [{RhCl(cod)}2], the monocationic counterpart was shown to retain the P‐coordinating ability toward a {RhCl(cod)} center, thus competing with the N‐coordinating ability of the nonmethylated imidazolyl substituent. The ethyl phosphinite version of the dication, thus possessing an extremely electron‐poor PIII center, was also characterized. According to the difference between the calculated homolytic and heterolytic dissociation energies, the N2C???P bond of imidazoliophosphanes with aryl, amino, or alkoxy substituents on the P atom is shown to be of dative nature. The P‐coordinating properties of imidazoliophosphanes with various combinations of phenyl or ethoxy substituents on the P atom and those of six diimidazolophosphane derivatives with zero, one, or two methylium substituents on the N atom, were analyzed by comparison of the corresponding HOMOs and LUMOs and by calculation of the IR C?O stretching frequencies of their [RhCl(CO)2] complexes. Comparison of the νCO values allows the family of the electron‐poor Im+PRR′ (Im=imidazolyl) potential ligands to be ranked in the following order versus (R,R′): P(OEt)3<(Ph,Ph)<(Ph,OEt)<(OEt,OEt)<PF3<(Ph,Im)<(Ph,Im+)<(OEt,Im+). The (Ph,Im) representative is therefore the least electron‐donating phosphane for which coordinating behavior toward a RhI center has been experimentally evidenced to date. Ultimate applications in catalysis could be envisaged.  相似文献   

11.
Visible‐light‐driven H2 evolution based on Dye/TiO2/Pt hybrid photocatalysts was investigated for a series of (E)‐3‐(5′‐{4‐[bis(4‐R1‐phenyl)amino]phenyl}‐4,4′‐(R2)2‐2,2′‐bithiophen‐5‐yl)‐2‐cyanoacrylic acid dyes. Efficiencies of hydrogen evolution from aqueous suspensions in the presence of ethylenediaminetetraacetic acid as electron donor under illumination at λ>420 nm were found to considerably depend on the hydrophilic character of R1, varying in the order MOD (R1=CH3OCH2, R2=H)≈ MO4D (R1=R2=CH3OCH2)> HD (R1=R2=H)> PD (R1=C3H7, R2=H). In the case of MOD /TiO2/Pt, the apparent quantum yield for photocatalyzed H2 generation at 436 nm was 0.27±0.03. Transient absorption measurements for MOD ‐ or PD ‐grafted transparent films of TiO2 nanoparticles dipped into water at pH 3 commonly revealed ultrafast formation (<100 fs) of the dye radical cation (Dye.+) followed by multicomponent decays, which involve minor fast decays (<5 ps) almost independent of R1 and major slower decays with significant differences between the two samples: 1) the early decay of the major components for MOD is about 2.5 times slower than that for PD and 2) a redshift of the spectrum occurred for MOD with a time constant of 17 ps, but not for PD . The substituent effects on H2 generation as well as on transient behavior have been discussed in terms substituent‐dependent charge recombination (CR) of Dye.+ with electrons in bulk, inner‐trap, and/or interstitial‐trap states, arising from different solvent reorganization.  相似文献   

12.
The integrated extinction coefficients (A) of the C≡C stretching modes in the IR spectra of 12 germylacetylenes Me3GeC≡CR are determined by the resonance interactions of substituents with the triple bond. TheA 1/2 values change linearly with change in the difference between the effective π-electron charges on the atoms at the triple bond and σ0 R constants of organic substituents R. The average value of the σ0 R constant of the Me3Ge substituent in the compounds studied is +0.06. The resonance acceptor effect of the Me3Ge substituent toward the triple bond (d,π-conjugation) is stronger than the donor effect (σ,π-conjugation). Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 8, pp. 1569–1574, August, 1998.  相似文献   

13.
The redox reactions of thiosulfate with four iron(III) complexes having phenolate-amide-amine coordination, FeIII(L){L = 1,2-bis(2-hydroxybenzamido)ethane, L1; 1,3-bis(2-hydroxybenzamido)propane, L2; 1,5-bis(2-hydroxybenzamido)3-azapentane, L3; and 1,8-bis(2-hydroxybenzamido)3,6-diazaoctane, L4} have been investigated in 10% v/v MeOH + H2O and I = 0.3 mol dm−3. At constant pH (~ 4.8) and under pseudo-first order conditions of [S2O 3 2− ] the reaction obeyed the rate law : − d[FeIII(L)]/dt = k obs [FeIII(L)] + k obs where k obs denotes the observed rate constant of thiosulfate decomposition; k obs = a[S2O 3 2− ] + b[S2O 3 2− ] T 2 is valid for all the complexes, particularly at pH < 6, while k obs = [H+][S2O 3 2− ] T 2 is consistent with the rate law for thiosulfate decomposition proposed earlier. The rate data (k obs) were analysed on the basis of the reactivities of various species of FeIII(L) generated by the equilibrium protonation of the sec-NH of dien and trien spacer units resulting in the ring opening (for [FeIII(L3/L4)]), and acid base equilibrium of the aqua ligand bound to the iron(III) centre ([FeIII(L)(OH2) n ]). The redox activities, both for second and third order paths, show the ligand dependencies : L4<L3<L1<L2 conforming to the fact that the complexes tend to be less susceptible to electron transfer from S2O 3 2− with (i) the increase of the number of chelate rings, (ii) the decrease of overall charge, and (iii) the decrease of ring size offered by the amine moiety (from six membered to five membered one as for [FeIII(L1/L2)(OH2)2]+. There was no evidence for the formation of inner sphere thiosulfato complex, the possibility of the formation of the outer sphere ion-pairs, [Fe(L/HL)(OH2)n +/2+, S2O 3 2− ] with low equilibrium constant value may not be excluded. In view of this, the outer sphere electron transfer (ET) mechanism is the most likely possibility.  相似文献   

14.
Chemical ionization was used to study gas-phase electrophilic addition reactions of chloromethyl ions ([CHxCl3-x]+, x = 0, 1, 2) with a number of substituted benzenes (C6H5Y, Y = NH2, OH, CHO, CN, NO2). Mass-analyzed ion kinetic energy spectrometry was used to characterize the reaction products with respect to the site of electrophilic addition (ring v. substituent). In some cases, examination of secondary reaction products (ion–molecule adduct which has undergone an elimination reaction in the ion source) aided in establishing the original site of electrophilic addition. Aniline, benzonitrile and nitrobenzene exhibited preferential substituent interaction, while phenol and benzaldehyde gave a mixture of ring and substituent reaction products. These gas-phase results differ considerably from solution-phase Friedel–Crafts alkylation; however, they are consistent with the notion of preferential σ-bond formation at polarizable centers of negative charge.  相似文献   

15.
The metal ligand stability constants of violuric acid [H2VA], N-methyl violuric acid [H2MVA], N-phenyl violuric acid [H2PVA] and N-(o-m-p) tolyl violuric acids [N-H2(o-m-p)TVA] with La(III), Ce(III), Pr(III), Nd(III), Sm(III), Eu(III), Gd(III), Dy(III), and Ho(III) have been determined potentiometrically in 50 Vol% ethanol water media at 25°C and at an ionic strength of 0.1 M NaClO4. The stability of the complexes follow the order of basicities of ligands and also the electron affinities of rare earths as measured by their overall ionisation potential. The order of stabilities of rare earths with violuric acids is, La3+ < Ce3+ < Pr3+ < Nd3+ < Sm3+ < Gd3+ < Eu3+ < Dy3+ < Ho3+.  相似文献   

16.
The gas-phase reaction of ozone with eight alkenes including six 1,1-disubstituted alkenes has been investigated at ambient T (285–298 K) and p = 1 atm. of air. The reaction rate constants are, in units of 10−18 cm3 molecule−1 s−1, 9.50 ± 1.23 for 3-methyl-1-butane, 13.1. ± 1.8 for 2-methyl-1-pentene, 11.3 ± 3.2 for 2-methyl-1,3-butadiene (isoprene), 7.75 ± 1.08 for 2,3,3-trimethyl-1-butene, 3.02 ± 0.52 for 3-methyl-2-isopropyl-1-butene, 3.98 ± 0.43 for 3,4-diethyl-2-hexene, 1.39 ± 17 for 2,4,4-trimethyl-2-pentene, and >370 for (cis + trans)-3,4-dimethyl-3-hexene. For isoprene, results from this study and earlier literature data are consistent with: k (cm3 molecule−1 s−1) = 5.59 (+ 3.51, &minus 2.16) × 10−15 e(−3606±279/RT), n = 28, and R = 0.930. The reactivity of the other alkenes, six of which have not been studied before, is discussed in terms of alkyl substituent inductive and steric effects. For alkenes (except 1,1-disubstituted alkenes) that bear H, CH3, and C2H5 substituents, reactivity towards ozone is related to the alkene ionization potential: In k<(10−18 cm3 molecule−1 s−1) = (32.89 ± 1.84) − (3.09 ± 0.20) IP (eV), n = 12, and R = 0.979. This relationship overpredicts the reactivity of C≥3 1-alkenes, of 1,1-disubstituted alkenes, and of alkenes with bulky substituents, for which reactivity towards ozone is lower due to substituent steric effects. The atmospheric persistence of the alkenes studied is briefly discussed. © 1996 John Wiley & Sons, Inc.  相似文献   

17.
A Rebek imide receptor with an acetylene‐linked phenyl ring complexes 2,6‐di(isobutyramido)pyridine in (CDCl2)2 via triple H‐bonding and π–π‐stacking interactions, and the influence of para‐substituents on both rings was investigated by 1H NMR binding titrations. When the phenyl ring was extended to biphenyl and the C(4)‐pyridine substituent varied, interaction energies increased in the order CH3CH2???phenyl<CH3S???phenyl<phenyl???phenyl?N‐methylcarboxamide???phenyl, highlighting the energetic gain from π stacking on amide fragments. The predicted preference of amide–π stacking for an antiparallel alignment of the local dipoles could not be confirmed with the studied system. Different substituents were introduced in the para position of the phenyl ring and their interaction with bound 2,6‐di(isobutyramido)pyridine was investigated. Theoretical predictions that the mere introduction of a substituent has a stabilizing effect on π–π stacking, regardless of its electronic nature, were experimentally confirmed.  相似文献   

18.
Systematic analysis of the effect of para-substituents (H, Cl, Br and OMe) on the meso-phenyl group in vanadyl meso-tetraphenylporphyrins ([VIVO(TPP)] (R=H, 1 ), [VIVO(TCPP)] (R=Cl, 2 ), [ VIVO(TBPP)] (R=Br, 3 ) and [VIVO(TMPP)] (R=OMe, 4 )) on their properties and catalytic oxygen atom transfer (OAT) for oxidation of benzoin to benzil using DMSO as well as 30 % aqueous H2O2 as the sacrificial oxygen source have been studied. Electrochemical and theoretical (density functional theory) studies are in good agreement with the influence of these substituents on the catalytic property of these complexes. Complex [VIVO(TCPP)] ( 2 ) displayed the best catalytic activity for the conversion (92 %) of benzoin to benzil in 30 h with >99 % product selectivity when DMSO was used as an oxygen source, whereas excellent conversion (~100 %) of benzoin to benzil was noticed in 18 h with 95 % product selectivity when 30 % aqueous H2O2 was used as a source of oxygen. Furthermore, among these complexes, the electron-withdrawing nature of the chloro substituent at the p-position of meso-phenyl group significantly influences the oxygen atom transfer. Experimental and simulated EPR studies confirmed the +4 oxidation of vanadium in these complexes. The structure of 2 , 3 and 4 , confirmed by single crystal X-ray diffraction method, are domed in shape, and the displacement of V(IV) ion from the mean porphyrin plane follows the order: 2 (0.458 Å) < 3 (0.459 Å) < 4 (0.479 Å). We observed that the electron-withdrawing nature of chloro substituent at the p-position of meso-phenyl group influence the oxygen atom transfer from vanadyl porphyrin to dimethyl sulfide much.  相似文献   

19.
The σR0 and σp parameters of Me3SiOCR2 and HOCR2 substituents at the triple bond were determined using the IR spectra of individual acetylene derivatives and their H-complexes. These parameters vary as the effective charge on the atoms of the C≡C fragment of terminal acetylenic alcohols and their trimethylsilyl ethers changes due to intermolecular interaction. The most reliable values of σR0 and σp parameters (−0.02 and −0.03, respectively) for the Me3SiOCH2 substituent were established; they indicate a sharp decrease in σ,π-conjugation of the Me3SiOCH2 substituent with the triple bond as compared to the Me3SiCH2 substituent. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 9, pp. 1759–1762, September, 1998.  相似文献   

20.
Using complete orthonormal sets of ψ (α*) ‐self‐frictional exponential type orbitals (ψ (α*) ‐SFETOs) and Qq‐noninteger auxiliary functions (Qq‐NIAFs) introduced by the author, the combined formulas for the one‐ and two‐center one‐range addition theorems of χ‐noninteger Slater type orbitals (χ‐NISTOs) with arbitrary values of distances between centers Rab (for Rab = 0 and Rab ≠ 0), and of integer (for α* = α, –∞ < α ≤ 2) and noninteger (for α* ≠ α, –∞ < α* < 3) self‐frictional (SF) quantum numbers are suggested. The presented relations for the one‐range addition theorems can be useful tools especially in the electronic structure studies of atoms, molecules and solids when χ‐NISTOs are employed as basis functions.  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