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 31 毫秒
1.
Cyclopropyllithium, C3H5Li ( 1 ), was studied in the presence of one equivalent lithium bromide (LiBr) in diethylether (DEE)/tetrahydrofuran (THF) mixtures and in THF as solvents. Increasing the THF concentration in DEE/THF leads in the 6Li NMR spectrum to a main signal (S1) at δ0.85 (rel. to ext. LiBr/THF) and a second resonance (S2) at δ0.26 aside from a minor component at δ0.07. In pure THF, the ratio of these signals was 66: 28:6. 6Li and 13C NMR allowed to identify the main signal as belonging to a mixed dimer, 1 •LiBr, and the signal at 0.26 ppm to a fluxional mixed tetramer, 1 2•(LiBr)2. 1J(13C,6Li) coupling constants of 11.0 and 9.8 Hz were measured at 168 K for S1 and S2, respectively, and chemical exchange between both signals was detected by 2D 6Li,6Li exchange spectroscopy and analyzed by temperature-dependent 1D 6Li line-shape calculations. These yielded the equilibrium constants Keq for the chemical exchange Li4(C3H5)2Br2 ⇌ 2 Li2C3H5Br. Their temperature dependence leads to van't Hoff parameters of ΔH° = 4.6 kJ/mol, ΔS° = 41.4 J/mol K, and ΔG°298 = −7.8 kJ/mol. From the rate constants k, Eyring parameters of ΔH* = 42.0 kJ/mol, ΔS* = 33.0 J/mol K, and ΔG*298 = 32.2 kJ/mol were calculated for the forward reaction Li4(C3H5)2Br2 → 2 Li2C3H5Br and ΔH* = 37.5 kJ/mol, ΔS* = −8.4 J/mol K, and ΔG*238 = 40.0 kJ/mol for the reverse reaction 2Li2C3H5Br → Li4(C3H5)2Br2.  相似文献   

2.
The solution structure and the aggregation behavior of (E)-2-lithio-1-(2-lithiophenyl)-1-phenylpent-1-ene ( 1 ) and (Z)-2-lithio-1-(2-lithiophenyl)ethene ( 2 ) were investigated by one- and two-dimensional 1H-, 13C-, and 6Li-NMR spectroscopy. In Et2O, both systems form dimers which show homonuclear scalar 6Li,6Li spin-spin coupling. In the case of 2 , extensive 6Li,1H coupling is observed. In tetrahdrofuran and in the presence of 2 mol of N,N,N′,N′-tetramethylethylylenediamine (tmeda), the dimeric structure of 1 coexists with a monomer. The activation parameters for intra-aggregate exchange in the dimers of 1 and 2 ( 1 (Et2O): ΔH≠ = 62.6 ± 13.9 kJ/mol, ΔS≠ = 5.8 ± 14.0 J/mol K, ΔG≠(263) = 61.1 kJ/mol; 2 (dimethoxyethane): ΔH≠ = 36.9 ± 6.5 kJ/mol, ΔS≠ = ?61 ± 25 J/mol K, ΔG≠(263) = 54.0 kJ/mol) and the thermodynamic parameters for the dimer-monomer equilibrium for 1 (ΔH°; = 26.7 ± 5.5 kJ/mol, ΔS° = 63 ± 27 J/mol K), where the monomer is favored at low temperature, were determined by dynamic NMR studies.  相似文献   

3.
Improved full ab initio optimizations of the molecular structure of biphenyl in twisted minimum energy, coplanar, and perpendicular conformations by use of Poles's GAUSSIAN 82 program have been performed in the 6-31G basis set. These lead to geometries and energies of much higher reliability than our earlier STO-3G results. The torsional angle Φmin obtained now is 45.41° in close agreement with the recent experimental value of 44.4° ± 1.2°. Calculated CC distances may be converted to experimental ED rg-values by means of independently determined linear regression correlations with very high statistical confidence, although they agree better with experimental x ray data for coplanar biphenyl without this correction. Calculated intramolecular angles are very similar for both STO-3G and 6-31G basis sets. The calculated torsional energy barrier towards Φ = 90° (ΔE90) is 6.76 kJ/mol in close agreement with the experimental-31G value of 6.5 ± 2.0 kJ/mol. For coplanar biphenyl with D2h-symmetry the calculated torsional energy barrier ΔE0 is 13.26 kJ/mol which is surprisingly much higher than the experimental value of 6.0 ± 2.1 kJ/mol. This discrepancy could not be resolved by optimizations assumed for two kinds of distortions of planarity of orthohydrogens from the molecular plane of the coplanar carbon atoms. But for the twisted minimum energy conformation asymmetric bending of ortho-H atoms lead to a torsional angle Φmin = 44.74° together with a dihedral angle towards ortho-H of 1.22°, and consequently even to an increase of torsional energy barriers to ΔE0 = 13.51 and ΔE90 = 6.91 kJ/mol.  相似文献   

4.
The 1H and 13C NMR behavior of the monosulfoxides of 1,2-, 1,3-and 1,4-dithianes (1–3) were studied in order to determine the conformational preference of the S→O bond in these heterocycles. From the results of variable temperature, double irradiation, solvent effects and shift reagent experiments, It is concluded that the axial conformers dominate the conformational equilibria of 1 and 3. On the other hand, 2-equatorial is more stable than 2-axial by 0.64 kcal/mol (ΔG°) at -80°, in CD3OD. This value is essentially identical with the one determined in CHClF2, and the lack of a solvent effect appears to indicate that dipole/dipole interactions do not control this equilibrium. AΔGc3 = 11.0 kcal/mol was determined for the inversion process of 2. Complete 1H and 13C NMR assignments for 1–3 are presented.  相似文献   

5.
Methylmaleic (citraconic, CTA) acid and methylfumaric (measaconic, MSA) acid in aqueous sulfuric acid solution undergo bromine-catalyzed reversible cis-trans isomerization in the presence of ceric and bromide ions. The positional isomerization of CTA or MSA to itaconic acid (ITA) is not observed. The method of high performance liquid chromatography (HPLC) was applied to study the kinetics of this catalyzed isomerization. The major catalytic species is best expressed as the Br?2 · radical anion. Under suitable catalytic conditions, there is a tendency for the [MSA]/[CTA] ratio to reach an equilibrium value of 4.10 at 25° for the CTA+Br?2 · ? MSA+Br?2 · reaction. Chloromaleic (CMA) and chlorofumaric (CFA) acids undergo similar isomerization with an equilibrium [CFA]/[CMA] ratio of 10.3 at 25°. The isomerization of maleic acid (MA) to fumaric acid (FA) is essentially irreversible with 50 as the lower limit of the equilibrium [FA]/[MA] ratio. The substituent has an important effect on the reversibility of this catalyzed isomerization of butenedicarboxylic acids. The thermodynamic parameters ΔH° and ΔS° at 25° for the CTA+Br?2 · ? MSA+Br?2 · reaction were found to be ?5.1±0.7 kj/mol and ?6.0±3.3 J/mol K, respectively. The present method gives a plausible way to measure the differences in enthalpy and entropy between the trans- and cis-isomers of butenedicarboxylic acids (CRCO2H=CR'CO2H) in aqueous solution.  相似文献   

6.
The energetic complex, [Cd(HTRTR)2(H2O)4](HTNR)2 {HTRTR = 4‐[3‐(1,2,4‐triazol‐yl)‐1,2,4‐triaozle; HTNR = styphnic acid anion) was synthesized and characterized by FT‐IR spectroscopy, elemental analysis, and single‐crystal X‐ray diffraction. It crystallizes triclinic in space group P$\bar{1}$ [a = 8.156(2) Å, b = 8.374(2) Å, c = 13.267(4) Å, α = 84.925(11)°, β = 87.016(11)°, γ = 63.683(5)°, V = 808.9(4) Å3, ρ = 1.940 g · cm–3]. The CdII ion is six‐coordinate with two HTRTRs and four water molecules. The thermal stabilities were investigated by differential scanning calorimetry (DSC). Non‐isothermal reaction kinetic parameters were calculated by Kissinger's and Ozawa‐Doyle's methods to obtain EK = 144.0 kJ · mol–1, lgAK = 14.22, and EO = 144.3 kJ · mol–1. The values of thermodynamic parameters, the peak temperature while β→0 (Tp0), free energy of activation (ΔG), entropy of activation (ΔS), and enthalpy of activation (ΔH) were obtained. Additionally, the enthalpy of formation was calculated by Hess's law on the basis of the experimental constant‐volume heat of combustion measured by bomb calorimetry, obtaining ΔfH°298 = 4985.5 kJ · mol–1. Finally, the sensitivities toward impact and friction were assessed according to relevant methods. The result indicates it as an insensitive energetic material.  相似文献   

7.
Abstract— At 293 K the long-wavelength absorption and emission band of 1.4 μM allophycocyanin is decreased by estriol (Δ1-3-5(10)-estratriene-3,16α, 17β-trio!) in the range 0.8-6.6 μM in the presence of 11% alcohol (vol/vol). The binding of estriol is shown to be of high affinity, 1:1 with allophycocyanin. The free energy of this binding process (ΔG°) is -33.6 kJ mol' and single binding site dissociation constant (KD) 1.0 ×10–6M. Estriol at 21 μM effectively quenches the fluorescence of 1.4 M large molecular weight phytochrome in its red absorbing form at 77 K while having little or no effect on the phototransformation difference spectrum at 293 K.  相似文献   

8.
γ-tubulin plays crucial role in the nucleation and organization of microtubules during cell division. Recent studies have also indicated its role in the regulation of microtubule dynamics at the plus end of the microtubules. Moreover, γ-tubulin has been found to be over-expressed in many cancer types, such as carcinomas of the breast and glioblastoma multiforme. These studies have led to immense interest in the identification of chemical leads that might interact with γ-tubulin and disrupt its function in order to explore γ-tubulin as potential chemotherapeutic target. Recently a colchicine-interacting cavity was identified at the interface of γ-tubulin dimer that might also interact with other similar compounds. In the same direction we theoretically investigated binding of a class of compounds, noscapinoids (noscapine and its derivatives) at the interface of the γ-tubulin dimer. Molecular interaction of noscapine and two of its derivatives, amino-noscapine and bromo-noscapine, was investigated by molecular docking, molecular dynamics simulation and binding free energy calculation. All noscapinoids displayed stable interaction throughout simulation of 25 ns. The predictive binding free energy (ΔGbind) indicates that noscapinoids bind strongly with the γ-tubulin dimer. However, bromo-noscapine showed the best binding affinity (ΔGbind = –37.6 kcal/mol) followed by noscapine (ΔGbind = –29.85 kcal/mol) and amino-noscapine (ΔGbind = –23.99 kcal/mol) using the MM-PBSA method. Similarly using the MM-GBSA method, bromo-noscapine showed highest binding affinity (ΔGbind = –43.64 kcal/mol) followed by amino-noscapine (ΔGbind = –37.56 kcal/mol) and noscapine (ΔGbind = –34.57 kcal/mol). The results thus generate compelling evidence that these noscapinoids may hold great potential for preclinical and clinical evaluation.  相似文献   

9.
The kinetics of isothermal anionic homopolymerization of ß–myrcene (MYR) and 4–methylstyrene (4MS) in cyclohexane, initiated by n–butyllithium was studied at different temperatures (55, 63, and 71 ° C). The kinetic information obtained from the homopolymerizations was used to estimate the parameters of the Eyring equation, ΔH?= 84080 J/mol and ΔS?= ?21.9 J/mol·K for the MYR, and ΔH?= 51250 J/mol and ΔS?= ?116.8 J/mol·K for the 4MS, to calculate the apparent propagation coefficients k p app as a function of temperature. Finally, the parameters obtained for the Eyring equation were validated by applying them in a mathematical model representing the kinetics of nonisothermal (quasi‐adiabatic) polymerization experiments of MYR and 4MS. © 2019 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2019, 57, 2157–2165  相似文献   

10.
The application of a chiral ligand‐exchange column for the direct high‐performance liquid chromatographic enantioseparation of unusual β‐amino acids with a sodium N‐((R)‐2‐hydroxy‐1‐phenylethyl)‐N‐undecylaminoacetate‐Cu(II) complex as chiral selector is reported. The investigated amino acids were isoxazoline‐fused 2‐aminocyclopentanecarboxylic acid analogs. The chromatographic conditions were varied to achieve optimal separation. The effects of temperature were studied at constant mobile phase compositions in the temperature range 5–45°C, and thermodynamic parameters were calculated from plots of lnk or lnα versus 1/T. Δ(ΔH°) ranged from –2.3 to 2.2 kJ/mol, Δ(ΔS°) from –3.0 to 7.8 J mol?1 K?1 and –Δ(ΔG°) from 0.1 to 1.7 kJ/mol, and both enthalpy‐ and entropy‐controlled enantioseparations were observed. The latter was advantageous with regard to the shorter retention and greater selectivity at high temperature. Some mechanistic aspects of the chiral recognition process are discussed with respect to the structures of the analytes. The sequence of elution of the enantiomers was determined in all cases.  相似文献   

11.
This paper describes the synthesis of four tricyclic heterocycles, furo[2,3–6:4,5-c']- ( 5a ), furo[3,2-b:4,5-c']- ( 5b ), furo[2,3-c:4,5-c']- ( 5c ) and furo[3,2-c:4,5-c']dipyridine ( 5d ). Starting with 2-formylfuropyridines ( 1a-d ), β-(2-furopyridyl)acrylic acids 2a-d were obtained by condensing with malonic acid. The acrylic acids were converted to the acid azides by reaction with ethyl chloroformate and the subsequent reaction with sodium azide. Heating of the acid azides at 230–240° with diphenylmethane and tributylamine gave tricyclic pyridinones 3a-d , which were converted to the respective chloro derivatives 4a-d by reaction with phosphorus oxychloride. Reduction of the chloro compounds over palladium-charcoal yielded compounds 5a-d respectively. All the compounds 2 to 5 were characterized by elemental analysis and spectral data. The H and 13C nmr and electronic spectral features of the furodipyridines were discussed comparing with those of the parent furopyridines.  相似文献   

12.
Critical concentrations of α-(1→3)-D-glucan L-FV-Ⅱ from Lentinus edodes were studied by viscometry andfluorescence probe techniques. The dependence of the reduced viscosity on concentration of the glucan in 0.5 mol/L NaOHaqueous solutions with or without urea showed two turning points corresponding to the dynamic contact concentration c_s andthe overlap concentration c~* of the polymer. The values of c_s and c~* were found to be 1×10~(-3) g cm~(-3) and 1.1×10~(-2) g cm~(-3),respectively, for L-FV-Ⅱ in 0.5 mol/L NaOH aqueous solutions. The two critical concentrations of L-FV-Ⅱ in 0.5 mol/LNaOH aqueous solutions were also found to be 1.2×10~(-3) g cm~(-3) fbr c_s and 9.2×10~(-3) g cm~(-3) for c~* from the concentrationdependence of phenanthrene fluorescence intensities. The overlap concentration c~* of L-FV-Ⅱ in 0.5 mol/L NaOH aqueoussolutions was lower than that of polystyrene with same molecular weight in benzene, owing to the fact that polysaccharidetends to undergo aggregation caused by intermolecular hydrogen bonding. A normal viscosity behavior of L-FV-Ⅱ in 0.5 mol/L urea/0.5 mol/L NaOH aqueous solutions can still be observed in an extremely low concentration range at 25℃.  相似文献   

13.
The intriguing multi‐ligand compound [Cu(IMI)4Cl]Cl ( 1 ) with the ligand imidazole (IMI) was synthesized and characterized by elemental analysis and FT‐IR spectroscopy. The crystal structure was determined by X‐ray single crystal diffraction and the crystallographic data showed that the compound belongs to the monoclinic P21/n space group [α = 8.847(2) Å, b = 13.210(3) Å, c = 13.870(3) Å, and β = 90.164(3)°]. Furthermore, the CuII ion is five‐coordinated by four nitrogen atoms from four imidazole ligands and a chlorine atom. The thermal decomposition mechanism was determined based on differential scanning calorimetry (DSC) and thermogravimetric (TG‐DTG) analysis. The non‐isothermal kinetics parameters were calculated by the Kissinger's method and Ozawa's method, respectively. The energy of combustion, enthalpy of formation, critical temperature of thermal explosion, entropy of activation (ΔS), enthalpy of activation (ΔH), and free energy of activation (ΔG) were measured and calculated.  相似文献   

14.
A silver nanoparticle decorated poly(thiophene) modified glassy carbon electrode (GCE) was prepared for determination of caffeic acid. The Ag/PTh/GCE surface was characterized by scanning electron microscopy (SEM) and energy‐dispersive X‐ray (EDX) spectroscopy. The modified electrode has shown higher electrocatalytic activity towards the oxidation of caffeic acid. The peak current of was found linear in the concentration range from 1.00×10?8 to 4.83×10?6 M with a detection limit of 5.3×10?9 M (S/N=3). The modified electrode was used for determination of CA concentration in red wine samples. The thermodynamic constants, entropy change (ΔS), enthalpy change (ΔH) and Gibbs free energy change (ΔG) were calculated as ?166.34 J/(mol K), ?154.24 kJ/mol and ?104.75 kJ/mol at 25 °C, respectively.  相似文献   

15.
In order to enhance the thermal stability of the barium salt of 5,5′‐bistetrazole (H2BT), carbohydrazide (CHZ) was used to build [Ba(CHZ)(BT)(H2O)2]n as a new energetic coordination compound by using a simple aqueous solution method. It was characterized by FT‐IR spectroscopy, elemental analysis, and single‐crystal X‐ray diffraction. The crystal belongs to the monoclinic P21/c space group [a = 8.6827(18) Å, b = 17.945(4) Å, c = 7.2525 Å, β = 94.395(2)°, V = 1126.7(4) Å3, and ρ = 2.356 g · cm–3]. The BaII cation is ten‐coordinated with one BT2–, two shared carbohydrazides, and four shared water molecules. The thermal stabilities were investigated by differential scanning calorimetry (DSC) and thermal gravity analysis (TGA). The dehyration temperature (Tdehydro) is at 187 °C, whereas the decomposition temperature (Td) is 432 °C. Non‐isothermal reaction kinetics parameters were calculated by Kissinger's method and Ozawa's method to work out EK = 155.2 kJ · mol–1, lgAK = 9.25, and EO = 158.8 kJ · mol–1. The values of thermodynamic parameters, the peak temperature (while β → 0) (Tp0 = 674.85 K), the critical temperature of thermal explosion (Tb = 700.5 K), the free energy of activation (ΔG = 194.6 kJ · mol–1), the entropy of activation (ΔS = –66.7 J · mol–1), and the enthalpy of activation (ΔH = 149.6 kJ · mol–1) were obtained. Additionally, the enthalpy of formation was calculated with density functional theory (DFT), obtaining ΔfH°298 ≈ 1962.6 kJ · mol–1. Finally, the sensitivities toward impact and friction were assessed according to relevant methods. The result indicates the compound as an insensitive energetic material.  相似文献   

16.
The formation of 1 : 2 titanium(III) complex with chromotropic acid (4, 5-dihydroxy-2, 7-naphthalene-disulfonic acid) was observed by spectrophotometric measurements at various ionic strengths. An expression, [Ti(III)]/D=1/Δ? + αH2+/KΔ?[H2R2?]2, was derived for the determination of the formation constant, K=7.2×102 liter2 mol?2 for the Ti(III).(HR)2 ion in the pH range of 1.3–1.8 at constant ionic strength, I=0.2 M, at 25°C. The thermodynamic data for the reaction, Ti(III)+2H3R2?=Ti(III) (HR)2+2H+, were calculated to be ΔG° = ?16 kJ mol?1 ΔH° = 18 kJ mol?1, ΔS° = 110 JK?1 mol?1, at 25°C.  相似文献   

17.
Isothermal kinetics of copper (ion) binding to poly(acrylic acid) (PAA) hydrogel at 20, 25, 35 and 45°C was investigated. Isothermal conversions and kinetic curves of Cu2+ binding to the PAA hydrogel were determined. It was found that the well-known kinetic models of Peppas cannot be applied to describing the entire process of Cu2+ binding. The new method for the determination of the kinetic model of the Cu2+ binding process, as well as the activation energy density distribution functions of PAA hydrogel interaction with Cu2+, were established. It was found that Cu2+ diffusion to the active centers (with E a = 9 kJ/mol) has a dominant influence on the kinetics of the process at temperatures T ≥ 30°C, but at T ≥ 30°C and for the degree of bound Cu2+ α ≥ 0.2, the interaction of Cu2+ from the adsorption center with E a = 26 kJ/mol is dominant. The text was submitted by the authors in English.  相似文献   

18.
The inhibition of the β‐glucosidases from sweet almonds and Caldocellum saccharolyticum at varying pH values by the glucosamine‐related inhibitors 1 – 7 has been compared to the inhibition by the known glucose analogues 8 – 14 . The amino derivatives 3 , 4 , 6 , and 7 were prepared in one step from the known 15 – 18 (Scheme 1), and the amino‐1,2,3‐triazole 5 by a variant of the synthesis leading to the glucose analogue 12 (Scheme 2). The key step for the preparation of the aminoimidazole 1 and of the amino‐1,2,4‐triazole 2 is the regioselective cleavage of the benzyloxy group at C(2) of the gluconolactam 35 and the mannonolactam 57 , respectively, by BCl3 and Bu4NBr (Schemes 3 and 4, resp.). The pH optimum for the inhibition by the amines is lower than their pKHA values, evidencing that they are bound as ammonium salts and that H‐bonding between C(2)−NH and the cat. base B contributes more strongly to binding than any possible H‐bond to the NH2−C(2) group. The influence of the ammonium group on the inhibitory strength correlates with the basicity of the `glycosidic heteroatom'. The strongest increase of the inhibitory strength is observed for the amines lacking a `glycosidic heteroatom' (ΔΔG(OH→NH)=−1.5 to −2.9 kcal/mol). The increase is less pronounced for the amino derivatives 3 – 4 , which possess a weakly basic `glycosidic heteroatom' (ΔΔG(OH→NH)=−0.6 to −1.1 kcal/mol); the amino compounds 1 and 2 , which possess a strongly basic `glycosidic heteroatom', are weaker inhibitors than the corresponding hydroxy compounds, as expressed by ΔΔG(OH→NH) between +4.3 and +4.7 kcal/mol for the amino‐imidazole 1 , and between +2.3 and 2.8 kcal/mol for the amino‐1,2,4‐triazole 2 , denoting the dominant detrimental influence of a C(2)−NH group on the H‐bond acceptor properties of a sufficiently basic `glycosidic heteroatom'.  相似文献   

19.
The multi‐ligand coordination compound copper(II) 1,2‐diaminopropane (pn) azide, [Cu(pn)(N3)2]n ( 1 ), was synthesized using pn and azido groups. It was characterized by X‐ray single crystal diffraction, elemental analysis, and FT‐IR spectroscopy. The crystal structure of 1 belongs to the monoclinic system, space group C2/c. The copper(II) cation is six‐coordinated by one pn molecule and four azido ligands with μ‐1 and μ‐1,1,3 coordination modes. Thermogravimetric investigations with a heating rate of 10 K · min–1 under nitrogen showed one main exothermic stage with a peak temperature of 215.7 °C in the DSC curve. The non‐isothermal kinetics parameters were calculated by Kissinger and Ozawa methods, respectively. The heat of combustion was measured by oxygen bomb calorimetry, and the enthalpy of formation, the critical temperature of thermal explosion, the entropy of activation (ΔS), the enthalpy of activation (ΔH), and the free energy of activation (ΔG) were calculated. The measurements showed that 1 has very high impact, friction, and flame sensitivities.  相似文献   

20.
The flash photolysis of azo?n?propane and of azoisopropane has been studied by kinetic spectroscopy. Transient absorption spectra in theregion of 220–260 nm have been assigned to the n-propyl and isopropyl radicals. For the n-propyl radical, ?max = 744 ± 39 l/mol cm at 245 nm and the rate constants for the mutual reactions were measured to be kc = (1.0 ± 0.1) × 1010 l/mol sec (combination) and kd = (1.9 ± 0.2) × 109 l/mol sec (disproportionation). For the isopropyl radical, ?max = 1280 ± 110 l/mol cm at 238 nm, with kc = (7.7 ± 1.6) × 109 l/mol sec and kd = (5.0 ± 1.2) × 109 l/mol sec The rate constant for the dissociation of the vibrationally excited triplet state of the azopropanes into radicals was measured from the variation in the quantum yield of radicals with pressure. For azo-n-propane k = (6.6 ± 1.3) × 107 sec?1, and for azoisopropane k = (1.6 ± 0.4) × 108 sec?1. Collisional deactivation of the vibrationally excited singlet and triplet states was found to occur on every collision for n-pentane; but nitrogen and argon were inefficient with a rate constant of 1.1 × 1010 l/mol sec. Spectra observed in the region of 220–260 and 370–400 nm areattributed to the cis isomers of the parent trans-azopropanes. These are formed, as permanent products, in increasing amounts as the pressure is increased.  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