首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到19条相似文献,搜索用时 203 毫秒
1.
对于非离子表面活性剂聚氧乙烯月桂醚(Brij-35)/N,N-二甲基甲酰胺(DMF)/长链醇(庚醇,辛醇,壬醇,癸醇)体系,利用滴定微量量热仪测定了胶束形成过程的热功率-时间曲线.根据热力学理论,测定了临界胶束浓度和胶束形成热(ΔHmθ),计算了热力学函数(ΔGmθ和ΔSmθ).讨论了温度、醇中的碳原子数、醇的浓度与临界胶束浓度和热力学函数之间的关系.结果表明:聚氧乙烯月桂醚(Brij-35)/DMF/长链醇体系:(1)在含有相同浓度的各种醇的体系中,ΔHmθ和ΔSmθ的值随着温度的升高而增大;CMC,ΔGmθ的值随着温度的升高而降低;(2)在相同温度及相同浓度的醇体系中,CMC,ΔHmθ,ΔGmθ和ΔSmθ的值都随着醇中碳原子数的增加而降低;(3)在相同温度及相同醇的体系中,CMC,ΔHmθ,ΔSmθ和ΔGmθ的值随着醇的浓度的增加都减小.  相似文献   

2.
用微量量热法测定了阳离子表面活性剂十六烷基三甲基溴化铵(CTAB)在非水溶剂N,N二甲基乙酰胺(DMA)中,分别加入长链醇(正庚醇、正辛醇、正壬醇、正癸醇)体系的热功率一时间曲线,由测得的曲线上的数据得到了临界胶束浓度(CMC)和形成热(ΔHθm).根据热力学理论,计算了热力学函数(ΔGθm,ΔSθm),讨论了温度、醇中的碳原子数、醇的浓度与临界胶束浓度(CMC)和热力学函数之间的关系.结果表明:十六烷基三甲基溴化铵(CTAB)的DMA溶液,在含有相同浓度的各种醇的体系中,CMC,ΔHθm和ΔSθm的值随着温度的升高而增加;而ΔGθm的值随着温度的升高而略有降低.在相同温度及相同浓度的醇体系中,CMC,ΔHθm,ΔGθm和ΔSθm的值都随着醇中碳原子数的增加而降低.在相同温度及相同醇的体系中,CMC和ΔGθm的值随着醇的浓度的增加而增加;而ΔHθm和ΔSθm的值随着醇的浓度的增加而减少.  相似文献   

3.
醇溶剂对羰基(C=O)紫外吸收的定量影响   总被引:1,自引:0,他引:1  
以各种醇为溶剂测定了丙酮和环己酮的羰基紫外吸收能量△E,讨论了△E与醇分子结构的关系。结果表明△E与醇分子中烷基的极化效应指数PEI有良好的线性关系:△E=a bPEI  相似文献   

4.
本文计算的σ_(ΔE)~(HO)和σ_(ΔE)~(LU)值,应用于苯系Ph(x)_n六十多个分子的亲电或亲核反应活性比较,结果同文献实验数据。σ_(ΔE)~(HO)值与分子间或分子内亲电反应活性呈正平行关系,而σ_(ΔE)~(LU)值则正好相反,但是CN、CF_3和F取代基则例外。  相似文献   

5.
张洪林  孔哲  闫咏梅  李干佐  于丽  李真 《化学学报》2007,65(10):906-912
N,N-二甲基乙酰胺(DMA)/长链醇非水溶液体系中, 利用微量量热仪, 研究阴离子表面活性剂十二烷基羧酸钠(SLA)、十二烷基硫酸钠(SDS)的临界胶束浓度(CMC)和热力学函数. 本文在十二烷基羧酸钠, 十二烷基硫酸钠的N,N-二甲基乙酰胺溶液中, 分别加入长链醇(庚醇、辛醇、壬醇、癸醇), 测定体系的热功率-时间曲线. 借助热力学理论, 由测得曲线, 进一步得到临界胶束浓度和热力学函数(ΔHm0, ΔGm0和ΔSm0). 讨论了温度、醇的碳原子数目、醇的浓度与热力学参数之间的关系. 结果表明, 对十二烷基羧酸钠或十二烷基硫酸钠的DMA溶液, 在含有相同浓度的各种醇的体系中, CMC, ΔHm0和ΔSm0的值随着温度的升高而增加, 而ΔGm0的值随着温度的升高而降低. 在相同温度及相同浓度的醇体系中, CMC, ΔHm0Gm0和ΔSm0的值都随着醇中碳原子数目的增加而降低. 在相同温度及相同醇的体系中, CMC, ΔGm0的值随着醇的浓度的增加而增大, 而ΔHm0, ΔSm0的值随着醇的浓度的增加而减少.  相似文献   

6.
为了探索密度泛函理论(DFT)方法中氮苄叉基苯胺分子π电子离域的本质, 介绍了将非平面分子氮苄叉基苯胺分子的DFT能量分成π和σ的方法, 并将π和σ电子能量分成单电子能部分: 动能ΔEπK(θ), ΔEσK(θ)和位能ΔEπP(θ), ΔEσP(θ); 双电子相互作用部分: 库仑作用ΔEππJ(θ), ΔEσσJ(θ), ΔEπσJ(θ)和交换相关作用ΔEππXC(θ), ΔEσσXC(Δ)以及ΔEπσXC(θ), 分析了垂直离域能ΔEV的稳定性及π电子离域对π和σ体系的影响. 在B3LYP/6-31G*, 6-311G*, 6-31G(2d), 6-311G(2d)水平下的计算结果表明, 与经典观点不同, π电子的离域是失稳定的, 且平面时失稳定性最强; 分析各个能量分量表明, 在π电子的离域过程中, π和σ体系均对基组较敏感, π体系本身单电子能的影响大于σ体系, π电子离域对双电子部分作用的影响主要体现在π-σ的耦合作用上.  相似文献   

7.
研制出以固态AgCl~-KCl~-聚乙烯醇-琼脂混合物为导电凝胶的全固态Ag/AgCl参比电极(AllState-Referencr Electrode,简写为ASRE)。将ASRE与pH玻璃电极组成电极组,直接电位法测定pH2.00~12.00磷酸盐缓冲溶液的pH值,与以饱和甘汞电极为参比电极(SCE)的测定结果比较,相对误差为-0.8%~+0.8%;与氯离子选择电极组合测定1.00×10~(-1)~1.00×10~(-4) mol/L NaCl的电池电动势ΔE,-lgc(Cl~-)与ΔE呈良好的线性关系,斜率为-49.3,与以双盐桥SCE为参比电极时的斜率(-49.9)基本一致,表明ASRE能够代替双盐桥SCE运用于Cl~-浓度的测定。在10~80℃范围内,以KNO_3溶液(1.000 mol/L)为模型,测定ASRE相对于SCE的ΔE,ΔE与T呈线性关系,温度系数为0.123mV/℃。ASRE有望代替传统参比电极,应用于离线或在线电化学测定中。  相似文献   

8.
溶胶-凝胶非标记免疫传感器检测乙肝表面抗原   总被引:5,自引:0,他引:5  
采用溶胶 凝胶 (sol gel)技术包埋乙肝表面抗体 (HBsAb) ,涂布于金盘电极表面 ,构成sol gel HBsAb/Au非标记免疫传感器 ,用于检测人血清中乙肝表面抗原 (HBsAg)。该传感器对HBsAg的电位响应遵循Nernst方程 ,在 1~ 330 μg/L浓度范围内 ,传感器的电位响应值ΔE与HBsAg浓度C的对数呈线性关系 ,线性回归方程为ΔE =1 8.1 7+79.84lgC。响应时间为 3min。癌胚抗原、甲胎蛋白等对测定无明显影响。对于HBsAg阴性血清 ,电位响应值ΔE <30mV ,而对于阳性血清则ΔE >30mV ,据此 ,作为临床判别的依据。对1 0 0例临床血清分别用传感器和酶联免疫法 (ELISA)进行双盲检验 ,两法的符合率为 86 %。  相似文献   

9.
将高分离快速液相色谱-四极杆-飞行时间质谱(RRLC-Q-TOF-MS)联用技术用于人参二醇型(PPD)皂苷Rb_1,Rb_2和Rc在酸性条件下的化学转化研究,并对人参炮制过程中人参二醇型皂苷Rb_1,Rb_2和Rc及其转化产物的相对含量进行了分析.利用RRLC-Q-TOF-MS联用和串联质谱(MS/MS)技术对化合物的保留时间、精确分子量及串联质谱碎片信息进行分析,以鉴定化合物的结构.研究结果表明,人参二醇型皂苷在酸性条件下的化学转化包括:取代糖基的水解反应、Δ20(21)或Δ20(22)位的脱水反应和C24,C25位的水合加成反应.在MS/MS分析中,质谱峰m/z 459,477和441分别为人参二醇苷元、C24,C25位水合人参二醇苷元和Δ20(21)或Δ20(22)位脱水人参二醇型苷元的特征离子,这为人参二醇型皂苷及其转化产物的结构鉴定提供了依据,并以此总结了人参二醇型皂苷的化学转化途径.还利用所建立的方法研究了生晒参和红参(100和120℃)中PPD人参皂苷在炮制过程中的变化.  相似文献   

10.
合成了2种含硅二胺单体3,3'-(1,3-二甲基-1,3-二乙烯基-1,3-二硅氧烷)二苯胺(S1)和4,4'-(1,3-二甲基-1,3-二乙烯基-1,3-二硅氧烷)二苯胺(S2),并对其作为环氧树脂的新型固化剂进行了研究.采用非等温示差扫描量热技术(DSC)研究了其与双酚A型环氧树脂E51体系的固化反应动力学,根据不同升温速率下E51/S1和E51/S2体系的特征温度的变化,分别确定了该两体系的固化反应工艺条件,即:E51/S1体系为100℃/1h+160℃/2h+190℃/3h;E51/S2体系为110℃/1h+170℃/2h+190℃/3h.通过Kissinger方程、Crane方程以及Arrheninus方程对固化反应进行了固化动力学行为研究,得到了两个反应体系的表观活化能ΔE、Arrhenius因子A以及反应级数n等动力学参数.E51/S1体系的ΔE为50.65 k J/mol、A为1.83×105,n为0.87;E51/S2体系的ΔE为51.39 k J/mol,A为1.44×105,n为0.87,由ΔE和A表明,E51和S2的反应活性高于E51和S1的反应活性,即,氨基位于苯环的间位时与环氧基团更容易反应.二者的反应级数相同并且小于1,说明E51与S1和S2之间的反应是复杂反应.在动力学参数的基础上,得到了n级固化动力学模型.  相似文献   

11.
自溶液中的吸附 VII: 硅胶自环己烷中吸附醇,酮和酯   总被引:3,自引:0,他引:3  
赵振国  顾惕人 《化学学报》1983,41(12):1091-1099
The adsorption isotherms of some monofunctional alcohols, ketones and esters from cyclohexane onto silica gel have been determined at 30`C and 10`C. The silica gel used bad a BET area of 417 m2/g and an average pore radius of 45A. The concentrations of free and associated hydroxyls on the silica gel were 1.4 and 4.1/100A^2 respectively. The adsorption order is cyclohexanol>n-octyl alcohol>cyclohexanone>methyl isobutyl ketone>n-propyl acetate=n-amyl acetate. The adsorption decreases with increasing temperature as normal. Except at very low concentrations, the isotherms can be represented by the Langmuir equation. The limiting adsorption, nms, on the silica gel does not accord with the stoichiometric ratio (1:1) between the free surface hydroxyl groups of the adsorbent and the adsorbate. In addition to surface conditions and the functional group of the adsorbate, it seems that the limiting adsorption is also controlled by the other factors, including temperature, solvent, and, sometimes the chain length of the adsorbate. The standard free energy (ΔG0) and standard enthalpy (ΔH0) of the adsorbates in adsorptien processes have been determined from the Langmuir parameters. The results indicate that the absolute value of ΔH0 is higher than that of ΔG0 (in other words, standard entropy ΔS0 is negative), as in the case of the adsorption of gases. Since there is practically no difference in ΔG0 or ΔH0 of adsorption between alcohols or esters, it is suggested that in dilute solution only the polar groups take part in adsorption, and the hydrocarben chains remain in solution during the adsorption process. For ketones, the absolute values of ΔG0 and ΔH0 are somewhat lower for methyl isobutyl ketone than that for cyclohexanone. A possible explanation is that in the adsorbed state the isobutyl chain of the methyl isobutyl ketone may somewhat close to the surface and thus decreases the adsorption and changes the ΔG0 and ΔH0.  相似文献   

12.
We calculate the ground state and excited state second-order dispersion interactions between parallel π-conjugated polymers. The unperturbed eigenstates and energies are calculated from the Pariser-Parr-Pople model using CI-singles theory. Based on large-scale calculations using the molecular structure of trans-polyacetylene as a model system and by exploiting dimensional analysis, we find that: (1) For inter-chain separations, R, greater than a few lattice spacings, the ground-state dispersion interaction, ΔE(GS), satisfies, ΔE(GS)~L(2)/R(6) for L ? R and ΔE(GS)~L/R(5) for R ? L, where L is the chain length. The former is the London fluctuating dipole-dipole interaction while the latter is a fluctuating line dipole-line dipole interaction. (2) The excited state screening interaction exhibits a crossover from fluctuating monopole-line dipole interactions to either fluctuating dipole-dipole or fluctuating line dipole-line dipole interactions when R exceeds a threshold R(c), where R(c) is related to the root-mean-square separation of the electron-hole excitation. Specifically, the excited state screening interaction, ΔE(n), satisfies, ΔE(n) ~ L∕R(6) for R(c) < L ? R and ΔE(n) ~ L(0)∕R(5) for R(c) < R ? L. For R < R(c) < L, ΔE(n) ~ R(-ν), where ν ? 3. We also investigate the relative screening of the primary excited states in conjugated polymers, namely the n = 1, 2, and 3 excitons. We find that a larger value of n corresponds to a larger value of ΔE(n). For example, for poly(para-phenylene), ΔE(n = 1) ? 0.1 eV, ΔE(n = 2) ? 0.6 eV, and ΔE(n = 3) ? 1.2 eV (where n = 1 is the 1(1)B(1) state, n = 2 is the m(1)A state, and n = 3 is the n(1)B(1) state). Finally, we find that the strong dependence of ΔE(n) on inter-chain separation implies a strong dependency of ΔE(n) on density fluctuations. In particular, a 10% density fluctuation implies a fluctuation of 13 meV, 66 meV, and 120 meV for the 1(1)B(1), m(1)A state, and n(1)B(1) states of poly(para-phenylene), respectively. Our results for the ground-state dispersion are applicable to all types of conjugated polymers. However, our excited state results are only applicable to conjugated polymers, such as the phenyl-based class of light emitting polymers, in which the primary excitations are particle-hole (or ionic) states.  相似文献   

13.
In this study, our investigations showed that the increasing concentrations of all examined mono alcohols caused a decrease in the V m, k cat and k cat/K m values of Bacillus clausii GMBE 42 serine alkaline protease for casein hydrolysis. However, the K m value of the enzyme remained almost the same, which was an indicator of non-competitive inhibition. Whereas inhibition by methanol was partial non-competitive, inhibition by the rest of the alcohols tested was simple non-competitive. The inhibition constants (K I) were in the range of 1.32–3.10 M, and the order of the inhibitory effect was 1-propanol>2-propanol>methanol>ethanol. The ΔG and ΔG E???T values of the enzyme increased at increasing concentrations of all alcohols examined, but the ΔG ES value of the enzyme remained almost the same. The constant K m and ΔG ES values in the presence and absence of mono alcohols indicated the existence of different binding sites for mono alcohols and casein on enzyme the molecule. The k cat of the enzyme decreased linearly by increasing log P and decreasing dielectric constant (D) values, but the ΔG and ΔG E???T values of the enzyme increased by increasing log P and decreasing D values of the reaction medium containing mono alcohols.  相似文献   

14.
The exciplex is a charge transfer species formed in the process of electron transfer between an electron donor and an electron acceptor and hence is very sensitive to solvent polarity. In order to understand the role of solvent in exciplex formation between pyrene (PY) and 4,4′‐bis(dimethylamino)diphenylmethane (DMDPM), we used two types of solvent approximations: an implicit solvent model and an explicit solvent model. The difference in energies between the excited and the meta‐stable Frank–Condon state (ΔE) of the structures were assumed to correspond to the emission maximum of the exciplex in different solvents. The ΔE values show the trend of stabilization of the exciplex with an increase in solvent polarity. This trend in stabilization is substantially more prominent in the explicit solvent model than that with the implicit solvent model. The ΔE value obtained in methanol reflects equal stabilization compared to that in a more polar solvent, N,N‐dimethylformamide. This extra stabilization of the exciplex may be explained on the basis of the H‐bonding capability of the protic solvent, methanol. © 2004 Wiley Periodicals, Inc. Int J Quantum Chem, 2005  相似文献   

15.
Quantum-chemical calculations using DFT and ab initio methods have been carried out for 32 carbenes RR'C which comprise different classes of compounds and the associated ketenes RR'C═C═O. The calculated singlet-triplet gaps ΔE(S-T) of the carbenes exhibit a very high correlation with the bond dissociation energies (BDEs) of the ketenes. An energy decomposition analysis of the RR'C-CO bond using the triplet states of the carbene and CO as interacting fragments supports the assignment of ΔE(S-T) as the dominant factor for the BDE but also shows that the specific interactions of the carbene may sometimes compensate for the S/T gap. The trend of the interaction energy ΔE(int) values is mainly determined by the Pauli repulsion between the carbene and CO. The stability of amino-substituted ketenes strongly depends on the destabilizing conjugation between the nitrogen lone-pair orbital and the ketene double bonds. There is a ketene structure of the unsaturated N-heterocyclic carbene parent compound NHC1 with CO as a local energy minimum on the potential-energy surface. However, the compound NHC1-CO is thermodynamically unstable toward dissociation. The saturated homologue NHC2-CO has only a very small bond dissociation energy of D(e) = 3.2 kcal/mol. The [3]ferrocenophane-type compound FeNHC-CO has a BDE of D(e) = 16.0 kcal/mol.  相似文献   

16.
The molecular and electronic structures and bonding analysis of terminal cationic metal-ylyne complexes (MeCN)(PMe(3))(4)M≡EMes](+) (M = Mo, W; E = Si, Ge, Sn, Pb) were investigated using DFT/BP86/TZ2P/ZORA level of theory. The calculated geometrical parameters for the model complexes are in good agreement with the reported experimental values. The M-E σ-bonding orbitals are slightly polarized toward E except in the complex [(MeCN)(PMe(3))(4)W(SnMes)](+), where the M-E σ-bonding orbital is slightly polarized toward the W atom. The M-E π-bonding orbitals are highly polarized toward the metal atom. In all complexes, the π-bonding contribution to the total M≡EMes bond is greater than that of the σ-bonding contribution and increases upon going from M = Mo to W. The values of orbital interaction ΔE(orb) are significantly larger in all studied complexes I-VIII than the electrostatic interaction ΔE(elstat). The absolute values of the interaction energy, as well as the bond dissociation energy, decrease in the order Si > Ge > Sn > Pb, and the tungsten complexes have stronger bonding than the molybdenum complexes.  相似文献   

17.
Racemic amino alcohols have been separated as perfluoroacylated derivatives by gas chromatography using either improved Chirasil-Val or heptakis(2,6-di-O-methyl-3-O-pentyl)-β-cyclodextrin as stationary phase. Using Chirasil-Val all the amino alcohols investigated were separated to baseline (α values between 1.03 and 1.08) whereas only a few amino alcohols were resolved on the modified cyclodextrin column. The enantioselectivity obtained on the latter phase was, however, significantly higher. The separations were performed as trifluoroacetyl, pentafluoropropionyl, and heptafluorobutyryl derivatives and the chiral discrimination observed for the different derivatives was significantly different for both stationary phases. In oder to obtain a better understanding of the separation mechanism, the Gibbs-Helmholtz parameters Δ(R,S)ΔH° and Δ(R,S)ΔS° were determined. The most extraordinary result was obtained for the trifluoroacetyl derivative of allo-threoninol. In addition to the order of elution of the enantiomers being the opposite of that for the other compounds, the separation seems to be entropy controlled (the sign Δ(R,S)ΔH° is positive), i.e. the separation improved at higher temperatures.  相似文献   

18.
Low concentrations of alcohols have proven to be able to enlarge the stability curve of globular proteins, by decreasing the cold denaturation temperature and increasing the hot denaturation temperature [S. R. Martin, V. Esposito, P. De Los Rios, A. Pastore and P. A. Temussi, J. Am. Chem. Soc., 2008, 130, 9963-9970]. In order to try to explain these data, I have considered that: (1) an aqueous 2 M MeOH solution can be treated as a uniform liquid, constituted by water molecules, whose density, above the temperature of maximum density, has the same values of neat water, simply shifted by 2 °C toward lower temperatures, whereas, below the temperature of maximum density, it decreases to a slightly lesser extent than the density of neat water; (2) the ΔE(a)(2 M MeOH) quantity, a balance between intra-protein energetic attractions and those with the surrounding solvent molecules, both water and methanol, assumes a constant positive value. These physically-based assumptions, when inserted into the theoretical model developed to rationalize the occurrence of cold denaturation in neat water [G. Graziano, Phys. Chem. Chem. Phys., 2010, 12, 14245-14252], reproduce in a qualitatively correct manner the effect of low concentrations of alcohols.  相似文献   

19.
Density functional theory calculations have been performed for the dimethylgallyl complexes of iron, ruthenium, and osmium [(η(5)-C(5)H(5))(L)(2)M(GaMe(2)] (M = Fe, Ru, Os; L = CO, PMe(3)) at the DFT/BP86/TZ2P/ZORA level of theory. The calculated geometry of the iron complex [(η(5)-C(5)H(5))(CO)(2)Fe(GaMe(2))] is in excellent agreement with structurally characterized complex [(η(5)-C(5)H(5))(CO)(2)Fe(Ga(t)Bu(2))]. The Pauling bond order of the optimized structures shows that the M-Ga bonds in these complexes are nearly M-Ga single bond. Upon going from M = Fe to M = Os, the calculated M-Ga bond distance increases, while on substitution of the CO ligand by PMe(3), the calculated M-Ga bond distances decrease. The π-bonding component of the total orbital contribution is significantly smaller than that of σ-bonding. Thus, in these complexes the GaX(2) ligand behaves predominantly as a σ-donor. The contributions of the electrostatic interaction terms ΔE(elstat) are significantly smaller in all gallyl complexes than the covalent bonding ΔE(orb) term. The absolute values of the ΔE(Pauli), ΔE(int), and ΔE(elstat) contributions to the M-Ga bonds increases in both sets of complexes via the order Fe < Ru < Os. The Ga-C(CO) and Ga-P bond distances are smaller than the sum of van der Waal radii and, thus, suggest the presence of weak intermolecular Ga-C(CO) and Ga-P interactions.  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