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.
The solvent effect on the acidic dissociation of the protonated base tris(hydroxymethyl)aminomethane (Tris) has been studied in aqueous mixtures of several organic constituents with qualitatively similar results. The dissociation constant of protonated 2-amino-2-methyl-1,3-propanediol (Bis), the analog of Tris with two hydroxy groups, has now been determined in 50 mass % methanol by emf measurements of cells without liquid junction in the temperature range 10 to 40°C. The conventional treatment of the data indicated abnormally low values for the activity coefficient of BisHCl, but an alternate method of calculation with allowance for ion pairing provided no advantage. The pKa in the range from T=283.15 to 313.15 K is given by $$pK_a = 3755.63/T - 24.5127 + 3.56855 In T$$ with a standard deviation of 0.0006 at the seven temperatures. Thermodynamic quantities for the dissociation process and corresponding quantities for the transfer from water to 50 mass % methanol were derived. The implications of the results in terms of selective solute-solvent stabilization are discussed.  相似文献   

2.
In order to determine the optimum conditions of potentiometric titration, an investigation has been made of the relative acidities of 13-hydroxyanthraquinones in water, methanol, acetone (ac), dimethylformamide (DMFA), and dimethyl sulfoxide (DMSO) on the basis of a calculation of the indices of the relative acidity constants (pKa) by Henderson's method. The existence of a relationship between pKa in water and pKa in acetone, dimethylformamide, and dimethyl sulfoxide has been established which is characterized by the linear equations $$pK_a^{DMSO} = 1.54pK_a^{H_2 O} + 11.88$$ , $$pK_a^{DMFA} = 1.38pK_a^{H_2 O} + 8.50$$ , $$pK_a^{ac} = 1.11pK_a^{H_2 O} + 10.26$$ . The sequence of neutralization of the hydroxyls in the titration of polyhydroxy-anthraquinones has been determined from the pKa values in DMSO and the results of a calculation of electronic structures by the Pariser-Parr-Pople method. A quantitative evaluation of the conditions of titration in the five solvents on the basis of indices of the titration constants (pKt) has shown that the optimum conditions for the quantitative determination by potentiometric titration are achieved in dimethyl sulfoxide.  相似文献   

3.
The solubility of carefully characterized UO2 in pOH 1.5 and pOH 2.5 aqueous solutions has been determined from 25°C to 300°C using a flow apparatus. Data were analyzed in terms of the reaction $$UO_2 + 2H_2 O + OH^ - \rightleftharpoons U(OH); log K = - 5.86 + 32/T$$ The extreme sensitivity of both the UO2 surface and aqueous U(IV) to oxidation is discussed.  相似文献   

4.
Reactions between CF2 and O(3P) have been studied at 295 K in a gas flow reactor sampled by a mass spectrometer. The major reaction for CF2 has been found to be $$CF_2 + O \to COF + F$$ with $$CF_2 + O \to CO + 2F(F_2 )$$ more than a factor of three slower. The rate coefficient for all loss processes for CF2 on reaction with O is (1.8±0.4)×10?11 cm3 s?1. The COF produced in (18) undergoes a fast reaction with O to produce predominantly CO2. $$COF + O \to CO_2 + F$$ It is uncertain from the results whether or not $$COF + O \to CO + FO$$ occurs, but in any event (19) is the major route. The rate coefficient for the loss of COF in this system [i.e., the combined rate coefficients for (19) and (20)] is (9.3±2.1)×10?11 cm3 s?1. Stable product analysis reveals that for each CF2 radical consumed, the following distribution of stable products is obtained: COF2 (0.04±0.02), CO (0.21±0.04), and CO2 (0.75±0.05). Thus COF2, which we assume is produced via $$CF_2 + O \xrightarrow{M} COF_2$$ is a very minor product in this reaction sequence. The measured rate coefficients demonstrate that reactions (18) and (19) are important sources of F atoms in CF4/O2 plasmas.  相似文献   

5.
The thermal expansion of VN1?x was determined from measurements of the lattice parameters in the temperature range of 298–1000 K and in the composition range of VN0.707–VN0.996. Within the accuracy of the results the expansion of the lattice parameter with temperature is not dependent on the composition. The lattice parameter as a function of composition ([N]/[V]=0.707?0.996) and temperature (298–1000 K) is given by $$\begin{gathered} a([N]/[V],T) = 0.38872 + 0.02488([N]/[V]) - \hfill \\ - (1.083 \pm 0.021) \cdot 10^{ - 4} T^{1/2} + (6.2 \pm 0.1) \cdot 10^{ - 6} T. \hfill \\ \end{gathered} $$ . The coefficient of linear thermal expansion as a function of temperature (in the same range) is given by $$\alpha (T) = a([N]/[V],T)^{ - 1} [( - 5.04 \pm 0.01) \cdot 10^5 T^{ - 1/2} + (6.2 \pm 0.1) \cdot 10^{ - 6} ].$$ . The average linear thermal expansion coefficient is $$\alpha _{av} = 9.70 \pm 0.15 \cdot 10^{ - 6} K^{ - 1} (298 - 1 000K).$$ . The data are compared with those of several fcc transition metal nitrides collected and evaluated from the literature.  相似文献   

6.
In acetonitrile (AN) solutions the gross constants are determined for the reactions $$Ph_3 CCl + MCl_n ANPh_3 C^ + MCl_{n + 1}^ - + AN$$ (MCl n =SbCl5, GaCl3, InCl3, and FeCl3). The relaxation spectra are interpreted for the reactions of metal(III) chlorides according to the equilibria $$\begin{gathered} 2 MCl_3 AN + 6AN \rightleftharpoons [MCl_2 (AN)_4 ]^ + [MCl_4 ]^ - + 4 AN \rightleftharpoons \hfill \\ 2 [MCl_2 (AN)_4 ]^ + Cl - \hfill \\ \end{gathered} $$   相似文献   

7.
A modified topological index \(\tilde Z_G \) is proposed to be defined as $$\tilde Z_G = \sum\limits_{k = 0}^{[N/2]} {( - 1)^k } a_{2k} $$ for characterising theπ-electronic system of a conjugated hydrocarbonG withN carbon atoms, wherea 2k is the coefficient of the characteristic polynomial ofG defined as $$P_G (X) = ( - 1)^N \det |A - XE| = \sum\limits_{k = 0}^N { a_k X^{N - k} } $$ with an adjacency matrixA and the unit matrixE. \(\tilde Z_G \) is identical toZ G for a tree graph, or a chain hydrocarbon.Z G increases with a (4n+2)-membered ring formation and decreases with a 4n-membered ring formation. The totalπ-electron energyE π of the Hückel molecular orbital is shown to be related with \(\tilde Z_G \) asE π =Cln \(\tilde Z_G \) . With this relation generalised and extended Hückel rules for predicting the stability of an arbitrary network are proved.  相似文献   

8.
The formation of complexes between iron(II) and tartrate ion (L) has been studied at 25° C in 1m-NaClO4, by using a glass electrode. The e.m.f. data are explained with the following equilibria: $$\begin{gathered} Fe^{2 + } + L \rightleftarrows FeL log \beta _1 = 1,43 \pm 0,05 \hfill \\ Fe^{2 + } + 2L \rightleftarrows FeL_2 log \beta _2 = 2,50 \pm 0,05 \hfill \\\end{gathered} $$ The protonation constants of the tartaric acid have been determinated: $$\begin{gathered} H^ + + L \rightleftarrows HL logk_1 = 3,84 \pm 0,03 \hfill \\ 2H^ + + L \rightleftarrows H_2 L logk_2 = 6,43 \pm 0,02 \hfill \\\end{gathered}$$ .  相似文献   

9.
Two general equations for estimation of excess enthalpies of ternary systems consisting of an alcohol and two hydrocarbons from observed excess properties of the various binary combinations have been developed. The first expression is based on the Kretschmer-Wiebe association model and takes the form $$\Delta \overline H _{ABC}^{ex} = h_A x_A K_A (\phi _{A1} - \phi _{A1}^o ) + Q_{ABC}$$ where $$\begin{gathered} Q_{ABC} = (x_A + x_B )(\phi _A + \phi _B )(\Delta \overline H _{AB}^{ex} )_{phys}^ \bullet + (x_A + x_C )(\phi _A + \phi _C )(\Delta \overline H _{AC}^{ex} )_{phys}^ \bullet \hfill \\ + (x_B + x_C )(\phi _B + \phi _C )(\Delta \overline H _{BC}^{ex} )_{phys}^ \bullet \hfill \\ \end{gathered}$$ \((\Delta \overline H _{ij}^{ex} )_{phys}^ \bullet\) represents the physical interactions in each of the individual binary systems, and the term involving φ A1 o represents the chemical contributions (caused by self-association) to the excess enthalpies of mixing. The second predictive expression is based on the Mecke-Kempter association model and is given by $$\Delta \overline H _{ABC}^{ex} = - h_A x_A [In(1 + K_A \phi _A )/K_A \phi _A - In(1 + K_A )/K_A ] + Q_{ABC}$$ where the first term (contained within brackets) represetns the chemical contributions to the enthalpies of mixing. The predictions of both expressions are compared with experimental data for the excess enthalpies of six ternary systems.  相似文献   

10.
Pressure dependent UV-visible spectrophotometric measurements were used to determine \(\Delta \bar V^ * \) and \(\Delta \bar \kappa ^ * \) for the formation of alkali metal borate ion pairs. The association constant for each ion pair was measured at 25°C and at ionic strengths of 0.1 and 1.0m over a pressure range of 1 to 2000 atm. The pressure dependence of the apparent association constants, K A (P)/K A (1), have been fitted to $$[RT/(P - 1)]ln[K_A (P)]/[K_A (1)] = - \Delta \bar V^0 + \Delta \bar \kappa ^0 [(P - 1)/2]$$ to determine \(\Delta \bar V^0 and \Delta \bar \kappa ^0 \) . The \(\Delta \bar V^0 \) for the alkali metal borate ion pairs range from 5–9 cm3-mol?1. The association constants were also measured as a function of ionic strength at 1 atm. Extrapolation to I=0 yielded K A of 2.12, 0.66, 0.76 and 1.12 for [LiB(OH)4], [KB(OH)4], [RbB(OH)4] and [CsB(OH)4], respectively. The trend generally indicates less ion pairing and a smaller volume change for the ion pair formation as the size of the cation increases. The concept of localized hydrolysis is used to explain the trend observed in the equilibrium constant of the ion pair as the cation size is changed.  相似文献   

11.
A mechanism for the thermal decomposition of ionic oxalates has been proposed on the basis of three quantitative relationships linking the quantitiesr c/r i (the ratio of the Pauling covalent radius and the cation radius of the metal atom in hexacoordination) andΣI i (the sum of the ionization potentials of the metal atom in kJ mol?1) with the onset oxalate decomposition temperature (T d) (Eq. 1) the average C-C bond distance (¯d) (Eq. 2), and the activation energy of oxalate decomposition (E a) (Eq. 3): (1) $$T_d = 516 - 1.4006\frac{{r_c }}{{r_i }}(\sum I_i )^{\frac{1}{2}}$$ (2) $$\bar d = 1.527 + 5.553 \times 10^{ - 6} \left( {122 - \frac{{r_c }}{{r_i }}(\sum I_i )^{\frac{1}{2}} } \right)^2$$ (3) $$E_a = 127 + 1.4853 \times 10^{ - 6} \left( {\left( {\frac{{r_c }}{{r_i }}} \right)^2 \sum I_i - 9800} \right)^2$$ On the basis of these results it is proposed that the thermal decomposition of ionic oxalates follows a mechanism in which the C-O bond ruptures first. From Eq. 3 it is further proposed that strong mutual electronic interactions between the oxalate and the cations restrict the essential electronic reorganization leading to the products, thereby increasingE a.  相似文献   

12.
Values of the integral heat of solution of iodine in carbon disulfide were determined at different mole ratiosr=n(CS2)/n(I2) in the range 34<r<2650 and 298,15 K by isoperibol calorimetry. The experimental data may be expressed by the empirical equation $$\Delta H_m \prime /cal Mol^{ - 1} = 2973 + 1759\frac{r}{{r + 1}} - 0,0821{\text{ }}r,$$ where ΔH m' is the molar enthalpy change for the process $$I_2 (c) + r{\text{ }}CS_2 (l) = [I_2 ,r{\text{ }}CS_2 ](sol).$$ Since the last term in the above equation can be explained by assuming the presence of a trace impurity in the solvent, the “true” heat of solution is given by $$\Delta H_m /cal Mol^{ - 1} = 2973 + 1759\frac{r}{{r + 1}}.$$ Smoothed values of this quantity are given in table 3 for selected values of the mole ratio,r. The results are discussed in terms of the regular solution theory.  相似文献   

13.
Precise conductance measurements are reported for four xylic acids, 2,3-2,5-2,6 and 3,5-dimethylbenzoics. Limiting molar conductances ΛO and pK a obtained from a Fuoss type analysis, are reported for each of these acids at five degree intervals covering the range from 0° to 100°C. The ΛO values for each acid are described by a polynomial in the Celsius temperature. The pK a were smoothed as a function of the Kelvin temperature T with an equation of the form: $$pK_{\text{a}} = A{\text{ }} + {\text{ }}B/T{\text{ }} + {\text{ }}C{\text{ }}logT{\text{ }} + {\text{ }}DT$$ where the term linear in T was required only for the 2,3-acid. Standard enthalpy, entropy, and heat capacity changes were calculated by suitable differentiation of this equation. Walden products were calculated for the four anions at each of the temperatures and are compared with earlier data for the toluate and the benzoate ions. Those acids with ortho groups undergo a large decrease in enthalpy on ionization and are substantially more acidic than benzoic acid. These effects are especially large for 2,6-dimethylbenzoic acid. A methyl group in the meta position lowers acidity slightly both in m-toluic and in 2,5-dimethylbenzoic acid. However, 2,3-dimethylbenzoic acid is more acidic than o-toluic.  相似文献   

14.
The kinetic equation which describes many electronic as well as atomic or chemical reactions under the condition of a steadily linear raise of the temperature, is considered in a mathematically exact and straightforward way. Therefore, the equation has been transformed into a dimensionsless form, using with profit the maximum condition for the intensity peak. The two temperatures T1 and T2, corresponding to the half-height of the intensity peak, are found as unique polynomials of the small argument \(\bar y \equiv {{k\bar T} \mathord{\left/ {\vphantom {{k\bar T} E}} \right. \kern-0em} E}\) only ( \(\bar T\) =temperature of peak maximum). Thereupon, further combinations give half-widthδ, peak asymmetryA21 or \(\tilde A = {{\bar C} \mathord{\left/ {\vphantom {{\bar C} {(1 - \bar C)}}} \right. \kern-0em} {(1 - \bar C)}}\) and the maximum of the intensity peakJ; they again all depend only on¯y. In some cases this dependence is weak, so that e.g. it is deduced that the half-width energy product divided by \(\bar T^2 \) is an invariant, different for every kinetic orderπ: $$\frac{{\delta \cdot E[eV]}}{{\bar T^2 }} = \left\{ {\begin{array}{*{20}c} {{1 \mathord{\left/ {\vphantom {1 {4998 K for monomolecular process}}} \right. \kern-\nulldelimiterspace} {4998 K for monomolecular process}}} \\ {{1 \mathord{\left/ {\vphantom {1 {3542 K for bimolecular process}}} \right. \kern-\nulldelimiterspace} {3542 K for bimolecular process}}} \\ {{1 \mathord{\left/ {\vphantom {1 {2872 K for trimolecular process}}} \right. \kern-\nulldelimiterspace} {2872 K for trimolecular process}}} \\ \end{array} } \right.$$ By means of these correlations, activation energy valuesE [eV] can be determined accurately to within 0.5 %, so that for most experiments the inaccuracy of theδ values becomes dominant and limiting. A special nomogram for the express estimation ofE from experimentally observedδ and \(\bar T\) is demonstrated.  相似文献   

15.
A solubility cell which can be operated continuously over the temperature range 5–95 °C has been developed. The solubility of Fe0.88S (monoclinic pyrrhotite) in solutions $$S_0 = ([H^ + ]) = H{\text{ }}m,{\text{ }}[Na^ + ] = (1.00---H) m,{\text{ }}[ClO_{4^ - } ] = 1.00 m)$$ at fixed partial pressures of H2S has been investigated at 50.7 °C. The hydrogen ion concentration and the total concentration of iron(II) ion in equilibrium with the solid phase was determined by e.m.f. and analytical methods respectively. The data were consistent with $$\log ^* K_{pso} = \log \frac{{[Fe^{2 + } ]pH_2 S}}{{[H^ + ]^2 }} = 3.80 \pm {\text{ }}0.10{\text{ }}[50.7^\circ C,{\text{ }}1 m(Na)ClO_4 ]$$ according to the overall reaction $$1.14{\text{ }}Fe_{0.88} S_{(s)} {\text{ }} + {\text{ }}2H_{(I = 1m)}^ + {\text{ }} \rightleftharpoons {\text{ }}Fe_{(I = 1m)}^{2 + } {\text{ }} + {\text{ H}}_{\text{2}} S_{(g)} {\text{ }} + {\text{ }}0.14{\text{ }}S_{(s)} $$   相似文献   

16.
The finite set of rate equations C m,n ' n,n-1 C m,n-1 (t)+α n,n C m,n (t)+α n,n+1 C m,n+1 (t), $$0 \leqslant m \leqslant N,0 \leqslant n \leqslant N,$$ where $$\alpha _{i,j}$$ are $\alpha _{j,j - 1} = A,\alpha _{j,j} = - \left( {A + B} \right),\alpha _{j,j + 1} = B$ , with $\alpha _{0,0} = - \alpha _{1,0} = - \alpha$ and $\alpha _{N,N} = - \alpha _{N - 1,N} = - b,\alpha _{0, - 1} = \alpha _{N,N + 1} = 0$ , subject to the initial condition $C_{m,n} \left( 0 \right) = \delta _{n,m}$ (Kronecker delta) for some $m$ , arises in a number of applications of mathematics and mathematical physics. We show that there are five sets of values of $a$ and $b$ for which the above system admits exact transient solutions.  相似文献   

17.
A very precise and accurate new method is described for determination of the Henry coefficient k and the isotopic fractionation of gases dissolved in liquids. It yields fully corrected values for k at essentially infinite dilution. For oxygen the random error for k is less than 0.02%, which is an order of magnitude better than the best previous measurements on that or any other gas. Extensive tests and comparison with other work indicate that systematic errors probably are negligible and that the accuracy is determined by the precision of the measurements. In the virial correction factor (1+λPt), where Pt is the total pressure of the vapor phase, the coefficient λ for oxygen empirically is a linear decreasing function of the temperature over the range 0–60°C. The simple three-term power series in 1/T proposed by Benson and Krause, $$\ln k = a_0 + a_1 /T + a_2 /T^2 $$ provides a much better form for the variation of k with temperature than any previous expression. With a0=3.71814, a1=5596.17, and a2=?1049668, the precision of fit to it of 37 data points for oxygen from 0–60°C is 0.018% (one standard deviation). The three-term series in 1/T also yields the best fit for the most accurate data on equilibrium constants for other types of systems, which suggests that the function may have broader applications. The oxygen results support the idea that when the function is rewritten as $$\ln k = - (A_1 + A_2 ) + A_1 \left( {\frac{{T_1 }}{T}} \right) + A_2 \left( {\frac{{T_1 }}{T}} \right)^2 $$ it becomes a universal solubility equation in the sense that A2 is common to all gases, with T1 and A1 characteristic of the specific gas. Accurate values are presented for the partial molal thermodynamic function changes for the solution of oxygen in water between the usual standard states for the liquid and vapor phases. These include the change in heat capacity, which varies inversely with the square of the absolute temperature and for which the random error is 0.15%. Analysis of the high-temperature data of Stephan et al., in combination with our values from 0–60°C, shows that for oxygen the fourterm series in 1/T, $$\ln k = - 4.1741 + 1.3104 \times 10^4 /T - 3.4170 \times 10^6 /T^2 + 2.4749 \times 10^8 /T^3 $$ where p=kx and p is the partial pressure in atmospheres of the gas, probably provides the best and easiest way presently available to calculate values for k in the range 100–288°C, but more precise measurements at elevated temperatures are needed. The new method permits direct mass spectrometric comparison of the isotopic ratio34O2/32O2 in the dissolved gas to that in the gas above the solution. The fractionation factor α=32k/34k varies from approximately 1.00085 (±0.00002) at 0°C to 1.00055 (±0.00002) at 60°C. Although the results provide the first quantitative determination of α vs. temperature for oxygen, it is not possible from these data to choose among several functions for the variation ofInα with temperature. If the isotopic fractionation is assumed to be due to a difference in the zero-point energy of the two species of oxygen molecules, the size of the solvent cage is calculated to be approximately 2.5 Å. The isotopic measurements indicate that substitution of a34O2 molecule for a32O2 molecule in solution involves a change in enthalpy with a relatively small change in entropy.  相似文献   

18.
Eyring equations for the coefficient of the water catalysisk w and for the coefficient of the catalysis of a series of anions with the same activation enthalpy as waterk A ? give for the difference of the activation entropies $$\Delta S_{A^ - }^* - \Delta S_w^* = \Delta S = R ln\frac{{k_{A^ - } }}{{k_w }}.$$ While at the general acid-base catalysis of the mutarotation of α-glucose ΔS is the entropy change in the reaction of the hydrated oxonium ion with the hydrated anion to the hydrated acid, ΔS for the decomposition of nitramide is the entropy change in the reaction of the hydrated oxonium ion with the hydrated anion to the unhydrated acid. The mechanism of reaction leads over the proton transfer from the tautomer of nitramide to the anion by the water molecule of the hydrated anion forming the unhydrated acid. As the conjugated acid for the hydroxide ion catalysis of the mutarotation of α-glucose and for the decomposition of nitramide is water, the entropy change of the breaking of the hydrogen bonds of water H2O...H?O?H...OH2 can be calculated. This entropy value agrees with that found byWalrafen from the Raman-spectrum of water. “Brönsted catalysis law” is replaced-according to the author — by the linear relation of ΔS to the logarithm of the base constantK B , or to the change of the free enthalpy of the acid ionization, resp. For the general base catalysis of the α-glucose mutarotation $$\Delta S = \beta R \ln K_B ,$$ for the general base catalysis of the decomposition of nitramide, for which no general acid catalysis exists, $$\begin{gathered} \Delta S = R \ln K_B , \hfill \\ therefore \beta = 1. \hfill \\\end{gathered}$$ The deviations from this relation were cleared up by thermodynamic calculations. The coefficient of the general base catalysis of the mutarotation of α-glucose β is due to the hydrogen bond of water.  相似文献   

19.
Two general relation between bond orderl and bond distance d (Å) are proposed:
  1. between atomssp 2-hybridised of second and third row: $$d_{PQ} = \left[ {0,731 + 0,3181\left( {n_P + n_Q } \right) - 0,1477\left( {\zeta _P + \zeta _Q } \right)} \right] - 0,020 + 0,0523\left( {\zeta _P + \zeta _Q } \right)l_{PQ} $$ ,ζ=Z/n,Z=Slater's effective nuclear charge of theπ-orbital).
  2. between atomssp-hybridised of the second row: $$d_{PQ} = \left[ {1,904 - 0,123\left( {\zeta _P + \zeta _Q } \right)} \right] - \left[ {0,075 + 0,023\left( {\zeta _P + \zeta _Q } \right)} \right]l_{PQ} $$ (l=total bond orderπ+π′).
  相似文献   

20.
A method for preparing chlorotungstic acid $$(H_3 O)_2 [W_6 Cl_8 i]Cl_6 a \cdot 6H_2 O$$ in good yield is given. On thermal degradation of the acid, the stages $$(H_2 O)_2 [W_6 Cl_8 ]Cl_6 ,[W_6 Cl_8 ]Cl_4 \cdot 2H_2 O and [W_6 Cl_8 ]Cl_4 $$ are isolable. Chlorotungstic acid and its partial Br i -substitution products can be precipitated almost quantitatively as $$(Oxin \cdot H)_2 [W_6 X_8 ]X_6 $$ When boiled with strong aqueous or aqueous-ethanolic HBr the substitution of Cl a and also partial Cl i /Br substitution occurs. In the same way I i can be introduced. The inverse reaction (substitution of Br i by Cl) is not possible. In ethanolic HB in the case of Cl i /Br substitution an induction period is observed.  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