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.
Electrical conductance data at 25°C for Li2SO4, Rb2SO4, Cs2SO4, and (NH4)2SO4 aqueous solutions are reported at concentrations up to 0.01 eq.-liter?1 and as a function of pressure up to 2000 atm. The molal dissociation constants are as follows: $$\begin{gathered} LiSO_4^ - : - log K_m = - 1.02 + 1.03 \times 10^4 P \pm 0.019 \Delta \bar V^o = - 5.8 \hfill \\ RbSO_4^ - : - log K_m = - 1.12 + 0.58 \times 10^4 P \pm 0.020 \Delta \bar V^o = - 3.3 \hfill \\ CsSO_4^ - : - log K_m = - 1.08 + 1.10 \times 10^4 P \pm 0.014 \Delta \bar V^o = - 6.2 \hfill \\ \left( {NH4} \right)SO_4^ - : - log K_m = - 1.12 + 0.58 \times 10^4 P \pm 0.020 \Delta \bar V^o = - 3.3 \hfill \\ \end{gathered} $$ whereP is in atmospheres and \(\Delta \bar V^o \) is in cm3-mole?1. These values were obtained by using the Davies-Otter-Prue conductance equation and Bjerrum distance parameters. A simultaneous Λ°,K m search was used to determine the equilibrium constantK m, a different procedure than used earlier for KSO 4 ? , NaSO 4 ? , and MgCl+. Recalculated values for these salts are as follows: $$\begin{gathered} KSO_4^ - : - log K_m = - 1.03 + 1.04 \times 10^4 P \pm 0.020 \Delta \bar V^o = - 5.9 \hfill \\ NaSO_4^ - : - log K_m = - 1.00 + 1.30 \times 10^4 P \pm 0.019 \Delta \bar V^o = - 7.3 \hfill \\ MgCl^ + : - log K_m = - 0.75 + 0.71 \times 10^4 P \pm 0.028 \Delta \bar V^o = - 4.0 \hfill \\ \end{gathered} $$   相似文献   

2.
The complexation kinetics of Mg2+ with CO 3 = and HCO 3 ? has been studied in methanol and water by means of the stopped-flow and temperature-jump methods. Kinetic parameters were obtained in methanol by coupling the magnesium-carbonato reactions with the metal-ion indicator Murexide. Relatively high stability constants were found in methanol (K=1.0×105 liters-mole?1 for Mg2+-Murexide,K=7.0×104 liters-mole?1 for Mg2+?HCO 3 ? , andK=2.0×105 for Mg2+?CO 3 = liters-mole?1). The corresponding, observed formation rate constants were determined to be $$\begin{gathered} k_f = 4.0 \times 10^6 M^{ - 1} - sec^{ - 1} (Mg^{2 + } - Murexide) \hfill \\ k_f = 5.0 \times 10^5 M^{ - 1} - sec^{ - 1} (Mg^{2 + } - HCO_3^ - ) \hfill \\ k_f = 6.8 \times 10^5 M^{ - 1} - sec^{ - 1} (Mg^{2 + } - CO_3^ = ) \hfill \\ \end{gathered} $$ The relaxation times were found to be much shorter (τ≈5–20 μsec) in aqueous solutions, primarily due to the relatively high dissociation rate constants. The data could be interpreted on the basis of a coupled reaction scheme in which the protolytic equilibria are established relatively rapidly, followed by a single relaxation process due to the formation of MgHCO 3 + and MgCO3 between pH 8.7 and 9.3. The observed formation rate constants were determined to be $$\begin{gathered} k_f = 5.0 \times 10^5 M^{ - 1} - sec^{ - 1} (Mg^{2 + } - HCO_3^ - ) \hfill \\ k_f = 1.5 \times 10^6 M^{ - 1} - sec^{ - 1} (Mg^{2 + } - CO_3^ = ) \hfill \\ \end{gathered} $$ These results, in conjunction with NMR solvent exchange rate constants, are analyzed in terms of a dissociative (S N1) mechanism for the rate of complex formation. The significance of these kinetic parameters in understanding the excess sound absorption in seawater is discussed.  相似文献   

3.
Cross-section values for 14.7 MeV neutrons have been measured for the following reactions: $$\begin{gathered} ^{24} Mg(n,p)^{24} Na - (187 \pm 7)mb;^{25} Mg(n,p)^{25} Na - (74 \pm 9)mb; \hfill \\ ^{26} Mg(n,\alpha )^{23} Ne - (55 \pm 6)mb;^{35} Cl(n,p)^{37} S - (22 \pm 3)mb; \hfill \\ ^{35} Cl(n,2n)^{34m} Cl - (9.3 \pm 1.5)mb;^{86} Sr(n,p)^{86m} Rb - (14 \pm 2)mb; \hfill \\ ^{88} Sr(n,p)^{88} Rb - (19 \pm 3)mb;^{86} Sr(n,2n)^{85m} Sr - (244 \pm 32)mb; \hfill \\ ^{88} Sr(n,2n)^{87m} Sr - (289 \pm 33)mb. \hfill \\ \end{gathered}$$ An analysis of available cross-section data for these reactions has been performed and preferred mean values for each reaction are given.  相似文献   

4.
Cross-section values for 14.7 MeV neutrons are measured for the following reactions: $$\begin{gathered} ^{31} P(n,\alpha )^{28} Al,(132 \pm 10)mb;^{42} Ca(n,p)^{42} K,(173 \pm 19)mb; \hfill \\ ^{43} Ca(n,p)^{43} K,(111 \pm 9)mb;^{44} Ca(n,p)^{44} K,(42 \pm 2)mb; \hfill \\ ^{44} Ca(n,\alpha )^{41} Ar,(27 \pm 2)mb;^{48} Ca(n,2n)^{47} Ca,(616 \pm 54)mb. \hfill \\ \end{gathered}$$ An analysis of available cross-section data for these reactions is performed and preferred mean values for each reaction are given.  相似文献   

5.
DTA, TG and DTG curves obtained in various atmospheres using different heating rates were used together with X-ray examinations to study the thermal decomposition mechanisms of two types of gelled UO3 microspheres: ammonia-washed (UN) and hot water-washed (UH) microspheres. The kinetics of the thermal decompositions were studied. The specific reaction rate constantk r for the decomposition of UO3 to U3O8 could be expressed in terms of the activation energy and the pre-exponential factor by the expressions: $$\begin{gathered} K_r (s^{ - 1} ) = 1.277 \times 10^{18} \exp \frac{{ - 295.4}}{{RT}}for the UN spheres, \hfill \\ K_r (s^{ - 1} ) = 8.406 \times 10^{19} \exp \frac{{ - 263.2}}{{RT}}for the UH spheres. \hfill \\ \end{gathered} $$   相似文献   

6.
Values of pa H o for 0.05 mole-kg?1 aqueous solutions of sodium hydrogen diglycolate in the temperature range 5–65°C have been obtained from cells without transport, and can be fitted to the equation $$\begin{gathered} pa^\circ _H = 3.5098 + 2.222 \times 10^{ - 3} ({T \mathord{\left/ {\vphantom {T {K - 298.15}}} \right. \kern-\nulldelimiterspace} {K - 298.15}}) \hfill \\ + 2.628 \times 10^{ - 5} ({T \mathord{\left/ {\vphantom {T {K - 298.15}}} \right. \kern-\nulldelimiterspace} {K - 298.15}})^2 \hfill \\ \end{gathered} $$ The analysis has been carried out by a multilinear regression procedure using a form of the Clarke and Glew equation. This buffer standard may be a useful alternative to the saturated potassium hydrogen tartrate buffer.  相似文献   

7.
The complex formation between Cu(II) and 8-hydroxyquinolinat (Ox) was studied with the liquid-liquid distribution method, between 1M-Na(ClO4) and CHCl3 at 25°C. The experimental data were explained by the equilibria: $$\begin{gathered} \operatorname{Cu} ^{2 + } + Ox \rightleftharpoons \operatorname{Cu} Ox \log \beta _1 = 12.38 \pm 0.13 \hfill \\ \operatorname{Cu} ^{2 + } + 2 Ox \rightleftharpoons \operatorname{Cu} Ox_2 \log \beta _2 = 23.80 \pm 0.10 \hfill \\ \operatorname{Cu} Ox_{2aq} \rightleftharpoons \operatorname{Cu} Ox_{2\operatorname{org} } \log \lambda = 2.06 \pm 0.08 \hfill \\ \end{gathered} $$ The equilibria between Cu(II) and o-aminophenolate (AF) were studied potentiometrically with a glass electrode at 25°C and in 1M-Na(ClO4). The experimental data were explained by the equilibria: $$\begin{gathered} \operatorname{Cu} ^{2 + } + AF \rightleftharpoons \operatorname{Cu} AF \log \beta _1 = 8.08 \pm 0.08 \hfill \\ \operatorname{Cu} ^{2 + } + 2AF \rightleftharpoons \operatorname{Cu} AF_2 \log \beta _2 = 14.60 \pm 0.06 \hfill \\ \end{gathered} $$ The protonation constants ofAF and the distribution constants between CHCl3?H2O and (C2H5)2O?H2O were also determined.  相似文献   

8.
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   

9.
The temperature dependencies of europium carbonate stability constants were examined at 15, 25, and 35°C in 0.68 molal Na+(ClO 4 ? , HCO 3 ? ) using a tributyl phosphate solvent extration technique. Our distribution data can be explained by the equilibria $$\begin{gathered} Eu^{3 + } + H_2 O + CO_2 (g)_ \leftarrow ^ \to EuCO_3^ + + 2H^ + \hfill \\ - log\beta _{12} = 9.607 + 496(t + 273.16)^{ - 1} \hfill \\ Eu^{3 + } + 2H_2 O + 2CO_2 (g)_ \leftarrow ^ \to Eu(CO_3 )_2^ - + 4H^ + \hfill \\ - log\beta _{24} = 21.951 + 670(t + 273.16)^{ - 1} \hfill \\ Eu^{3 + } + H_2 O + CO_2 (g)_ \leftarrow ^ \to EuHCO_3^{2 + } + H^ + \hfill \\ - log\beta _{11} = 1.688 + 1397(t + 273.16)^{ - 1} \hfill \\ \end{gathered}$$   相似文献   

10.
Electrical conductance data at 25°C for K2SO4, Na2SO4, and MgCl2 solutions are reported at concentrations up to 0.01 eq-liter?1 and as a function of pressure up to 2000 atm. The molal dissociation constants are as follows: $$\begin{gathered} {\text{ }}KSO_4^ - :log K_m = ( - 1.02{\text{ }} + 1.6 \times 10^{ - 4} P - {\text{ }}2.5 \times 10^{ - 8p2} ) \pm 0.03 \hfill \\ NaSO_4^ - :log K_m = ( - 1.02 + 9.6 \times 10^{ - 5} P - {\text{ }}4.3 \times 10^{ - 9p2} ) \pm 0.03 \hfill \\ MgCl^ + :log K_m = ( - 0.64 + 1.1 \times 10^{ - 4} P - {\text{ }}1.7 \times 10^{ - 8p2} ) \pm 0.04 \hfill \\ \end{gathered} $$ withP in atmospheres. These values cannot be chosen solely on the basis of minimizing errors in fitting conductance data to theoretical equations. For the values cited above, the Bjerrum distances for 1–2 (or 2-1) and 1-1 salts were used. However, the conductance fits for KSO 4 ? and NaSO 4 ? were equally good for half-Bjerrum distances and resulted in higher dissociation constants. Ultrasonic data are used to argue in favor of the lower dissociation values derived by using Bjerrum distances. Our results for MgCl+ disagree with those of Havel and Högfeldt.  相似文献   

11.
Isotopic abundance values for50Cr,58Fe and109Ag and the absolute gamma-intensities for51Cr,59Fe and110mAg were evaluated. These evaluated data, together with experimental k0-determinations (i.e. from the “activation method”), made it possible to calculate the following 2200 m.s?1 cross-sections, which considerably deviate from the hitherto generally published ones [between brackets]: $$\begin{gathered} {}^{5 0}Cr(n,\gamma )^{5 1} Cr; \sigma _0 = (15.2 \pm 0.2) barn [cf.:15.8 - 16.0] \hfill \\ {}^{5 8}Fe(n,\gamma )^{5 9} Fe; \sigma _0 = (1.31 \pm 0.03) barn [cf.:1.14 - 1.16] \hfill \\ {}^{1 0 9}Ag(n,\gamma )^{1 1 0 m} Ag;\sigma _0 = (3.89 \pm 0.05) barn [cf.:4.4 - 5.0] \hfill \\ \end{gathered} $$   相似文献   

12.
In the present work, the conductivity of Standard Sea Water was studied over the temperature range 12–40°C using high-precision conductance techniques. One of the purposes of this investigation was to extend the temperature range of the equation that describes the conductivities tabulated by Cox. The conductivities reported in this work differ from the Cox equation by no more than 0.1% in the range of temperature overlap (to 30°C). The combined data base leads to the expression $$\begin{gathered} \kappa \left( {mmho - cm^{ - 1} } \right) = 29.05128 + 0.88082t - 1.98312 \times 10^{ - 4} t^2 + 3.33663 \hfill \\ \times 10^{ - 4} t^3 - 1.0776 \times 10^{ - 5} t^4 + 1.12518 \times 10^{ - 7} t^5 \hfill \\ \end{gathered} $$ for the electrical conductance of Copenhagen Standard Sea Water, with an overall uncertainty estimate (absolute accuracy) of ±0.1% for the temperature range of 0–40°C.  相似文献   

13.
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   

14.
The pK 2 * for the dissociation of sulfurous acid from I=0.5 to 6.0 molal at 25°C has been determined from emf measurements in NaCl solutions with added concentrations of NiCl2, CoCl2, McCl2 and CdCl2 (m=0.1). These experimental results have been treated using both the ion pairing and Pitzer's specific ion-interaction models. The Pitzer parameters for the interaction of M2+ with SO 3 2? yielded $$\begin{gathered} \beta _{NiSO_3 }^{(0)} = - 5.5, \beta _{NiSO_3 }^{(1)} = 5.8, and \beta _{NiSO_3 }^{(2)} = - 138 \hfill \\ \beta _{CoSO_3 }^{(0)} = - 12.3, \beta _{CoSO_3 }^{(1)} = 31.6, and \beta _{CoSO_3 }^{(2)} = - 562 \hfill \\ \beta _{MnSO_3 }^{(0)} = - 8.9, \beta _{MnSO_3 }^{(1)} = 18.7, and \beta _{MnSO_3 }^{(2)} = - 353 \hfill \\ \beta _{CdSO_3 }^{(0)} = - 7.2, \beta _{CdSO_3 }^{(1)} = 13.8, and \beta _{CdSO_3 }^{(2)} = - 489 \hfill \\ \end{gathered} $$ The calculated values of pK 2 * using Pitzer's equations reproduce the measured values to within ±0.01 pK units. The ion pairing model yielded $$\begin{gathered} logK_{NiSO_3 } = 2.88 and log\gamma _{NiSO_3 } = 0.111 \hfill \\ logK_{CoSO_3 } = 3.08 and log\gamma _{CoSO_3 } = 0.051 \hfill \\ logK_{MnSO_3 } = 3.00 and log\gamma _{MnSO_3 } = 0.041 \hfill \\ logK_{CdSO_3 } = 3.29 and log\gamma _{CdSO_3 } = 0.171 \hfill \\ \end{gathered} $$ for the formation of the complex MSO3. The stability constants for the formation of MSO3 complexes were found to correlate with the literature values for the formation of MSO4 complexes.  相似文献   

15.
The kinetics of the 2,2′-azobisisobutyronitrile-initiated oxidation of methyl oleate (MO) in the medium of the oxidized substrate itself (homogeneous system) and in an aqueous solution of cetyltrimethylammonium bromide (aqueous emulsion system, AES) was studied. The oxidation was found to occur as a nonbranched chain reaction with quadratic-law chain termination in both neat methyl oleate ([MO] ≈ 2.6 mol/l) and AES. The temperature dependence of the oxidizability parameter for methyl oleate in the temperature range from 303 to 333 K was described by the following expressions:
$\begin{gathered} in an Mo medium, log(k_2 (2k_6 )^{ - 0.5} ) = (2.0 \pm 0.5) - (27.5 \pm 3.2)\theta ^{ - 1} [l^{0.5} mol^{ - 0.5} s^{ - 0.5} ]; \hfill \\ and in the AES, log(k_2 (2k_6 )^{ - 0.5} ) = (2.4 \pm 0.4) - (29.4 \pm 2.1)\theta ^{ - 1} [l^{0.5} mol^{ - 0.5} s^{ - 0.5} ], \hfill \\ \end{gathered}$\begin{gathered} in an Mo medium, log(k_2 (2k_6 )^{ - 0.5} ) = (2.0 \pm 0.5) - (27.5 \pm 3.2)\theta ^{ - 1} [l^{0.5} mol^{ - 0.5} s^{ - 0.5} ]; \hfill \\ and in the AES, log(k_2 (2k_6 )^{ - 0.5} ) = (2.4 \pm 0.4) - (29.4 \pm 2.1)\theta ^{ - 1} [l^{0.5} mol^{ - 0.5} s^{ - 0.5} ], \hfill \\ \end{gathered}  相似文献   

16.
The complex formation between copper(II) and acetylacetonate (L)* has been studied by potentiometry and distribution between CHCl3 and water. The experimental data are interpreted by postulating the following equilibria: $$\begin{gathered} Cu^{2 + } + L \rightleftharpoons CuL1g \beta _1 = 8.42 \pm 0.10 \hfill \\ Cu^{2 + } + 2 L \rightleftharpoons CuL_2 1g \beta _2 = 15.47 \pm 0.10 \hfill \\ \left( {CuL_2 } \right)_{aq} \rightleftharpoons \left( {CuL_2 } \right)_0 1g \lambda _B = 1.80 \pm 0.10 \hfill \\ \end{gathered} $$ In order to study the complex formation, the protonation constant (k) of acetylacetonate and the distribution coefficient λ A of acetylacetone in the same experimental conditions were required. It was found: lgk=9.05±0.03; λ A = 1.20 ± 0.02.  相似文献   

17.
The solvent extraction of Yb(III) and Ho(III) by 1-(2-pyridylazo)-2-naphthol (PAN or HL) in carbon tetrachloride from aqueous-methanol phase has been studied as a function ofpH × and the concentration ofPAN or methanol (MeOH) in the organic phase. When the aqueous phase contains above ~25%v/v of methanol the synergistic effect was increased. The equation for the extraction reaction has been suggested as: $$\begin{gathered} Ln(H_2 0)_{m(p)}^{3 + } + 3 HL_{(o)} + t MeOH_{(o)} \mathop \rightleftharpoons \limits^{K_{ex} } \hfill \\ LnL_3 (MeOH)_{t(o)} + 3 H_{(p)}^ + + m H_2 0 \hfill \\ \end{gathered} $$ where:Ln 3+=Yb, Ho; $$\begin{gathered} t = 3 for C_{MeOH in.} \varepsilon \left( { \sim 25 - 50} \right)\% {\upsilon \mathord{\left/ {\vphantom {\upsilon \upsilon }} \right. \kern-\nulldelimiterspace} \upsilon }; \hfill \\ t = 0 for C_{MeOH in.} \varepsilon \left( { \sim 5 - 25} \right)\% {\upsilon \mathord{\left/ {\vphantom {\upsilon \upsilon }} \right. \kern-\nulldelimiterspace} \upsilon } \hfill \\ \end{gathered} $$ . The extraction equilibrium constants (K ex ) and the two-phase stability constants (β 3 × ) for theLnL 3(MeOH)3 complexes have been evaluated.  相似文献   

18.
The crystal structure of Cr4As3 has been determined by single crystal photographs: $$\begin{gathered} space group Cm - C_s ^3 \hfill \\ \alpha = 13.16_8 {\AA} \hfill \\ b = 3.54_2 {\AA} \hfill \\ c = 9.30_2 {\AA} \hfill \\ \beta = 102.1_9 \circ \hfill \\ \end{gathered}$$ Cr4As3 crystallizes with a novel structure type, which can be derived from the MnP-structure type.  相似文献   

19.
Complex formation between lead(II) and ethylenedithio diacetic acid (H2 L) has been studied at 25°C in aqueous 0.5M sodium perchlorate medium. Measurements have been carried out with a glass electrode and with a lead amalgam electrode. In acidic medium and in the investigated concentration range experimental data can be explained by assuming the following equilibria: $$\begin{gathered} Pb^{2 + } + L^{2 - } \rightleftharpoons PbL log\beta _{101} = 3.62 \pm 0.03 \hfill \\ Pb^{2 + } + H^ + + L^{2 - } \rightleftharpoons PbHL^ - log\beta _{111} = 6.30 \pm 0.07 \hfill \\ \end{gathered} $$   相似文献   

20.
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