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.
The Wheeler–Jonas equation (WJ equation) is widely used to predict breakthrough of volatile organic compounds on granular activated carbon (GAC) and the most important criteria is to calculate the overall adsorption rate constant (k v ) and amount adsorbed based on breakthrough curve. The operational factors, including temperatures, concentrations and flow rates, for packed bed, can affect the values of k v more or less and to what degree has not yet been systematically investigated. What’s more, the relation between the dynamic adsorption coefficient (k d) or Henry constant, which is used for the design of packed bed, and WJ equation is not clear. In order to solve these problems, we performed xenon dynamic adsorption on GAC adsorber experiments under different values of xenon concentrations, flow rates and temperatures, obtained the breakthrough curves for elution times versus xenon concentrations, and then employed WJ equation to explain breakthrough curves. The experimental results indicate that the WJ equation can fit the breakthrough curve very well and k d be integrated into the WJ equation. The values of k v are proportional to the values of flow rates and k d, but independent of that of temperatures and xenon concentrations.  相似文献   

2.
Thermal decomposition kinetics of magnesite were investigated using non-isothermal TG-DSC technique at heating rate (β) of 15, 20, 25, 35, and 40 K min−1. The method combined Friedman equation and Kissinger equation was applied to calculate the E and lgA values. A new multiple rate iso-temperature method was used to determine the magnesite thermal decomposition mechanism function, based on the assumption of a series of mechanism functions. The mechanism corresponding to this value of F(a), which with high correlation coefficient (r-squared value) of linear regression analysis and the slope was equal to −1.000, was selected. And the Malek method was also used to further study the magnesite decomposition kinetics. The research results showed that the decomposition of magnesite was controlled by three-dimension diffusion; mechanism function was the anti-Jander equation, the apparent activation energy (E), and the pre-exponential term (A) were 156.12 kJ mol−1 and 105.61 s−1, respectively. The kinetic equation was
\frac\textda\textdT = \frac105. 6 1 bexp( - \frac18777.9T ){ \frac32(1 + a)2/3 [(1 + a)1/3 - 1] - 1 }, \frac{{{\text{d}}\alpha }}{{{\text{d}}T}} = \frac{{10^{5. 6 1} }}{\beta }\exp \left( { - \frac{18777.9}{T}} \right)\left\{ {\frac{3}{2}(1 + \alpha )^{2/3} [(1 + \alpha )^{1/3} - 1]^{ - 1} } \right\},  相似文献   

3.
A new approach for determining the activation energy of amorphous alloys is developed. Setting the second order differential coefficient of heterogeneous reaction rate equation of non-isothermal heating as zero at extreme points of DSC curve, we obtain the new correlation taking form:
g1 = Lambertw( g3 e - g2 ) + g2 \gamma_{1} = Lambertw\left( {\gamma_{3} e^{{ - \gamma_{2} }} } \right) + \gamma_{2}  相似文献   

4.
The curing kinetics of a bi-component system about o-cresol-formaldehyde epoxy resin (o-CFER) modified by liquid crystalline p-phenylene di[4-(2,3-epoxypropyl) benzoate] (p-PEPB), with 3-methyl-tetrahydrophthalic anhydride (MeTHPA) as a curing agent, were studied by non-isothermal differential scanning calorimetry (DSC) method. The relationship between apparent activation energy E a and the conversion α was obtained by the isoconversional method of Ozawa. The reaction molecular mechanism was proposed. The results show that the values of E a in the initial stage are higher than other time, and E a tend to decrease slightly with the reaction processing. There is a phase separation in the cure process with LC phase formation. These curing reactions can be described by the Šesták–Berggren (S–B) equation, the kinetic equation of cure reaction as follows: \frac\textda\textdt = Aexp( - \fracE\texta RT )am ( 1 - a )n {\frac{{{\text{d}}\alpha }}{{{\text{d}}t}}} = A\exp \left( { - {\frac{{E_{\text{a}} }}{RT}}} \right)\alpha^{m} \left( {1 - a} \right)^{n} .  相似文献   

5.
The solution of the exponential integral at linear heating for the general case that the activation energy linearly depends on temperature according toE(T)=E 0+RBT is
\fracAqò0T TB exp( - \fracE0 RT ) dT = \fracAq( \fracRTB + 2 E0 + (B + 2)RT ) exp( - \fracE0 RT ).\frac{A}{q}\int\limits_0^T {T^B \exp \left( { - \frac{{E_0 }}{{RT}}} \right) dT = \frac{A}{q}\left( {\frac{{RT^{B + 2} }}{{E_0 + (B + 2)RT}}} \right)} \exp \left( { - \frac{{E_0 }}{{RT}}} \right).  相似文献   

6.
The differential capacitance curves were measured with an ac bridge in the Ga/[N-MF + 0.1 m M KBr + 0.1 (1 − m) M KClO4] and Ga/[N-MF + 0.1 m M KI + 0.1 (1 − m) M KClO4] systems at the following fractions m of surface-active anions: 0, 0.01, 0.02, 0.05, 0.1, 0.2, 0.5, and 1. As compared with other solvents, N-methyl formamide (N-MF) enables one to realize the largest positive charges of Ga electrode, at which it remains ideally polarizable (up to 20 μ/cm2). The data on the specific adsorption of Br and I anions in the system can be quantitatively described by the Frumkin’s isotherm; to the first approximation, free energy of halide ion (Hal) adsorption DGadsHal - 1 \Delta G_{adsHal^{ - 1} } is a linear function of electrode charge. It is found that, in contrast to the Hg/N-MF interface, DGadsHal - 1 \Delta G_{adsHal^{ - 1} } at the Ga/N-MF interface varies in the reverse order: Brt— ∼ I < Cl. From the measured results, we can conclude that the energy of metal-Hal interaction increases in series: $\Delta G_{M - Cl^ - } > \Delta G_{M - Br^ - } > \Delta G_{M - I^ - } $\Delta G_{M - Cl^ - } > \Delta G_{M - Br^ - } > \Delta G_{M - I^ - } and the difference (DGGa - Hal1- - DGGa - Hal2- )(\Delta G_{Ga - Hal_1^ - } - \Delta G_{Ga - Hal_2^ - } ) is larger than the difference between the solvation energies of Hal- (DGS - Hal1- - DGS - Hal2- )Hal^ - (\Delta G_{S - Hal_1^ - } - \Delta G_{S - Hal_2^ - } ).  相似文献   

7.
The enthalpies of dissolution of 1,2,3-triazole nitrate in water were measured using a RD496-2000 Calvet microcalorimeter at four different temperatures under atmospheric pressure. Differential enthalpies (Δdif H) and molar enthalpies (Δdiss H) of dissolution were determined. The corresponding kinetic equations that describe the dissolution rate at the four experimental temperatures are \fracdadt / s - 1 = 10 - 3.75( 1 - a)0.96\frac{d\alpha}{dt} / \mathrm{s}^{ - 1} =10^{ - 3.75}( 1 - \alpha)^{0.96} (T=298.15 K), \fracdadt /s - 1 = 10 - 3.73( 1 - a)1.00\frac{d\alpha}{dt} /\mathrm{s}^{ - 1} = 10^{ - 3.73}( 1 - \alpha)^{1.00} (T=303.15 K), \fracdadt / s - 1 = 10 - 3.72( 1 - a)0.98\frac{d\alpha}{dt} / \mathrm{s}^{ - 1} = 10^{ - 3.72}( 1 - \alpha)^{0.98} (T=308.15 K) and \fracdadt / s - 1 = 10 - 3.71( 1 -a)0.97\frac{d\alpha}{dt} / \mathrm{s}^{ - 1} = 10^{ - 3.71}( 1 -\alpha)^{0.97} (T=313.15 K). The determined values of the activation energy E and pre-exponential factor A for the dissolution process are 5.01 kJ⋅mol−1 and 10−2.87 s−1, respectively.  相似文献   

8.
The formation of large even-numbered carbon cluster anions, \textC\textn - {\text{C}}_{\text{n}}^{ - } , with n up to 500 were observed in the mass spectra generated by laser ablation of graphene and graphene oxide, and the signal intensity of the latter is much weaker than that of the former. The cluster distributions generated from graphene can be readily altered by changing the laser energy and the accumulation period in the FT - ICR cell. By choosing suitable experimental conditions, weak signals of odd-numbered anions from \textC125 - {\text{C}}_{{125}}^{ - } to \textC211 - {\text{C}}_{{211}}^{ - } , doubly charged anions from \textC702 - {\text{C}}_{{70}}^{{2 - }} to \textC2302 - {\text{C}}_{{230}}^{{2 - }} and triply charged cluster anions from \textC803 - {\text{C}}_{{80}}^{{3 - }} to \textC2243 - {\text{C}}_{{224}}^{{3 - }} can be observed. Tandem MS was applied to some selected cluster anions. Though no fragment anions larger than \textC20 - {\text{C}}_{{20}}^{ - } can be observed by the process of collisional activation with N2 gas for most cluster ions, several cluster anions can lose units of C2, C4, C6 or C8 in their collision process. The differences in their dissociation kinetics and structures require further calculations and experimental studies.  相似文献   

9.
The decomposition of nitromethane behind shock waves was studied at T = 1190–1490 K and P ≈ 1.5 atm; the reaction was monitored based on the formation and consumption of [(N)\dot]O2\dot NO_2, radicals, which were detected by their absorption in the region of λ = 405 nm. It was found that the curves of the yield of [(N)\dot]O2\dot NO_2 have a convex shape, which is characteristic of the formation of primary decomposition products. Based on an analysis of the initial sections of the experimental curves of the yield of [(N)\dot]O2\dot NO_2, the temperature dependence of the rate constants of formation of these radicals upon the decomposition of nitromethane was found for the first time: k1 ([(N)\dot]O2 ) = (6.3 ±2) ×1012 exp( - 48.9 ±2/RT)s - 1k_1 (\dot NO_2 ) = (6.3 \pm 2) \times 10^{12} \exp ( - 48.9 \pm 2/RT)s^{ - 1} (the dimensionality of E a is kcal/mol). It was found that the rate constants of nitromethane decomposition measured from the consumption of the parent substance and the yield of [(N)\dot]O2\dot NO_2 radicals almost coincide with each other. A kinetic simulation of the formation and consumption of [(N)\dot]O2\dot NO_2 upon the decomposition of nitromethane was performed. A good agreement between experimental and calculated data was achieved. A brief theoretical analysis of competition between the channels of direct disintegration and isomerization under conditions of the thermal decomposition of nitromethane was performed. The advantage of the direct disintegration channel with the rupture of the C-N bond was shown. Both experimental and published data on the isomerization of nitromethane into methyl nitrite upon its thermal decomposition and photolysis were analyzed.  相似文献   

10.
The 17O-NMR spin-lattice relaxation times (T 1) of water molecules in aqueous solutions of n-alkylsulfonate (C1 to C6) and arylsulfonic anions were determined as a function of concentration at 298 K. Values of the dynamic hydration number, (S-) = nh - (tc- /tc0 - 1)(\mathrm{S}^{-}) = n_{\mathrm{h}}^{ -} (\tau_{\mathrm{c}}^{-} /\tau_{\mathrm{c}}^{0} - 1), were determined from the concentration dependence of T 1. The ratios (tc -/tc0\tau_{\mathrm{c}}^{ -}/\tau_{\mathrm{c}}^{0}) of the rotational correlation times (tc -\tau_{\mathrm{c}}^{ -} ) of the water molecules around each sulfonate anion in the aqueous solutions to the rotational correlation time of pure water (tc0\tau_{\mathrm{c}}^{0}) were obtained from the n DHN(S) and the hydration number (nh -n_{\mathrm{h}}^{ -} ) results, which was calculated from the water accessible surface area (ASA) of the solute molecule. The tc -/tc0\tau_{\mathrm{c}}^{ -}/\tau_{\mathrm{c}}^{0} values for alkylsulfonate anions increase with increasing ASA in the homologous-series range of C1 to C4, but then become approximately constant. This result shows that the water structures of hydrophobic hydration near large size alkyl groups are less ordered. The rotational motions of water molecules around an aromatic group are faster than those around an n-alkyl group with the same ASA. That is, the number of water–water hydrogen bonds in the hydration water of aromatic groups is smaller in comparison with the hydration water of an n-alkyl group having the same ASA. Hydrophobic hydration is strongly disturbed by a sulfonate group, which acts as a water structure breaker. The disturbance effect decreases in the following order: $\mbox{--} \mathrm{SO}_{3}^{-} > \mbox{--} \mathrm{NH}_{3}^{ +} > \mathrm{OH}> \mathrm{NH}_{2}$\mbox{--} \mathrm{SO}_{3}^{-} > \mbox{--} \mathrm{NH}_{3}^{ +} > \mathrm{OH}> \mathrm{NH}_{2}. The partial molar volumes and viscosity B V coefficients for alkylsulfonate anions are linearly dependent on their n DHN(S) values.  相似文献   

11.
In order to evaluate the effectiveness of l-lactate dehydrogenase (LDH) from rabbit muscle as a regenerative catalyst of the biologically important cofactor nicotinamide adenine dinucleotide (NAD), the kinetics over broad concentrations were studied to develop a suitable kinetic rate expression. Despite robust literature describing the intricate complexations, the mammalian rabbit muscle LDH lacks a quantitative kinetic rate expression accounting for simultaneous inhibition parameters, specifically at high pyruvate concentrations. Product inhibition by l-lactate was observed to reduce activity at concentrations greater than 25 mM, while expected substrate inhibition by pyruvate was significant above 4.3 mM concentration. The combined effect of ternary and binary complexes of pyruvate and the coenzymes led to experimental rates as little as a third of expected activity. The convenience of the statistical software package JMP allowed for effective determination of experimental kinetic constants and simplification to a suitable rate expression:
v = \fracVmax( AB )KiaKb + KbA + KaB + AB + \fracPKI - Lac + \fracB2AKI - Pyr + \fracB2QKI - Pyr - NAD v = \frac{{{V_{max}}\left( {AB} \right)}}{{{K_{ia}}{K_b} + {K_b}A + {K_a}B + AB + \frac{P}{{{K_{I - Lac}}}} + \frac{{{B^2}A}}{{{K_{I - Pyr}}}} + \frac{{{B^2}Q}}{{{K_{I - Pyr - NAD}}}}}}  相似文献   

12.
The kinetics of oxidation of L-valine by a copper(III) periodate complex was studied spectrophotometrically. The inverse second-order dependency on [OH] was due to the formation of the protonated diperiodatocuprate(III) complex ([Cu(H3IO6)2]) from [Cu(H2IO6)2]3−. The retarding effect of initially added periodate suggests that the dissociation of copper(III) periodate complex occurs in a pre-equilibrium step in which it loses one periodate ligand. Among the various forms of copper(III) periodate complex occurring in alkaline solutions, the monoperiodatocuprate(III) appears to be the active form of copper(III) periodate complex. The observed second-order dependency of [L-valine] on the rate of reaction appears to result from formation of a complex with monoperiodatocuprate(III) followed by oxidation in a slow step. A suitable mechanism consistent with experimental results was proposed. The rate law was derived as:
- \fracd[DPC]dt = \frackK1K2K3[Cu(H2IO6)2]f3- [L -Val]f2[H3IO62 -]f[OH - ]f2.- \frac{\mathrm{d}[\mathrm{DPC}]}{\mathrm{d}t} =\frac{kK_{1}K_{2}K_{3}[\mathrm{Cu}(\mathrm{H}_{2}\mathrm{IO}_{6})_{2}]_{\mathrm{f}}^{3-} [\mathrm{L} -\mathrm{Val}]_{\mathrm{f}}^{2}}{[\mathrm{H}_{3}\mathrm{IO}_{6}^{2 -}]_{\mathrm{f}}[\mathrm{OH}^{ -} ]_{\mathrm{f}}^{2}}.  相似文献   

13.
The molar conductivities (Λ) of solutions of bis(2,2′-bipyridine)bis(thiocyanate)chromium(III) triiodide [CrIII(bipy)2(SCN)2]I3 (where bipy denotes 2,2′-bipyridine, C10H8N2), [ _3^-\mathrm{A}^{+}\mathrm{I}_{3}^{-} ], were measured in acetonitrile (ACN) at the temperatures 294.15, 299.15, and 305.15 K. In addition, cyclic voltammograms (CVs) of [ A+I3-\mathrm{A}^{+}\mathrm{I}_{3}^{-} ] were recorded on platinum, gold, and glassy carbon working electrodes in ACN, using n-tetrabutylammonium hexafluorophosphate (NBu4PF6) as the supporting electrolyte, at scan rates (v) ranging from 0.05 to 0.12 V⋅s−1. Furthermore, electrochemical impedance spectroscopic (EIS) measurements were carried out in the frequency range 50 Hz<f<50 kHz using these three working electrodes. The measured molar conductivities (Λ) demonstrate that [ A+I3-\mathrm{A}^{+}\mathrm{I}_{3}^{-} ] behaves as uni-univalent electrolyte in ACN over the investigated temperature range. The Λ values were analyzed by means of the Lee-Wheaton conductivity equation in order to estimate the limiting molar conductivities (Λ o), as well as the thermodynamic association constants (K A), at each experimental temperature for formation of [A+ I3-\mathrm{I}_{3}^{-} ] ion-pairs. The limiting ionic conductivities ( l±o\lambda_{\pm}^{\mathrm{o}} ), the diffusion coefficients at infinite dilution (D ±), as well as the Stokes’ radii (r St) were determined for both A+ and I3-\mathrm{I}_{3}^{-} ions. The thermodynamic parameters for the ionic association process, i.e. the Gibbs energy ( DGAo\Delta G_{\mathrm{A}}^{\mathrm{o}} ), enthalpy ( DHAo\Delta H_{\mathrm{A}}^{\mathrm{o}} ), and entropy ( DSAo\Delta S_{\mathrm{A}}^{\mathrm{o}} ), were also determined. The mobility and diffusivity of the A+ ion increase linearly with increasing temperature because the solvent medium becomes less viscous as the temperature increases. The K A values indicate that significant ion association occurs that is not influenced by temperature changes. The ion-pair formation process is exothermic ( DHAo < 0\Delta H_{\mathrm{A}}^{\mathrm{o}}<0 ), leading to the generation of additional entropy ( $\Delta S_{\mathrm{A}}^{\mathrm{o}}>0$\Delta S_{\mathrm{A}}^{\mathrm{o}}>0 ). As a result, the Gibbs energy DGAo\Delta G_{\mathrm{A}}^{\mathrm{o}} is negative ( DGAo < 0\Delta G_{\mathrm{A}}^{\mathrm{o}}<0 ) and the formation of [A+I3-][\mathrm{A}^{+}\mathrm{I}_{3}^{-}] becomes favorable. CV studies on [A+I3-][\mathrm{A}^{+}\mathrm{I}_{3}^{-}] solutions indicated that the redox pair Cr3+/2+ appears to be quasi-reversible on a glassy carbon electrode but is completely irreversible on platinum and gold electrodes. EIS experiments confirm that, among these three electrodes, the glassy carbon working electrode has the smallest resistance to electron transfer.  相似文献   

14.
The enthalpies of dissolution in ethyl acetate and acetone of hexanitrohexaazaisowurtzitane (CL-20) were measured by means of a RD496-2000 Calvet microcalorimeter at 298.15 K, respectively. Empirical formulae for the calculation of the enthalpy of dissolution (Δdiss H), relative partial molar enthalpy (Δdiss H partial), relative apparent molar enthalpy (Δdiss H apparent), and the enthalpy of dilution (Δdil H 1,2) of each process were obtained from the experimental data of the enthalpy of dissolution of CL-20. The corresponding kinetic equations describing the two dissolution processes were \frac\textda\textdt = 1.60 ×10 - 2 (1 - a)0.84 {\frac{{{\text{d}}\alpha }}{{{\text{d}}t}}} = 1.60 \times 10^{ - 2} (1 - \alpha )^{0.84} for dissolution process of CL-20 in ethyl acetate, and \frac\textda\textdt = 2.15 ×10 - 2 (1 - a)0.89 {\frac{{{\text{d}}\alpha }}{{{\text{d}}t}}} = 2.15 \times 10^{ - 2} (1 - \alpha )^{0.89} for dissolution process of CL-20 in acetone.  相似文献   

15.
Park  M.K.  Ryu  S.G.  Park  H.B.  Lee  H.W.  Hwang  K.C.  Lee  C.H. 《Plasma Chemistry and Plasma Processing》2004,24(1):117-136
The decomposition rate of CNCl in a BaTiO3-filled Packed Bed Plasma Reactor was studied as a function of AC input power, power frequency, residence time in the reactor, and inlet flow rate. The decomposition rate was compared with those of CH3CN and CCl2CHCl. Under the condition of 6.7 Wh/m3 specific energy den- sity, the decomposition rate of CNCl was found to be 50%, which is lower than those of CH3CN and CCl2CHCl at the same or similar conditions. At a higher frequency of the power input system, the decomposition rate of inlet gas becomes lower due to a decrease in field strength for the same level of power. And, under the same level of input power, a higher decomposition rate was obtained at an increased residence time. The relation between gas decomposition rates stemmed from the electron–molecule collision and bounding energy within the molecule. The decomposition ratio of CNCl was lower than those of CCl2CHCl and CH3CN because the bond strength of the weakest bond in the molecule is higher. In order to test the decomposition efficiency of CNCl with catalytic packing material in a plasma reactor, the catalyst of γ-Al2O3 and Pt/γ-Al2O3 was packed in the packed bed plasma reactor. Although byproducts were formed, the plasma-catalyst hybrid reactor containing Pt/γ-Al2O3 showed a higher efficiency in CNCl decomposition as shown in the decomposition rate of above 99% in 0.3 kWh/m3.  相似文献   

16.
The mer-[Ru(pic)3] isomer, where pic is 2-pyridinecarboxylic acid, undergoes base hydrolysis at pH > 12. The reaction was monitored spectrophotometrically within the UV–Vis spectral range. The product of the reaction, the [Ru(pic)2(OH)2] ion, is formed via a consecutive two-stage process. The chelate ring opening is proceeded by the nucleophilic attack of OH ion at the carbon atom of the carboxylic group and the deprotonation of the attached hydroxo group. In the second stage, the fast deprotonation of the coordinated OH ligand leads to liberation of the monodentato bonded picolinate. The dependence of the observed pseudo-first-order rate constant on [OH] is given by k\textobs1 = \frack + k1 [\textOH - ] + k + k2 K1 [\textOH - ]2 k - + k1 + ( k + + k2 K1 )[\textOH - ] + k + K1 [\textOH - ]2 k_{{{\text{obs}}1}} = \frac{{k_{ + } k_{1} [{\text{OH}}^{ - } ] + k_{ + } k_{2} K_{1} [{\text{OH}}^{ - } ]^{2} }}{{k_{ - } + k_{1} + \left( {k_{ + } + k_{2} K_{1} } \right)[{\text{OH}}^{ - } ] + k{}_{ + }K_{1} [{\text{OH}}^{ - } ]^{2} }} and ( k\textobs2 = \frackca + kcb K2 [\textOH - ]1 + K2 [\textOH - ] ) \left( {k_{{{\text{obs}}2}} = \frac{{k_{ca} + k_{cb} K_{2} [{\text{OH}}^{ - } ]}}{{1 + K_{2} [{\text{OH}}^{ - } ]}}} \right) for the first and the second stage, respectively, where k 1, k 2, k -, k ca , k cb are the first-order rate constants and k + is the second-order one, K 1 and K 2 are the protolytic equilibria constants.  相似文献   

17.
Wen-Rui Jin  Kun Liu 《中国化学》1985,3(4):321-331
The behaviour of the cobalt complex with dimethylglyoxime (DMG), Co(II)A2, at the mercury electrode has been investigated in details. The adsorption phenomena have been observed by both normal pulse polarography and voltammetry with linearly changing potential. Experimental results show that, under the condition of adsorption potentials ranging from ?0.60 to ?0.9 eV (vs. S.C.E.), Co(II)A2 can be adsorbed on the surface of hanging mercury drop electrode (HMDE) very well. The superficial concentrations represents a Langmuir isotherm with both concentration of Co(II)A2 and the preconcentration time. The superficial concentration equation for adsorption voltammetry, corresponding to the condition of the low coverage of the electrode surface, is deduced. The equation has been verified experimentally. The sensitivity of the proposed method, which has been analysed theoretically, is independent on the scan rate and the surface area of HMDE, but depends on the preconcentration time and the diffusion layer thickness. For the 120 sec accumulation, the lower limit of determination is 1.10?9 M.  相似文献   

18.
Apparent molar heat capacities C\textp,fC_{{\text{p}},\phi } and volumes VfV_\phi of aqueous trifluoromethanesulfonic (triflic) acid, HCF3SO3 (aq.) were determined with a Picker flow microcalorimeter and vibrating-tube densimeter at temperatures from 283 to 328 K and molalities from 0.05 to 9.5 mol-kg-1. Values of VfV_\phi and C\textp,fC_{{\text{p}},\phi } display a maximum near 0.8 mol-kg-1. VfV_\phi also displays a shallow minimum at ~5 mol-kg-1, while C\textp,fC_{{\text{p}},\phi } continues to decrease smoothly up to the limit of our measurements at 9.5 mol-kg-1. We attribute this behavior to ion–ion interaction between triflate and the hydrated proton to form the aqueous complex H2n+1On+ CF3SO3- (aq.), n = 5. Standard partial molar properties Cpo and Vo are consistent with results obtained from NaCF3SO3 (aq.) and yield values for the triflate anion CF3SO3- (aq.), over this range.  相似文献   

19.
Study on Adsorption and Separation of Naringin with Macroporous Resin   总被引:1,自引:0,他引:1  
X-5 resin, with higher adsorption and easier desorption of naringin, was selected from five kinds of macroporous resins through static adsorption and desorption experiments. Effects of concentration, pH value, and flow rate of naringin extract on the adsorption of naringin by X-5 resin were studied. Meanwhile, the effect of these factors on the desorption of naringin from X-5 resin was also investigated. The experimental results show that the adsorption isotherm of naringin by X-5 resin can be described by the Langmuir isotherm equation. The static maximum adsorption capacity of naringin is 32.6 mg/g with naringin concentration at 2.7 g/L, while the dynamic adsorption capacity of naringin is 23.8 mg/g with naringin extract flow rate at two times that of resin volume per hour. The optimal eluant is 60% (v/v) ethanol-water with pH value of 10. The desorption ratio will rise to more than 85% when the flow rate of this optimal eluant is one to two times that of resin volume per hour. Translated from Journal of Central South University (Science and Technology)  相似文献   

20.
The reaction mechanism of CH3SCH2CH3 with OH radical is studied at the CCSD(T)/6-311+G(3df,p)//MP2/6-31+G(2d,p) level of theory. Three hydrogen abstraction channels, one substitution process and five addition–elimination channels are identified in the title reaction. The result shows hydrogen abstraction is dominant. Substitution process and addition–elimination reactions may be negligible because of the high barrier heights. Enthalpies of formation [ \Updeltaf H(298.15\textK)o \Updelta_{f} H_{(298.15{\text{K}})}^{o} ] of the reactants and products are evaluated at the CBS-QB3, G3 and G3MP2 levels of theory, respectively. It is found that the calculated enthalpies of formation by the aforementioned three methods are in consistent with the available experimental data. Rate constants and branching ratios are estimated by means of the conventional transition state theory with the Wigner tunneling correction over the temperature range of 200–900 K. The calculation shows that the formations of P1 (CH2SCH2CH3 + H2O) and P2 (CH3SCHCH3 + H2O) are major products during 200–900 K. The three-parameter expressions for the total rate constant is fitted to be k\texttotal = 1.45 ×10 - 21 T3.24 exp( - 1384.54/T) k_{\text{total}} = 1.45 \times 10^{ - 21} T^{3.24} \exp ( - 1384.54/T) cm3 molecule−1 s−1 from 200 to 900 K.  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