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 46 毫秒
1.
The solubility of H2S at 25°C in solvents of the composition: [H+]=H M, [Na+]=(I?H)=A M, [ClO4 ?]=I M was investigated by iodometric determination of [H2S]tot in the saturated solutions. Kp12=[H2S]tot·p H2S ?1 was calculated. The results are consistent with the equation:
$$\begin{gathered} \lg [H_2 S]_{tot} \cdot p_{H_2 S}^{ - 1} = --- 0,991_8 --- 0,059_0 [Na + ] + 0,008_1 [H + ]--- \hfill \\ ---0,000_1 [H + ]^4 . \hfill \\ \end{gathered} $$  相似文献   

2.
The solubilities in the quaternary system K+, \( {\text{NH}}{_4^{+}} \)//Cl?, \( {\text{SO}}{_4^{2-}} \)H2O and its two ternary subsystems NH4Cl–KCl–H2O, (NH4)2SO4–K2SO4–H2O at 80.0 °C were measured using the isothermal dissolution equilibrium method under atmospheric pressure, and the corresponding phase diagrams were plotted. In the phase diagram of the NH4Cl–KCl–H2O system, there are three crystalline zones, which correspond to (K1?m,(NH4)m)Cl, ((NH4)n,K1?n)Cl and the co-existence zone of (K1?m,(NH4)m)Cl and ((NH4)n,K1?n)Cl, respectively. In the phase diagram of the (NH4)2SO4–K2SO4–H2O system, there is only one crystalline zone for (K1?t,(NH4)t)2SO4. In the phase diagram of the K+, \( {\text{NH}}{_4^{+}} \)//Cl?, \( {\text{SO}}{_4^{2-}} \)H2O system, there are three crystal zones, which correspond to (K1?t,(NH4)t)2SO4, (K1?m,(NH4)m)Cl and ((NH4)n,K1?n)Cl, respectively. According to the analysis and the calculations for the phase diagrams of the K+, \( {\text{NH}}{_4^{+}} \)//Cl?, \( {\text{SO}}{_4^{2 -}} \)H2O system at 80.0 °C and 50.0 °C, this paper proposes a technological process. In the process, the (K1?t,(NH4)t)2SO4 can be prepared at 80.0 °C and the ((NH4)n,K1?n)Cl can crystallize out at 50.0 °C. The mass fraction of K2SO4 in product L1 (K1?t,(NH4)t)2SO4 (t?=?0.1465) is 88.48%. The composition of solid solutions in the K+, \( {\text{NH}}{_4^{+}} \)//Cl?, \( {\text{SO}}{_4^{2 -}} \)H2O system was experimentally determined and then theoretical calculations about the process can be carried out.  相似文献   

3.
B-Nb2O5 was recrystallized from commercially available oxide, and XRD analyses indicated that it is stable in contact with solutions over the pH range 0 to 9, whereas solid polyniobates such as Na8Nb6O19?13H2O(s) appear to predominate at pH>9. Solubilities of the crystalline B-Nb2O5 were determined in five NaClO4 solutions (0.1≤I m /mol?kg?1≤1.0) over a wide pH range at (25.0±0.1)?°C and at 0.1 MPa. A limited number of measurements were also made at I m =6.0 mol?kg?1, whereas at I m =1.0 mol?kg?1 the full range of pH was also covered at (10, 50 and 70)?°C. The pH of these solutions was fixed using either HClO4 (pH≤4) or NaOH (pH≥10) and determined by mass balance, whereas the pH on the molality scale was measured in buffer mixtures of acetic acid?+?acetate (4≤pH≤6), Bis-Tris (pH≈7), Tris (pH≈8) and boric acid?+?borate (pH≈9). Treatment of the solubility results indicated the presence of four species, \(\mathrm{Nb(OH)}_{n}^{5-n}\) (where n=4–7), so that the molal solubility quotients were determined according to:
$0\mathrm{.5Nb}_{2}\mathrm{O}_{5}\mathrm{(cr)+0}\mathrm{.5(2}n-5\mathrm{)H}_{2}\mathrm{O(l)}_{\leftarrow}^{\to}\mathrm{Nb(OH)}_{n}^{5-n}+(n-5)\mathrm{H}^{+}\quad (n=4\mbox{--}7)$
and were fitted empirically as a function of ionic strength and temperature, including the appropriate Debye-Hückel term. A Specific Interaction Theory (SIT) approach was also attempted. The former approach yielded the following values of log?10 K sn (infinite dilution) at 25?°C: ?(7.4±0.2) for n=4; ?(9.1±0.1) for n=5; ?(14.1±0.3) for n=6; and ?(23.9±0.6) for n=7. Given the experimental uncertainties (2σ), it is interesting to note that the effect of ionic strength only exceeded the combined uncertainties significantly in the case of log?10 K s6 to I m =1.0 mol?kg?1, such that these values may be of use by defining their magnitudes in other media. Values of Δ f G o, Δ f H o, S o and \(C_{p}^{\mathrm{o}}\) (298.15 K, 0.1 MPa) for each hydrolysis product were calculated and tabulated.
  相似文献   

4.
A potentiometric method has been used for the determination of the protonation constants of N-(2-hydroxyethyl)iminodiacetic acid (HEIDA or L) at various temperatures 283.15?≤?T/K?≤?383.15 and different ionic strengths of NaCl(aq), 0.12?≤?I/mol·kg?1?≤?4.84. Ionic strength dependence parameters were calculated using a Debye–Hückel type equation, Specific Ion Interaction Theory and Pitzer equations. Protonation constants at infinite dilution calculated by the SIT model are \( \log_{10} \left( {{}^{T}K_{1}^{\text{H}} } \right) = 8.998 \pm 0.008 \) (amino group), \( \log_{10} \left( {{}^{T}K_{2}^{\text{H}} } \right) = 2.515 \pm 0.009 \) and \( \log_{10} \left( {{}^{T}K_{3}^{\text{H}} } \right) = 1.06 \pm 0.002 \) (carboxylic groups). The formation constants of HEIDA complexes with sodium, calcium and magnesium were determined. In the first case, the formation of a weak complex species, NaL, was found and the stability constant value at infinite dilution is log10KNaL?=?0.78?±?0.23. For Ca2+ and Mg2+, the CaL, CaHL, CaL2 and MgL species were found, respectively. The calculated stability constants for the calcium complexes at T?=?298.15 K and I?=?0.150 mol·dm?3 are: log10βCaL?=?4.92?±?0.01, log10βCaHL?=?11.11?±?0.02 and \( \log_{10} \beta_{\text{Ca{L}}_{2}} \)?=?7.84?±?0.03, while for the magnesium complex (at I?=?0.176 mol·dm?3): log10βMgL?=?2.928?±?0.006. Protonation thermodynamic functions have also been calculated and interpreted.  相似文献   

5.
Some equilibria involving gold(I) thiomalate (mercaptosuccinate, TM) complexes have been studied in the aqueous solution at 25 °C and I?=?0.2 mol·L?1 (NaCl). In the acidic region, the oxidation of TM by \( {\text{AuCl}}_{4}^{ - } \) proceeds with the formation of sulfinic acid, and gold(III) is reduced to gold(I). The interaction of gold(I) with TM at nTM/nAu?≤?1 leads to the formation of highly stable cyclic polymeric complexes \( {\text{Au}}_{m} \left( {\text{TM}} \right)_{m}^{*} \) with various degrees of protonation depending on pH. In general, the results agree with the tetrameric form of this complex proposed in the literature. At nTM/nAu?>?1, the processes of opening the cyclic structure, depolymerization and the formation of \( {\text{Au}}\left( {\text{TM}} \right)_{2}^{*} \) occur: \( {\text{Au}}_{4} ( {\text{TM)}}_{4}^{8 - } + {\text{TM}}^{3 - } \rightleftharpoons {\text{Au}}_{ 4} ( {\text{TM)}}_{5}^{11 - } \), log10 K45?=?10.1?±?0.5; 0.25 \( {\text{Au}}_{4} ( {\text{TM)}}_{4}^{8 - } + {\text{TM}}^{3 - } \rightleftharpoons {\text{Au(TM)}}_{2}^{5 - } \), log10 K12?=?4.9?±?0.2. The standard potential of \( {\text{Au(TM)}}_{2}^{5 - } \) is \( E_{1/0}^{ \circ } = -0. 2 5 5\pm 0.0 30{\text{ V}} \). The numerous protonation processes of complexes at pH?<?7 were described with the use of effective functions.  相似文献   

6.
The vaporization of the NaI-PrI3 quasi-binary system was studied by high-temperature mass spectrometry over the whole concentration range. At 623–994 K, saturated vapor contained not only (NaI) n and (PrI3) n molecules (n = 1, 2) and Na+(NaI) n (n = 0–4) and I?(PrI3) n (n = 1–2) ions but also mixed molecular and ionic associates recorded for the first time (NaPrI4, Na2PrI5, NaPrI 3 + , Na2PrI 4 + , Na3PrI 5 + , Na4PrI 6 + , NaPrI 5 ? , and NaPr2I 8 ? ). The partial vapor pressures of molecules were calculated, and the equilibrium constants of the dissociation of neutral and charged associates were measured. The enthalpies of molecular and ion-molecular reactions were determined, and the enthalpies of formation of gaseous molecules and ions were obtained.  相似文献   

7.
8.
The solubility of α-MnS (Alabandite) in solutions of the general compositionL ([H+]=HM, [Na+]=(3,000-H)M, [ClO4 ?]=3,000M) at fixed partial pressures of hydrogen sulfide has been investigated at 25°C. The hydrogen ion concentration and the total concentration of manganese(II) ions in equilibrium with the solid phase have been determined byemf and analytical methods respectively. The data could be explained by assuming the reaction
$$\alpha --MnS_{(s)} + 2H_{(l)}^ + = Mn_{(1)}^{2 + } + H_2 S_{(g)} $$  相似文献   

9.
The structural stabilities, bonding nature, electronic properties, and aromaticity of bare iridium trimers \(\rm{Ir}_3^{+/-}\) with different geometries and spin multiplicities are studied at the DFT/B3LYP level of theory. The ground state of the \(\rm{Ir}_3^{+}\) cation is found to be the 3A2 (C2v) triplet state and the ground state of the \(\rm{Ir}_3^{-}\) anion the 5A2 (C2v) quintet state. A detailed molecular orbital (MO) analysis indicates that the ground-state \(\rm{Ir}_3^{+}\) ion (C2v, 3A2) possesses double (σ and partial δ) aromaticity as well as the ground-state \(\rm{Ir}_3^{-}\) ion (C2v, 5A2). The multiple d-orbital aromaticity is responsible for the totally delocalized three-center metal-metal bond of the triangular Ir3 framework. \(\rm{Ir}_3^{-}\) (C2v, 1A1) structure motif is perfectly preserved in pyramidal Ir3M0/+ (Cs, 1A′) and bipyramidal \(\rm{Ir}_3M_2^{+/3+}\) (C2v, 1A1) (M = Li, Na, K and Be, Ca) bimetallic clusters which also possess the corresponding d-orbital aromatic characters.  相似文献   

10.
Densities, ρ, and viscosities, η, of pure isobutanol, 1-amino-2-propanol, and 1-propanol, along with their binary mixtures of {x 1isobutanol + x 21-propanol}, {x 11-amino-2-propanol + x 21-propanol}, and {x 11-amino-2-propanol + x 2isobutanol} were measured over the entire composition range and at temperatures (293.15–333.15) K at ambient pressure (81.5 kPa). Excess molar properties such as the excess molar volume, V m E , partial molar volumes, \( \bar{V}_{1} \) and \( \bar{V}_{2} \), excess partial molar volumes, \( \bar{V}_{1}^{\text{E}} \) and \( \bar{V}_{2}^{\text{E}} \), thermal expansion coefficient, α, excess thermal expansion coefficient, α E, viscosity deviation, Δη, and the excess Gibbs energy of activation, ?G E*, for the binary mixtures were calculated from the experimental values of densities and viscosities. The excess values of the binary mixtures are negative in the entire composition range and at all temperatures, and increase with increasing temperature. Viscosity deviations, Δη, are negative over the entire composition range and decrease with increasing temperature. The viscosities of the mixtures were correlated by the models of McAllister, Heric, Hind, Katti, and Nissan. The obtained data were correlated by Redlich–Kister equation and the fitting parameters and standard deviations were determined.  相似文献   

11.
By means of the formula \(e^{ - ((r^{XH} - r_0^{XH} )/b^{XHX} )^{5/3} } + e^{ - ((r^{YH} - r_0^{YH} )/b^{YHY} )^{5/3} } = 1\) that characterizes the correlation between the parameters of XH?Y linear fragments (r 0 XH , r 0 YH are the bond lengths in free molecules, b XHX, b YHY are the dimensional coefficients) r XX(r XH) and r XY(r YH) dependences are obtained. Given the length of a hydrogen bridge formed by O, N, and F atoms, they enable us to find the proton position in the bridge. The definition “a quasi-symmetric hydrogen bond,” based on the invariance of the r XX distance when the proton shifts by 0.1 Å, is established to be applicable to OHO, FHF, NHN, and ClHCl fragments. It is shown that the hydrogen bridge length remains almost constant (exceeds the minimum length by no more than 0.1 Å) if its bond orders are above 0.1. Here the displacement of the central proton can reach 0.2–0.3 Å.  相似文献   

12.
In this study, a hydrolysis model for lead, applicable to high ionic strength, is developed based on lead oxide solubilities as a function of ionic strength. Solubility measurements on lead oxide, α-PbO (tetragonal, red), mineral name litharge, as a function of ionic strength were conducted in NaClO4 solutions up to I?=?0.45 mol·kg?1, in NaCl solutions up to I?=?5.0 mol·kg?1, and in Na2SO4 solutions up to I?=?5.4 mol·kg?1, at room temperature (22.5?±?0.5 °C). The lead hydroxyl species considered in this work include the following,
$$ {\text{PbO}}\left( {\text{cr}} \right) \, + {\text{ 2H}}^{ + } \rightleftharpoons {\text{Pb}}^{ 2+ } + {\text{ H}}_{ 2} {\text{O}}\left( {\text{l}} \right) $$
(1)
$$ {\text{Pb}}^{ 2+ } + {\text{ H}}_{ 2} {\text{O}}\left( {\text{l}} \right) \rightleftharpoons {\text{PbOH}}^{ + } + {\text{ H}}^{ + } $$
(2)
$$ {\text{Pb}}^{ 2+ } + {\text{ 2H}}_{ 2} {\text{O}}\left( {\text{l}} \right) \rightleftharpoons {\text{Pb}}\left( {\text{OH}} \right)_{ 2} \left( {\text{aq}} \right) \, + {\text{ 2H}}^{ + } $$
(3)
$$ {\text{Pb}}^{ 2+ } + {\text{ 3H}}_{ 2} {\text{O}}\left( {\text{l}} \right) \rightleftharpoons {\text{Pb(OH}})_{3}^{ - } + 3{\text{H}}^{ + } $$
(4)
The equilibrium constants for Reactions (1) and (2) were taken from literature. The equilibrium constants in base 10 logarithmic units for Reactions (3) and (4) are determined in this study as ? 17.05?±?0.10 (2σ) and ? 27.99?±?0.15 (2σ), respectively, with a set of Pitzer parameters describing the interactions with Na+, Cl?, and \( {\text{SO}}_{4}^{2 - } .\) In combination with the parameters from literature including those that have already been published by our group, the solution chemistry of lead in a number of media including NaCl, MgCl2, NaHCO3, Na2CO3, Na2SO4, NaClO4, and their mixtures, can be accurately described in a wide range of ionic strengths.
  相似文献   

13.
Triple phosphates A2FeTi(PO4)3 (A = Na, Rb) were synthesized by the solid-phase method and studied by electronic microscopy, electron probe X-ray microanalysis, and IR and Mössbauer spectroscopy. The crystal structure of the obtained compounds was refined by X-ray powder diffraction (the Rietveld method). The unit-cell parameters are as follows: for Na2FeTi(PO4)3 (space group R \(\overline 3 \) c, Z = 6), a = 8.6015(1) Å, c = 21.718(1) Å, V = 1391.52(1) Å3; for Rb2FeTi(PO4)3 (space group P213, Z = 4), a = 9.8892(2) Å, V = 967.12(1) Å3. The base of the crystal structures is a mixed octahedral-tetrahedral framework {[FeTi(PO4)3]2?}3∞. Na+ and Rb+ cations are arranged in cavities of the framework. The influence of cationic substitutions on the change of the structural type of the isoformular compounds A2FeTi(PO4)3 (A = Na, Rb) was considered.  相似文献   

14.
Micro hydration structures of the sodium ion, [Na(H2O) n ]+, n = 1–12, were probed by density functional theory (DFT) at B3LYP/aug-cc-pVDZ level in both gaseous and aqueous phase. The predicted equilibrium sodium–oxygen distance of 0.240 nm at the present level of theory. The four-, five- and six-coordinated cluster can transform from each other at the ambient condition. The analysis of the successive water binding energy and natural charge population (NBO) on Na+ clearly shows that the influence of Na+ on the surrounding water molecules goes beyond the first hydration shell with the hydration number of 6. The Car-Parrinello molecular dynamic simulation shows that only the first hydration sphere can be found, and the hydration number of Na+ is 5.2 and the hydration distance (rNa–O) is 0.235 nm. All our simulations mentioned in the present paper show an excellent agreement with the diffraction result from X-ray scattering study.  相似文献   

15.
Densities, ρ, viscosities, η, and refractive indices, n D, for 1-hexyl-3-methyl imidazolium chloride ([hmim]Cl) (IL), 1-pentanol, and ethylene glycol (EG), and for the binary mixtures {x 1[hmim]Cl + x 21-pentanol} and {x 1[hmim]Cl + x 2EG} were measured over the entire composition range at temperatures (293.15–333.15) K and ambient pressure. The excess molar volumes, \( V_{\text{m}}^{\text{E}} \), and viscosity deviations, Δη, for the binary mixtures were calculated from the experimental data. The \( V_{\text{m}}^{\text{E}} \) values of {x 1[hmim]Cl + x 21-pentanol} mixtures are negative over the entire composition range at all temperatures, and increase with increasing temperature in the alcohol rich region and decrease with increasing temperature in the IL rich range. The \( V_{\text{m}}^{\text{E}} \) values of {x 1[hmim]Cl + x 2EG} mixture are positive in the alcohol rich range and negative in the IL rich range at all temperatures, and decrease with increasing temperature. Viscosity deviations of both mixtures are negative over the entire composition range at all temperatures and decrease with increasing temperature. The excess molar properties were correlated by Redlich–Kister equation, and the excess molar volumes were correlated using the PFP model. The fitting parameters and standard deviations were determined.  相似文献   

16.
Madelung's coefficientM a of aragonite has been calculated considering the non-spherical shape of the CO 3 2? -ions. As a result of the multipole expansionM a has been found as a function of the C?O-distanced and the charge on the oxygen atomq o to:
$$\begin{gathered} M_a = \frac{1}{4}\left\{ {10,4446---\left[ {0,65849 + \sum\limits_{n = 1}^{10} {A_n \left( {\frac{{d---0,8}}{a}} \right)^n } } \right]} \right\} \cdot q_o \hfill \\ \left. \begin{gathered} \hfill \\ ---\left[ {0,11066 + \sum\limits_{n = 1}^{12} {B_n \left( {\frac{{d---0,8}}{a}} \right)} ^n } \right] \cdot q_o^2 \hfill \\ \end{gathered} \right\}. \hfill \\ \end{gathered}$$  相似文献   

17.
The pure Znm (m?=?2–10), mixed ZnmOm (m?=?1–10), ZnmO10??m (m?=?1–9) clusters and the univalent and divalent ring-like ZnmOm (m?=?2–10) cluster ions are systematically investigated by using Amsterdam density functional (ADF) program with Triple-zeta with two polarization functions basis set in conjunction with self-consistent field. Our calculated results show that the Zn4 and Zn7 clusters are the magic clusters. The structures of the ZnmOm (m?=?1–10) clusters evolve from two-dimension to three-dimension after m?=?8. For the ZnmO10??m (m?=?1–9) clusters, the Zn-rich structures evolve gradually from three-dimension to plane with an increase in the O ratios. The Zn5O5 cluster with equal ratio has a two dimensional structure. In the O-rich clusters, the O dimers can be easily detached from them. The O and Zn atoms partly adopt sp2 and sp hybridization, respectively, in the ring-like ZnmOm (m?=?2–10) clusters and their ions. Gain and loss charge would affect the degree of hybridization and change their geometries. Their structural changes can be explained by valence bond theory.  相似文献   

18.
The dehydrated form of (Li,Na)-substituted analcime, Li1.30Na0.53[Al1.83Si4.17O12], has been prepared and investigated with single crystal X-ray diffraction: a = 32.167(6) Å, b = 18.551(2) Å, c = 11.693(2) Å; β = 90.06(1)°, V = 6978(1) Å3, Z = 24, space group C2. The structure was analyzed through considering the aluminosilicate framework as a system of tubes composed from corrugated 6-membered rings joint by triples of tetrahedra. Volume decrease by 6.5% and trigonal distortion of the structure are explained by the localization of the non-framework cations in new unusual positions. On dehydration of Li, Na-analcime, 67% of Na+ and 20% of Li+ migrated from the standard M-positions at the periphery of the tubes into essentially different positions NaW and LiL situated on the axes of the tubes. Among the total of the fixed tube positions— 12NaW and 16LiL — one half is aggregated in the tubes parallel to [001] and has a planar three-fold coordination by framework O-atoms. The configuration and cation population of the tubes in other directions follow the motif of the “basic” system.  相似文献   

19.
The kinetics and mechanism of base hydrolysis of tris(3-(2-pyridyl)-5,6-bis(4-phenyl sulphonic acid)-1,2,4-triazine)iron(II), \({\text{Fe}}({\text{PDTS}})_{3}^{4 - }\) have been studied in aqueous, sodium dodecyl sulphate (SDS) and cetyltrimethyl ammonium bromide (CTAB) media at 25, 35 and 45 °C under pseudo-first-order conditions, i.e. \(\left[ {\text{OH}^{ - } } \right]\) ? \({\text{Fe}}({\text{PDTS}})_{3}^{4 - }\). The reaction is first order each in \({\text{Fe}}({\text{PDTS}})_{3}^{4 - }\) and hydroxide ion. The rate increases with increasing ionic strength in aqueous and SDS media, whereas this parameter has little effect in CTAB. In SDS medium, the rate-determining step involves the reaction between \(\left[ {\text{OH}^{ - } } \right]\) and \({\text{Fe}}({\text{PDTS}})_{3}^{4 - }\), whereas in CTAB medium, it involves reaction between a neutral ion pair, {\({\text{Fe}}({\text{PDTS}})_{3}^{4 - }\)·4CTA+} and \(\left[ {\text{OH}^{ - } } \right]\) ions. The specific rate constants and thermodynamic parameters (E a, ΔH #, ΔS # and ΔG 35°C # ) have been evaluated in all three media. The near equal values of ΔG 35°C # obtained in aqueous and SDS media suggest that these reactions occur essentially by the same mechanism. Slightly lower ΔG 35°C # values in CTAB medium can be attributed to a higher concentration of reactants in the Stern layer. The reaction is inhibited in SDS medium but catalysed in CTAB. The former can be attributed to the anionic surfactant creating more repellent space between the reactants. Catalysis in CTAB medium is ascribed to electrophilic and hydrophilic interactions between hydroxide ion/substrate with the cationic Stern layer, resulting in increased local concentrations of both reactants.  相似文献   

20.
Extraction of Np4+ and \( {\text{NpO}}_{2}^{2 + } \) was carried out from nitric acid feeds using solutions of N,N,N′,N′-tetra-n-octyldiglycolamide (TODGA) in two imidazolium-based room temperature ionic liquids, viz., 1-butyl-3-methylimidazolium bis(trifluoromethanesulphonyl) imide ([C4mim][NTf2]) and 1-octyl-3-methylimidazolium bis(trifluoromethanesulphonyl) imide ([C8mim][NTf2]). The extraction equilibrium was attained within 2 h for both the metal ions in both the ionic liquids. While a cation exchange mechanism is proposed for the extraction of \( {\text{NpO}}_{2}^{2 + } , \) an ion-pair mechanism of extraction is proposed for the Np4+ ion. The nature of the extracted species was determined by carrying out experiments at varying concentrations of TODGA, and species of the type Np(L)2(NO3)4 and NpO2(L)2+ were found to be extracted in 3 mol·dm?3 HNO3. The identification of these extracted species was also supported from the variable nitrate and C4mim+ ion concentration experiments.  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