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 312 毫秒
1.
Bromo dimethyl sulfoxide osmium(II) complexes were synthesized: trans-[OsBr2(dmso-S)4] (1) was obtained by the reaction of K2[OsBr6] with DMSO in the presence of SnBr2 at 100°C and cis,fac-[OsBr2(dmso-S)3(dmso-O)] (2) was prepared by thermal isomerization of 1 in a DMSO solution at 150°C. The coordination mode of DMSO molecules was determined by IR and 1H and 13C NMR spectroscopy. X-ray diffraction analysis showed that compound 2 crystallizes in the monoclinic system, space group P21/n; a = 8.4711(5) Å, b = 27.7876(15) Å, c = 8.5569(5) Å, β = 115.7110(10)°; Z = 4. The coordination polyhedron of osmium is a distorted octahedron; the osmium environment is formed by two cis-arranged bromine atoms and three fac-S-coordinated and one O-coordinated DMSO molecules. The interconversion of complexes in solutions was studied by UV/Vis and 1H and 13C NMR spectroscopy. In chloroform and DMSO, complex 2 isomerizes to cis-[OsBr2(dmso-S)4] and (in the light) to 1. The complexes trans-[OsX2(dmso-d6)4], where X = Cl, Br, were isolated from DMSO-d 6 and characterized by the IR spectra.  相似文献   

2.
The binuclear copper(II) complex [Cu2L(CH3COO)] (I), where L3? is the azomethine trianion based on 3-methyl-4-formyl-1-phenylpyrazol-5-one and 1,3-diaminopropan-2-ol, and its DMSO adduct (II) in which the DMSO molecule acts as an additional bridging ligand are synthesized. The structure of complex II is determined by X-ray diffraction analysis, and the structure parameters of the coordination unit of complex I are determined by EXAFS spectroscopy. The μ2-coordination of the DMSO molecule in compound II results in a change in the sign of the exchange interaction parameter. In complex I, the antiferromagnetic exchange interaction (2J = ?169 cm?1) occurs between the copper(II) ions. The exchange interaction of the ferromagnetic type (2J = 174 cm?1) is observed in complex II. The quantum-chemical calculations of the magnetic exchange parameters by the density functional theory method show that the role of the DMSO molecule as a switch of the exchange interaction character is exclusively the stabilization of the “broken” conformation of the metallocycles.  相似文献   

3.
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   

4.
Vapor pressures for the system I (dimethyl sulfoxide/H2O=DMSO/H2O) and isotopic differential pressures I-II (II=DMSO/D2O) have been measured between 25 and 70°C at DMSO concentrations of 0.05, 0.15, 0.30, 0.45, 0.60, 0.70, 0.80, 0.87, and 0.92 mole fraction. A high-precision differential method was used. The total pressures over the solutions, I, have been fitted to a relation derived from the Duhem-Margules equation, P T =P 1 o X1γ1+P 2 o X2γ2, with γ1=exp[∑kαkX 2 k ] and $$\gamma _2 = exp[\sum \alpha _k X_2^k - \sum (\alpha _k /(k - 1))(kX_2^{k - 1} - 1)].$$ . The αk are parameters andk is a number ≥2. The αk were taken as temperature dependent. Four parameters sufficed to fit the data within experimental error. Excess partial molal properties derived from the fits are in quantitative agreement with earlier literature results derived from the directly measured partial pressures, but the present data extend over a wider temperature range. The isotopic differential pressures I-II were similarly fitted to the relation above. The excess free energies and enthalpiesG I E andH I E are large and negative. The isotope effects ΔG I,II E =G I E ?G II E and ΔH I,II E are negative. They are discussed in detail in terms of the theory of isotope effects in condensed phases and demonstrated to be consistent with that theory and with the available spectroscopic data. A small amount of enthalpy data for the solution of DMSO in HOH and DOD is reported.  相似文献   

5.
A titration method utilizing glass electrodes and silver-silver chloride electrodes in a cell without liquid junction has been used to determine the acidic dissociation constants at 15, 25, and 35°C of nine protonated nitrogen bases in mixtures of water and dimethyl sulfoxide (DMSO). The mole fraction of DMSO in the mixed solvents was 0.2, 0.4, 0.6, and 0.8. The cell was calibrated with HCl (molality=0.01 mole-kg?1) in the mixed solvents, and the ionic strength and chloride molality remained substantially unchanged during the titration with added base. This method minimizes the errors resulting from the formation of AgCl 2 ? in the media rich in DMSO. The pK a of all the protonated bases passes through a minimum at a solvent composition close to that at which H2O-DMSO mixtures display a maximum solvent structure. The results are discussed in terms of the preferential solvation of ions by the two types of solvent molecules. They are consistent with the hypothesis that increased solvent structure is accompanied by increased desolvation of the cation acids.  相似文献   

6.
The liquidus temperature was measured in the ammonium nitrate-dimethylsulfoxide-water system over in the concentration range 0–60 mole% ammonium nitrate. The probable formation of the NH4NO3·nDMSO solvate with n=1.3–1.5 and the mixed solvate NH4NO3·DMSO·H2O at 30 mole% ammonium nitrate and a DMSO:H2O ratio of 4∶1 are indicated. The glass transition temperatures T g were measured over a salt concentration range of 0–50 mol% ammonium nitrate and at various compositions of the mixed solvent (y DMSO =0.1–0.9 mole fraction). At a constant mixed solvent composition, the dependence of the glass transition temperature on the salt concentration can be approximated by a linear relationship, as can its dependence on the DMSO content in the solution at constant salt concentration. The glass-forming composition regions were found and the limits of this region are discussed.  相似文献   

7.
Aromatization of 4-aryl(hetaryl)tetrahydroimidazo[4,5-c]pyridine-6-carboxylic acids and their lithium salts by the action of dimethyl sulfoxide has been revealed for the first time. Heating of these compounds in DMSO for 5–7 h at 90–95°C leads to the formation of 4-aryl(hetaryl)imidazo[4,5-c]pyridine derivatives as a result of dehydrogenation and decarboxylation. Heating of the corresponding lithium salts generated in situ (DMSO, 90–95°C, 3–5 h) affords difficultly accessible 4-aryl(hetaryl)imidazo[4,5-c]pyridine-6-carboxylic acids.  相似文献   

8.
A competitive solvation study of Al(ClO4)3, Ga(ClO4)3, In(ClO4)3, UO2(ClO4)2, and UO2(NO3)2 in water-acetone-dimethylsulfoxide (DMSO) and water-acetone-hexamethylphosphoramide (HMPT) mixtures has been carried out by direct H1 and P31 nuclear magnetic resonance (NMR) techniques. At low temperature, proton and ligand exchange are slow enough in these systems to permit the observation of signals for bulk and coordinated molecules of water and the organic bases (DMSO and HMPT). Both DMSO and HMPT compete effectively with water for coordination sites in the Al3+, Ga3+, and In3+ systems, with steric effects dominating the HMPT results. Both Al3+ and In3+ are able to bind a maximum of two to three HMPT molecules, for example. In contrast, UO2+ is solvated selectively by the organic molecules to the allowed maximum of 4 molecules per cation. H1 and P31 NMR spectral results support the formation of only the mono-, tri-, and tetra-HMPT solvation complexes.  相似文献   

9.
Free energies of transfer of ions from water to mixtures of water with acetonitrile (AN), with dimethylformamide (DMF), with dimethylsulfoxide (DMSO), and with ethylene glycol have been determined using both the tetraphenylarsonium tetraphenylboride [TATB] and the negligible liquid junction potential [E j ] assumptions. By making use of ΔG tr (Ag+)[TATB]=12 kJ-mol?1 for transfer from DMSO to AN and by assuming negligible liquid junction potential in the cell $${\text{Ag|AgNO}}_{\text{3}} {\text{(0}}{\text{.01}}M{\text{),S}}\parallel {\text{Et}}_{\text{4}} {\text{NPic(0}}{\text{.1}}M{\text{),AN}}\parallel {\text{AgNO}}_{\text{3}} {\text{(0}}{\text{.01}}M{\text{),AN|Ag}}$$ single ion free energies of transfer of silver ion ΔG tr (Ag+)[E j ] from DMSO to 35 pure and mixed solvents show a standard deviation of only 2 kJ-mol?1 when compared with ΔG tr (Ag+) calculated from the TATB assumption that ΔG tr (Ph 4 As+)=ΔG tr (Ph 4 B?). The ferrocene assumption [Fc] also gives acceptable agreement with ΔG tr (Ag+)[TATB] provided that the solvents are not highly aqueous. Other cells with other junctions give less acceptable agreement between the E j and TATB assumptions. It is essential that the salt bridge is always tetraethylammonium picrate in AN, if the E j assumption is assumed. Because of the ease of making potentiometric measurements compared with the difficulty of measurements required for the TATB assumption, the negligible liquid junction potential method in the cell shown is recommended for estimating transfer free energies of single ions. The ferrocene assumption is acceptable only for non-structured aprotic solvents.  相似文献   

10.
Solvate Ph3Sb[OC(O)C6H4(OH-4)]2 · 1/2Et2O (I) has been synthesized by the reaction between triphenylantimony and 4-oxybenzoic acid in the presence of hydrogen peroxide in diethyl ether. Tetraphenylantimony 4-oxybenzoate, which crystallizes from DMSO in the form of solvate Ph4SbOC(O)C6H4(OH-4) · DMSO (II), has been synthesized from pentaphenylantimony and triphenylantimony bis(4-oxybenzoate) or 4-oxybenzoic acid. According to X-ray diffraction data, an antimony atom in molecules of compounds I and II has a trigonal bipyramidal coordination. Crystals of compound I contain two crystallographically independent types of molecules (A and B). The Sb-C and Sb-O distances, the equatorial CSbC and axial OSbO angles are, respectively, 2.083(9)–2.103(8)Å; 2.068(5), 2.128(5)Å; 117.6(3)°–124.2(3)° and 171.5(2)° (IA); 2.103(9)–2.135(8)Å; 2.086(5), 2.154(6)Å;110.2(3)°–138.0(4)° and 174.8(2)° (IB). In compound II, Sb-C is 2.117(2)–2.175(2) Å, Sb-O is 2.247(2) Å, and CeqSbCeq and OSbCax angles are 110.89(9)°–133.30(9)° and 177.05(7)°, respectively. The Sb…O=C intramolecular contacts are 3.151(7), 3.153(8) Å (IA), 2.985(8), 3.008(9) Å (IB), and 2.975(5) Å (II). Molecules IA and IB are conformation isomers, which differ from each other by the arrangement of carboxyl groups with respect to the equatorial plane.  相似文献   

11.
The thermochemical properties of 2,4-dinitroanisole (DNAN) in N-methy pyrrolidone (NMP) and dimethyl sulfoxide(DMSO) were studied using a RD496-2000 Calvet microcalorimeter at four different temperatures. The heat effects were measured for DNAN dissolved in NMP and DMSO and the relationships between the heat effects and the amounts of the substance were determined. The molar enthalpies and the differential molar enthalpies of dissolution processes were also obtained from the experimental data. The corresponding kinetic equations describing the two dissolution processes at different temperatures were discussed.  相似文献   

12.
Organic solvents are often used in mixture solvent systems to optimize synthetic reactions. However, they may also produce unexpected effects, some of which may be hazardous and cause a runaway reaction and/or lead to an accident. Thus, the proper accident scenarios and thermal risk assessment models are needed to use mixture solvents more safely and efficiently. For chemical process safety, Stoessel suggests the systematic assessment of accident scenarios. However, if scenarios are changed by the properties of mixture solvents, Stoessel’s concept does not cover them. Our previous study evaluated characteristic scenarios of mixture solvents based on Stoessel’s model. In this study, as a characteristic scenario pattern, we focused on the energy release of the solvent and the material derived from degraded solvent and investigated them experimentally using tetrahydrofuran (THF) and dimethylsulfoxide (DMSO) as a representative mixture solvent. From hazard and scenario identification of THF and DMSO, we assumed that THF peroxide and DMSO play roles in energy release. THF containing peroxide and DMSO were mixed, and thermal analysis and chemical composition analysis were performed. Our results indicated that DMSO promotes the decomposition of THF peroxide, and the decomposition temperature of DMSO decreases upon mixing with degraded THF. Therefore, we verified the scenario pattern of energy release of solvent and the material derived from degraded solvent.  相似文献   

13.
A reaction of alk-4-ynals with aliphatic amino alcohols or 2-aminoethanethiol in the system DMSO—KOH gives bicyclic N,O- and N,S-enaminals: 6-methylidenehexahydro-2H-pyrrolo[2,1-b][1,3]oxazines, 5-methylidenehexahydropyrrolo[2,1-b]oxazoles, or 5-methyl-idenehexahydropyrrolo[2,1-b]thiazoles. The reaction proceeds through the formation of equilibrium mixtures of the corresponding imines and monocyclic aminals with subsequent 5-exodig-cyclization catalyzed by the superbasic system DMSO-KOH.  相似文献   

14.
Bis[1-(methoxyphenyl)propan-2-yl]phosphines and bis[1-(methoxyphenyl)propan-2-yl]phosphine oxides were synthesized by phosphorylation of allyl(methoxy)benzenes in the system red phosphorus-KOH · 0.5 H2O-DMSO under microwave irradiation.  相似文献   

15.
The partial molal heats of solution ΔH s o and the partial molal heat capacities of solution ΔC p o of the bolaform salts [Et3N(CH2)nNEt3]Br2 and [allyl3N(CH2)nNallyl3]Br2 have been obtained at infinite dilution in dimethylsulfoxide (DMSO). A comparison of these data with the results of previous thermodynamic studies of the same solutes in aqueous solvents has been carried out. The observed differences have been interpreted in terms of solute-induced solvent structural effects occurring in aqueous solvent media. Partial molal heat capacities of the bolaform salts at infinite dilution in DMSO, H2O, and D2O have been calculated from ΔC p o data and previously reported values of the heat capacities of the crystalline state. The data clearly show that the structure-promoting capabilities of these salts in aqueous solvents increase with increasing hydrocarbon content. A comparison of contributions to partial molal heat capacities of methylene groups in the bolaform and R4N+ series of salts reveals that similarities exist between the solvation effects of CH2 groups in the normal alkyl chain of the R4N+ cations and in the bridging alkyl chain of the bolaform cation.  相似文献   

16.
The aggregation behavior of two silicone surfactants (monomeric and Gemini) was studied by surface tension measurements in a room temperature ionic liquid, ethylammonium nitrate (EAN), at various temperatures. A series of parameters, including critical micelle concentration (CMC), surface tension at the CMC (γ CMC), adsorption efficiency (pC 20), and effectiveness of surface tension reduction (Π CMC), were obtained. By comparing the silicone surfactants with traditional surfactants, we deduced that the surface activity of the silicone surfactants in EAN was superior to the activity of other surfactants. In addition, from the CMC values and their temperature dependence, we estimated the thermodynamic parameters of the micelle formation, $ \Delta G_m^0 $ , $ \Delta H_m^0 $ , and $ \Delta S_m^0 $ . It was revealed that the micellization of the silicone surfactants is entropy driven at low temperature and enthalpy driven at high temperature. Isothermal titration calorimetry measurements were also carried out to study the micellization of Gemini silicone surfactant. 1H NMR was performed to study the silicone surfactant micelle formation mechanism in EAN.  相似文献   

17.
Poly(2-vinylpyridine)s (P2VPs) are important polymers with extensive applications in modern day material science. P2VP is an exceptional case for liquid chromatography because of certain polar interactions with most of the stationary phases. In the present study, we established the critical adsorption point (CAP) of P2VP for the first time. The effectiveness of the method is demonstrated by analyses of blends and block copolymers of P2VP and PMMA. The CAP of PMMA is established for determination of molar mass of P2VP component of above mentioned blends and block copolymers. The methods successfully demonstrate the separation of both types of homopolymers from the rest of the samples in conjunction with the determination of molar mass distribution of noncritical block or component. Graphical Abstract
?  相似文献   

18.
To realize highly selective relay recognition of Fe3+ and H2PO4- ions, a simple benzimidazole-based fluorescent chemosensor(L) was designed and synthesized. Sensor L displays rapid, highly selective, and sensitive recognition to Fe3+ in H2O/DMSO(1:1, v/v) solutions. The in situ-generated L-Fe3+ complex solution exhibits a fast response and high selectivity toward dihydrogen phosphate anion via the Fe3+ displacement approach. The detection limits of sensor L to Fe3+ and L-Fe3+complex to H2PO4- anion were estimated to be 1.0 × 10-9 mol/L. Notably, the sensor was retrievable to indicate dihydrogen phosphate anions with Fe3+, and H2PO4-, in turn, increased. This successive recognition feature of sensor L makes it a potential utility for Fe3+ and H2PO4- anion detection in aqueous media.  相似文献   

19.
Application of size exclusion chromatography (SEC) for the analysis of cellulose samples is often limited due to poor solubility in the solvent system N,N-dimethylacetamide/lithium chloride (DMAc/LiCl). Hence different activation or derivatization methods have been developed and published. Most of these methods are laborious, influence the molar mass distribution or do not support dissolution of manmade fibers, such as viscose rayon. In this study, we have evaluated different activation methods for their applicability in viscose rayon dissolution and we present a novel method for activation. We found that an additional solvent exchange step with dimethyl sulfoxide (DMSO) increases and accelerates solubility of viscose fibers in DMAc/LiCl for subsequent SEC analysis. The improved dissolution by DMSO activation is mainly due to increased swelling and improved action towards the outer skin of the fiber. The novel approach has also been applied to the even more difficult dissolution of oxidized viscose fibers.  相似文献   

20.
2-(2-Furyl)-1H-naphto[2,3-d]imidazole was synthesised by interaction of 2,3-diaminonaphtalene with furfural in the conditions of Weidenhagen reaction. At its N-methylation in the KOH-DMSO system the 1-methyl-2-(2-furyl)-1H-naphto[2,3-d]imidazole was obtained, which we subjected to reactions of electrophilic substitution (nitration, bromination, sulfonation, formylation, and acylation). This compound possesses a specific reactivity: all reactions, as a rule, proceed at the naphthalene scaffold while the furan ring is not involved.  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