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 687 毫秒
1.
Li  Siwen  Yu  Hong  Ma  Yajie 《Chromatographia》2011,74(11):759-765

A method of ion-pair chromatography was developed on a reversed-phase silica-based monolithic column for the fast and simultaneous determination of trifluoromethanesulfonate (CF3SO3 ) and p-toluenesulfonate (C7H7SO3 ). The analysis was performed using a mobile phase of tetrabutylammonium hydroxide + citric acid + acetonitrile on the Chromolith Speed ROD RP-18e column with direct conductivity detection. The effects of the eluent, column temperature and flow rate on the retention of the anions were investigated. The experimental phenomenon was discussed according to hydrophobic interaction and ion-exchange mechanism in the separation. The optimized chromatographic conditions were selected. The optimized eluent for the separation consisted of 0.2 mmol L−1 tetrabutylammonium hydroxide + 0.10 mmol L−1 citric acid + 9% acetonitrile (pH 5.5). The flow rate was set at 6.0 mL min−1. The column temperature was 25 °C. Under the optimal conditions, the better separation of CF3SO3 and C7H7SO3 was achieved without any interference by other anions (Cl, Br, I, NO3 , SO4 2−, ClO3 , BF4 and PF6 ). The detection limit (S/N = 3) was 0.28 and 0.71 mg L−1 for CF3SO3 and C7H7SO3 , respectively. The method has been applied to the determination of CF3SO3 and C7H7SO3 in ionic liquids. The spiked recoveries of CF3SO3 and C7H7SO3 were 101.1 and 100.2%, respectively.

  相似文献   

2.
Isopiestic vapor pressure experiments are performed at 25°C with aqueous NaTcO4, mixed NaTcO4-NaCl, and NaCl reference solutions. The solubility of CsTcO4 is determined in 0–7.4m CsCl and in 0–5.6m NaCl solutions. The osmotic coefficients of the binary and ternary solutions are used to evaluate the binary Pitzer parameters Β0, Βl, and CΦ for NaTcO4 and the mixing parameters θTcO 4 - /cl- and ψNa +/TcO 4 - /cl-. The binary Pitzer parameters for the sparingly soluble CsTcO4 and CsC1O4 are calculated together with ψCs +/mo 4 - /cl- from their solubilities in CsCl solution. The solubilities of CsTcO4 in NaCl and CsClO4 in NaClO4 solution are also included in the parametrization of the reciprocal salt systems Na+/Cs+/Cl-/MO 4 - . The parameters Β0 and Βl of pertechnetate and perchlorate salts correlate well with the ionic radii.  相似文献   

3.
The mass spectra of 13C-labelled 2-phenylthiophenes and 2,5-diphenylthiophenes were studied. The label distributions for the [HCS]+, [C2H2S], [C8H6], [C9H7]+ and [C7H5]S+ ions from 2-phenylthiophene and the [HCS]+, [C9H7]+, [C7H5S], and [C15H11]+ ions from 2,5-diphenylthiophene were interpreted in terms of both carbon skeletal rearrangements in the thiophene ring and migration of the phenyl substituent. The degree of carbon scrambling in the thiophene ring appeared to be almost independent of the electron beam energy. The formation of some of the fragment ions studied seems to be so fast that no carbon scrambling could be detected at all; in neither case was complete scrambling of the carbon atoms of the thiophene ring observed.  相似文献   

4.
The complexes (η5-C5H5)Pd(η1-C5H5)PR3 which are prepared from [Cl(PR3)-Pd]2(μ-OCOCH3)2 and TlC5H5 are fluxional in solution. According to the 1H and 13C NMR spectra at various temperatures, two dynamic processes occur. The process with the higher activation energy is a π/σ (η51) exchange of the two different cyclopentadienyl ligands, whereas the second one with the lower activation energy presumably is a metallotropic rearrangement (1,2-shift). The coalescence temperature for the η51 exchange depends on the size of the phosphine. The X-ray structural analysis of (C5H5)2PdPPri3 proves that it exists as a “frozen” η5 + η1 structure in the crystal with the palladium approximately in a square-planar coordination. The η5-bonded cyclopentadienyl ring shows some unusual bonding patterns which are obviously electronic in nature. EHT-MO calculations for (η5-C5H5)PdCH3(PH3) indicate that in this model system alternating CC distances in the ring and a stronger bond of the metal to one of the five carbon atoms of the C5H5 ligand are to be expected. The calculations suggest that in similar complexes possessing a six-electron donor ligand like C5H5? and a metal fragment which is isolobal to PdCH3(PH3)+, analogous distortions should be observed. Some reactions of the compounds (η5-C5H5)Pd(η1-C5H5)PR3 are described.  相似文献   

5.
The distribution coefficients (DC) for HgCl 4 2– , Hg(SO4) 2 2– , Hg(NO3) 4 2– , Ag+, Ag(SCN) 2 and Ag(NH3) 2 + between aqueous solutions and Dowex A-1 were measured in varying hydrogen ion concentrations. The DC of Ag+ in the NO 3 media was very low (4 to 6). The DC for the Ag(SCN) 2 complex decreased as pH increased. The Ag(NH3) 2 + complex had a constant DC of about 65 from pH 8 and above. The trend observed for three mercury complexes in HCl, H2SO4 and HNO3 was similar; the DC decreased steadily from 0.1M to 6M. The HgCl 4 2– complex had the highest DC (9000) while the Hg(NO3) 4 2– complex had the lowest DC (2000).  相似文献   

6.
The paper presents a radiokinetic study on the appearance and growth of*Fe2S3,*Fe(OH)3,*Fe2(C2O4)3,*Fe(IO3)3 crystals in a colloidal medium of agar and gelatine. The values of the diffusion constants through gels of55+59Fe3+ radioactive cations and of the rate of global growth process of these crystals in agar or gelatine were calculated using the experimental data. A new method for the determination of the starting time for the complex nucleation process was proposed. The formation rate of crystals under study decreases in the order:*Fe(OH)3>*Fe(IO3)3>*Fe2S3>*Fe2(C2O4)3, in agar medium and*Fe(OH)3>*Fe(IO3)3>*FeC2O4)3>*FeS3, in gelatine medium.  相似文献   

7.
The gas‐phase reaction of CH3+ with NF3 was investigated by ion trap mass spectrometry (ITMS). The observed products include NF2+ and CH2F+. Under the same experimental conditions, SiH3+ reacts with NF3 and forms up to six ionic products, namely (in order of decreasing efficiency) NF2+, SiH2F+, SiHF2+, SiF+, SiHF+, and NHF+. The GeH3+ cation is instead totally unreactive toward NF3. The different reactivity of XH3+ (X = C, Si, Ge) toward NF3 has been rationalized by ab initio calculations performed at the MP2 and coupled cluster level of theory. In the reaction of both CH3+ and SiH3+, the kinetically relevant intermediate is the fluorine‐coordinated isomer H3X‐F‐NF2+ (X = C, Si). This species forms from the exoergic attack of XH3+ to one of the F atoms of NF3 and undergoes dissociation and isomerization processes which eventually result in the experimentally observed products. The nitrogen‐coordinated isomers H3X‐NF3+ (X = C, Si) were located as minimum‐energy structures but do not play an active role in the reaction mechanism. The inertness of GeH3+ toward NF3 is also explained by the endoergic character of the dissociation processes involving the H3Ge‐F‐NF2+ isomer. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

8.
The complexes [(ν3-RC3H4)Pd{R1NS(R2)NR1}] (R  H, CH3;R1  aryl; R2  CH3, t-C4H9) have been obtained from the reaction of [(ν3-RC3H4)PdCl]2 with [Li{R1NS(R2)NR1}]; two isomers are produced, differing in the orientation of the allyl group. The sulfurdiimino group has some π-allylic character. The compounds decompose in solution into azo—arenes and [(ν3-RC3H4)Pd(SR2)]2, and this is shown to be dependent upon steric and electronic factors.The properties of the sulfurdiimino compounds are compared with those of the compounds [(ν3-RC3H4)Pd(R3N3R3)]2 and [(ν3-RC3H4)Pd(R3NC(R4)NR3)]2 (R  H, CH3;R3  CH3, aryl; R4  H, CH3), which have been prepared by new methods.  相似文献   

9.
Reaction of Telluriumhexafluoride and Trimethylamine, Structures of the TeF5? and SeF5? Anions The reaction of TeF6 and (CH3)3N is of the redox kind, resulting in reduction of tellurium: X-ray single crystal analysis reveals the compounds (CH3)2N? CH2? N(CH3)2+TeF5? and [(CH3)3NH+]5(TeF5?)3(HF2?)2. By comparison with published data it can be shown that this mixture is identical to previously published [(CH3)3N]2TeF6. The latter was supposed to be one of the few examples of tellurium in a coordination state of eight. (CH3)4N+TeF5? and (CH3)4N4SeF5? are obtained and their structure is investigated by single crystal x-ray methods also. The anions SF5?, SeF5? and TeF5? are discussed in terms of weak interactions.  相似文献   

10.
The production of 1Δg O2 in microwave discharges in CO2 and NO2 has been confirmed by the observation of its EPR spectrum. Quantitative measurements of the EPR spectra of 1Δg and 3Σ+g O2 and 3P O atoms yield conversion efficiencies for these species. These combined measurements its show that the observed 1Δg O2 is not formed by simple excitation of 3Σ+g O2 formed in the discharge. Rather there must exist a direct mechanism for the production of 1Δg O2 from CO2 and NO2. The SO2 discharge products give rise to no 1Δg O2 EPR spectrum.  相似文献   

11.
Equilibrium constants for the fluorinated species HF, F-, HF-2 and H2F2 in formic acid and in a 1M potassium formate solution in formic acid have been studied by 19F NMR. The chemical shifts of these species have been determined from measurements of the shifts for various initial mixtures of differing concentrations of dissolved HF, F- and HF-2. From these values, relative concentrations of HF, F-, and HF-2 and H2F2 in each solution have been calculated through a numerical method. The following constants were obtained: K1 = [H+][F-]/[HF] = 1.1 x 10-5M; KD = [HF][F-]/[HF-2] = 0.5 M; K′1 = [H+][HF-2]/[H2F2]= 1.1 x 10-5 M; K′D = [HF]2/[H2F2]=0.5 M.  相似文献   

12.
The electronic absorption, fluorescence and phosphorescence spectra of s-tetrazine at low temperatures (4.2-1.5 K) are reported and analyzed in the neat crystal and in several mixed crystals. The 3B3u-1Ag (nπ*) origin is at 18414 ± 5 cm?1 for neat tetrazine. In the mixed crystal several sites identified. The lowest energy origin is at 17453 cm?1 for tetrazine in pyrazine; 17 701 cm?1 in pyrimidine; and 17 676 cm?1 in pyridazine. The eB3u-1Ag (nπ*) origin is at 14 096 ± 2 cm?1 for the neat crystal. The phosphorescence lifetime of neat tetrazine is measured to be 96.8 ± 2.1 μs at 4.2 and 1.8 K. All the spectra are predominately composed of members of progressions in a single totally symmetric mode (ν6a) built upon site origins and vibrational fundamentals. The ν6a interval is: 743 (1Ag), 715 (3B3u), and 709 cm?1 (1B3u) in the neat tetrazine crystal; 732 (1Ag) and 705 cm?1 (1B3u in pyrazine host, 737 (1Ag) and 701 cm?1 (1B3u) in pyrimidine host, and 732 (1Ag) and 703 cm?1 (1B3u) in pyridazine host mixed crystals. All emission spectra may be analyzed by Oi → (ν″6a)on (i), i indicating the observed s  相似文献   

13.
为探讨急性心肌梗死(AMI)患者血清中K+、Na+、Ca2+、Fe2+、Mg2+含量变化,并研究其与心肌梗死患者之间的关系。选取2022年5月至2023年2月收治的AMI患者37例,同时选取健康体检者35例作为对照组。依据入院时或体检时收集的抽血样本进行临床生化分析,比较两组间血清K+、Na+、Ca2+、Fe2+、Mg2+含量,采用判别方程、主成分分析法(PCA),判断分析哪种金属离子对于心肌梗死的诊断价值大。结果表明,AMI患者的血清中Ca2+和Fe2+含量低于健康对照组,差异具有统计学意义。基于血钙、铁水平两组具有显著性差异,以它们为基础进行判别分析,获得判别函数式。将血清中K+、Na+、Ca2+、Fe2+、Mg2+  相似文献   

14.
Efficiencies and rates of electron transfer from various electron donors to excited fullerenes (C60 and C70) have been determined by observing the transient absorption bands in the near-IR region, where the anion radicals of fullerenes appear. From the rise of the absorption bands of C60 −+ and C70 −+ in the near-IR region, electron transfer takes place via the triplet states (TC60 * and TC70 *) under appropriately low concentrations of electron donors. By analysis of the rise curves C60 −+ and C70 −+, contribution of the excited singlet states (SC60 * and SC70 *) in addition to the route of the triplet states (TC60 * and TC70 *) is confirmed. The quantum yield for electron transfer via the triplet states Φct T was evaluated by the ratio of [C60 −+]/[TC60 *] (or [C70 −+/[TC70 *]). The Φct T depends upon the donor-ability, donor concentration, and solvent polarity. The back electron-transfer process, which was evaluated by observing C60 −+, also depends upon the solvent polarity.  相似文献   

15.
The kinetics of the base catalysed racemization of [Co(EN3A)H2O]
  • 1 Abbreviations: EN3A3?=(?OOCCH2)2N(CH2)2NHCH2COO?; ME3A3?=(?OOCCH2)2N(CH2)2 N(CH3)CH2COO?; EDDA2?=?OOCCH2NH(CH2)2NHCH2COO?; EDTA4?=(?OOCCH2)2N(CH2)2N(CH2COO?)2;TNTA4?=(?OOCCH2)2N(CH2)3N(CH3COO?)2; HETA3?=(?OOCCH2)2N(CH2)2N(CH2COO?)CH2CH2OH; en=H2N(CH2)2NH2; Meen=H2N(CH2)2NHCH3; sar?=?OOCCH2NHCH3.
  • were studied polarimetrically in aqueous buffer solution. The reaction rate is first order in OH? and in complex, in weakly acidic medium. Activation parameters are ΔH≠=22 kcal · mol?1, ΔS≠=26 cal · K?1. The results are discussed in terms of an SN1CB mechanism involving exchange of the ligand water molecule. The N-methylated analogue [Co(ME3A)H2O] does not racemize in the pH-range investigated. Loss of optical activity occurs at a rate which is about 1,000 times slower than the racemization of [Co(EN3A)H2O](60°) and coincides with the decomposition of the complex.  相似文献   

    16.
    The ionic liquid (IL) trihalogen monoanions [N2221][X3] and [N2221][XY2] ([N2221]+=triethylmethylammonium, X=Cl, Br, I, Y=Cl, Br) were investigated electrochemically via temperature dependent conductance and cyclic voltammetry (CV) measurements. The polyhalogen monoanions were measured both as neat salts and as double salts in 1-butyl-1-methyl-pyrrolidinium trifluoromethane-sulfonate ([BMP][OTf], [X3]/[XY2] 0.5 M). Lighter IL trihalogen monoanions displayed higher conductivities than their heavier homologues, with [Cl3] being 1.1 and 3.7 times greater than [Br3] and [I3], respectively. The addition of [BMP][OTf] reduced the conductivity significantly. Within the group of polyhalogen monoanions, the oxidation potential develops in the series [Cl3]>[BrCl2]>[Br3]>[IBr2]>[ICl2]>[I3]. The redox potential of the interhalogen monoanions was found to be primarily determined by the central halogen, I in [ICl2] and [IBr2], and Br in [BrCl2]. Additionally, tetrafluorobromate(III) ([N2221]+[BrF4]) was analyzed via CV in MeCN at 0 °C, yielding a single reversible redox process ([BrF2]/[BrF4]).  相似文献   

    17.
    Abstract

    The mechanisms and kinetics of oxidation of ascorbate, AH?, by Ni(III)Li aq and by LiNi(III) (HPO4)2 ? complexes (L1 = meso-(5,12)-7,7,14,14-hexamethyl-1,4,8,11-tetraazacyclotetradecane; L2 = 1,8-dimethyl-1,3,6,8,10,13-hexaazacyclotetradecane) in neutral aqueous solutions have been investigated.

    The oxidation of ascorbate by the LiNi(III) (HPO4)2 ? and Ni(III)L1 aq proceeds via two consecutive reactions well separated in time. The products of the first reaction are the A.? radical anion and the corresponding Ni(II) complex. The oxidations by the LiNi(III)(HPO4)2 ? complexes proceed via the outer sphere mechanism, whereas the detailed mechanism of reaction of Ni(III)L1 aq cannot be determined. The rate of reaction decreases with the increase in the concentration of phosphate, thus indicating that LiNi(III)(HPO4)(H2O)+ and LiNi(III)OH2+ are stronger oxidizing agents than LiNi(III)(HPO4)? 2.

    The oxidation of ascorbate by Ni(III)L2 aq proceeds via three consecutive reactions which are well separated in time. Thus the results clearly point out that this process occurs via the inner sphere mechanism. The first transient observed is tentatively identified as L2(H2O)Ni(II)(A.?)2+, i.e., an unexpected complex of the ascorbate anion radical. Also in this process the last transient observed is the A.? anion radical. The stabilization of the ascorbyl radical in a transient complex might be of biological significance.  相似文献   

    18.
    A phenylenediamine‐capped conjugate of calix[4]arene ( Lamino ) was synthesized by reducing its precursor, Limino , with sodium borohydride in methanol. The Lamino sample binds to anions due to the more flexible and bent conformation of the capped aminophenolic binding core, compared to the precursor Limino . The Lamino sample showed selectivity towards H2PO4? by exhibiting a ratiometric increase in emission by about 11‐fold with a detection limit of (1.2±0.2) μm ((116±20) ppb) over 15 anions studied, including other phosphates, such as P2O74?, adenosine monophosphate (AMP2?), adenosine diphosphate (ADP2?), and adenosine triphosphate (ATP2?). The Lamino sample shows an increase in the absorbance at λ=315 nm in the presence of H2PO4?, CO32?, HCO3?, CH3CO2?, and F?. The 1H NMR spectroscopic titration of Lamino with H2PO4?, F?, and CH3CO2? showed major changes in the phenylene‐capped and salicyl moieties, and thereby, confirming the aminophenolic region as the binding core. However, the binding strength of these anions followed the trend H2PO4?>F??CH3CO2?>HSO4?. The heat changes observed by isothermal titration calorimetry support this trend. The Lamino sample showed reversible sensing towards H2PO4? and F? in the presence of Mg2+ and Ca2+, respectively. NOESY studies of Lamino , in comparison with its anionic complexes, revealed that major conformational changes occurred in the capping region to facilitate the binding of anion. ESI‐MS and the Job's method revealed 1:1 stoichiometry between Lamino and H2PO4? or F?. In the SEM micrographs of Lamino , the spherical particles are converted into spherical aggregates and further form large agglomerates and even branched sheets in the presence of anions, depending upon their binding strength.  相似文献   

    19.
    The permeability of rubber and PE gloves, PE bags and strippable PVC paint to HTO, Na125I, H2 35SO4, H3 32PO4,141CeCl3,89SrCl2, and,137CsCl in aqueous solution was tested. All materials were permeable to HTO and Na125I. Whereas rubber gloves were permeable the other materials were impermeable to H2 35SO4. The remaining radioactive compounds showed no penetration. The diffusion coefficients of the penetrating compounds were calculated on the basis of Fick's First Law for the respective materials. The permeability to Na125I and HTO decreases in the order rubber, PVC paint, and PE. Penetration increases in the order H2 35SO4, Na125I, and HTO.  相似文献   

    20.
    The 1H NMR chemical shifts of the C(α)? H protons of arylmethyl triphenylphosphonium ions in CD2Cl2 solution strongly depend on the counteranions X?. The values for the benzhydryl derivatives Ph2CH? PPh3+ X?, for example, range from δH=8.25 (X?=Cl?) over 6.23 (X?=BF4?) to 5.72 ppm (X?=BPh4?). Similar, albeit weaker, counterion‐induced shifts are observed for the ortho‐protons of all aryl groups. Concentration‐dependent NMR studies show that the large shifts result from the deshielding of the protons by the anions, which decreases in the order Cl? > Br? ? BF4? > SbF6?. For the less bulky derivatives PhCH2? PPh3+ X?, we also find C? H???Ph interactions between C(α)? H and a phenyl group of the BPh4? anion, which result in upfield NMR chemical shifts of the C(α)? H protons. These interactions could also be observed in crystals of (p‐CF3‐C6H4)CH2? PPh3+ BPh4?. However, the dominant effects causing the counterion‐induced shifts in the NMR spectra are the C? H???X? hydrogen bonds between the phosphonium ion and anions, in particular Cl? or Br?. This observation contradicts earlier interpretations which assigned these shifts predominantly to the ring current of the BPh4? anions. The concentration dependence of the 1H NMR chemical shifts allowed us to determine the dissociation constants of the phosphonium salts in CD2Cl2 solution. The cation–anion interactions increase with the acidity of the C(α)? H protons and the basicity of the anion. The existence of C? H???X? hydrogen bonds between the cations and anions is confirmed by quantum chemical calculations of the ion pair structures, as well as by X‐ray analyses of the crystals. The IR spectra of the Cl? and Br? salts in CD2Cl2 solution show strong red‐shifts of the C? H stretch bands. The C? H stretch bands of the tetrafluoroborate salt PhCH2? PPh3+ BF4? in CD2Cl2, however, show a blue‐shift compared to the corresponding BPh4? salt.  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