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 15 毫秒
1.
The selectivity of the cryptand [TriPip222], a per‐aza analogue of cryptand [2.2.2], in which each of the linking arms contains a piperazine ring for the endohedral complexation of metal cations of the I, II, and III main groups and group 12 of the periodic table of elements, was predicted on the basis of DFT [B3LYP/LANL2DZp (LANL2DZp = LANL2DZ augmented with polarization functions on non‐hydrogen atoms)] calculated structures and complex‐formation energies. The cavity size of the studied cryptand is similar to that of [bpy.bpy.bpy], [2.bpy.bpy] and [2.phen.phen], such that the complexation of K+ > Na+ and of Sr2+ ≈ Ca2+ > Ba2+ are most favorable. The essential flexibility for achieving the selectivity of the cryptand is mainly associated with a twist of the CH2–Nbridgehead ··· Nbridgehead–CH2 angle and not with the piperazine moiety.  相似文献   

2.
The synthesis and characterization of two bimetallic, cationic low‐valent gallium–cryptand[2.2.2] complexes is reported. The reaction of cryptand[2.2.2] with Ga2Cl4 gave two different cations, [Ga3Cl4(crypt‐222)]+ ( 1 ) or [Ga2Cl2(crypt‐222)]2+ ( 2 ), depending on whether or not trimethylsilyl triflate (Me3SiOTf) was added as a co‐reagent. Complexes 1 and 2 are the first examples of bimetallic cryptand[2.2.2] complexes, as well as the first low‐valent gallium–cryptand[2.2.2] complexes. Computational methods were used to evaluate the bonding in the gallium cores.  相似文献   

3.
Under Ammonia chemical Ionization conditions the source decompositions of [M + NH4]+ ions formed from epimeric tertiary steroid alchols 14 OHβ, 17OHα or 17 OHβ substituted at position 17 have been studied. They give rise to formation of [M + NH4? H2O]+ dentoed as [MHsH]+, [MsH? H2O]+, [MsH? NH3]+ and [MsH? NH3? H2O]+ ions. Stereochemical effects are observed in the ratios [MsH? H2O]+/[MsH? NH3]+. These effects are significant among metastable ions. In particular, only the [MsH]+ ions produced from trans-diol isomers lose a water molecule. The favoured loss of water can be accounted for by an SN2 mechanism in which the insertion of NH3 gives [MsH]+ with Walden inversion occurring during the ion-molecule reaction between [M + NH4]+ + NH3. The SN1 and SNi pathways have been rejected.  相似文献   

4.
A first principles methodology, aimed at understanding the roles of molecular conformation and energetics in host–guest binding interactions, is developed and tested on a system that pushes the practical limits of ab initio methods. The binding behavior between the [2.2.2]‐cryptand host (4,7,13,16,21,24‐hexaoxa‐1,10‐diaza‐bicyclo[8.8.8]hexacosane) and alkali metal cations (Li+, Na+, and K+) in gas, water, methanol, and acetonitrile is characterized. Hartree–Fock and density functional theory methods are used in concert with crystallographic information to identify gas phase, energy‐minimized conformations. Gas phase free energies of binding, with vibrational contributions, are compared to solution‐state binding constants through relative binding selectivity analysis. Calculated relative binding free energies qualitatively correlated with solution state experiments only after gas phase metal desolvation is considered. The B3LYP exchange–correlation functional improves theoretical correlations with experimental relative binding free energies. The relevance of gas phase calculations towards understanding binding in condensed phases is discussed. Natural bond orbital methods highlights previously unidentified intramolecular and intermolecular M+(222) chemistries, such as an intramolecular n′O→σ*CH hydrogen bond.  相似文献   

5.
Excess‐electron compounds can be considered as novel candidates for nonlinear optical (NLO) materials because of their large static first hyperpolarizabilities (β0). A room‐temperature‐stable, excess‐electron compound, that is, the organic electride Na@(TriPip222), was successfully synthesized by the Dye group (J. Am. Chem. Soc. 2005 , 127, 12416). In this work, the β0 of this electride was first evaluated to be 1.13×106 au, which revealed its potential as a high‐performance NLO material. In particular, the substituent effects of different substituents on the structure, electride character, and NLO response of this electride were systemically studied for the first time by density functional theory calculations. The results revealed that the β0 of Na@(TriPip222) could be further increased to 8.30×106 au by introducing a fluoro substituent, whereas its NLO response completely disappeared if one nitryl group was introduced because the nitro‐group substitution deprived the material of its electride identity. Moreover, herein the dependence of the NLO properties on the number of substituents and their relative positions was also detected in multifluoro‐substituted Na@(TriPip222) compounds.  相似文献   

6.
Thermodynamic quantities for the interactions of mono- and tri(2-methylenepropylene)-bridged cryptands, cryptand [3.3.1], cryptand [2.2.2], and 18-crown-6-with Na+, K+, Rb+, and Cs+ have been determined by calorimetric titration in an 80:20 (v/v) methanol: water solution at 25°C. Incorporation of the 2-methylenepropylene (–CH2C(=CH2)CH2–) bridge(s) into cryptand [2.2.2] results in a large change in the ligand-cation binding properties. Tri(2-methylenepropylene)-bridged cryptand [2.2.2] (2) shows high selectivity factors for Na+ over K+ and other alkali cations, while 2-methylenepropylene-bridged cryptand [2.2.2.] (1) selects K+ over Na+, as does cryptand [2.2.2]. The K+/Na+ selectivity is reversed with increasing number of 2-methylenepropylene bridges. This observation indicates that increasing the number of 2-methylenepropylene bridges on cryptand [2.2.2] favors complexation of a small cation over a large one. The logK values for the formation of 1 and 2 complexes (except 1-Cs+ and 2-Na+) decrease as compared with those for the corresponding [2.2.2] complexes. Formation of six-membered chelate ring(s) by the propyleneoxy unit(s) of 1 and 2 with a cation stabilizes the cryptate complexes of the small Na+ and destabilizes the complexes of large alkali metal cations. Thermodynamic data indicate that the stabilities of the cryptate complexes studied are dominated mostly by the enthalpy change. In most cases, both stabilization of Na+ complexes and destabilization of the complexes of large alkali metal cations by six-membered chelate ring(s) also result from an enthalpic effect. Cryptand [3.3.1] shows a selectivity for K+ over Cs+, despite its two long CH2(CH2OCH2)3CH2 bridges. The [3.1] macroring portion of [3.3.1]may be too small to effectively bind the Cs+, resulting in the low stability of the Cs+ complex.  相似文献   

7.
A non‐ionic cryptand‐22 surfactant consisting of a macrocyclic cryptand‐22 polar head and a long paraffinic chain (C10H21‐Cryptand‐22) was synthesized and characterized. The critical micellar concentration (CMC) of the cryptand surfactant in ROH/H2O mixed solvent was determined by the pyrene fluorescence probe method. In general, the cmc of the cryptand surfactant increased upon decreasing the polarity of the surfactant solution. The cryptand surfactant also can behave as a pseudo cationic surfactant by protonation of cryptand‐22 or complexation with metal ions. Effects of protonation and metal ions on the cmc of the cryptand surfactant were investigated. A preliminary application of the cryptand surfactant as an ion‐transport carrier for metal ions, e.g., Li+, Na+, K+ and Sr2+, through an organic liquid‐membrane was studied. The transport ability of the cryptand surfactant for these metal ions was in the order: K+ ≥ Na+ < Li+ < Sr2+. A comparison of the ion‐transport ability of the cryptand surfactant with other macrocyclic polyethers, e.g., dibenzo‐18‐crown‐6, 18‐crown‐6 and benzo‐15‐crown‐5, was studied and discussed. Among these macrocyclic polyethers, the cryptand surfactant was the best ion‐transport carrier for Na+, Li+ and Sr2+ ions. Furthermore, a foam extraction system using the cryptand surfactant to extract the cupric ion was also investigated.  相似文献   

8.
Redox‐inactive metal ions are one of the most important co‐factors involved in dioxygen activation and formation reactions by metalloenzymes. In this study, we have shown that the logarithm of the rate constants of electron‐transfer and C−H bond activation reactions by nonheme iron(III)–peroxo complexes binding redox‐inactive metal ions, [(TMC)FeIII(O2)]+‐Mn + (Mn +=Sc3+, Y3+, Lu3+, and La3+), increases linearly with the increase of the Lewis acidity of the redox‐inactive metal ions (ΔE ), which is determined from the gzz values of EPR spectra of O2.−‐Mn + complexes. In contrast, the logarithm of the rate constants of the [(TMC)FeIII(O2)]+‐Mn + complexes in nucleophilic reactions with aldehydes decreases linearly as the ΔE value increases. Thus, the Lewis acidity of the redox‐inactive metal ions bound to the mononuclear nonheme iron(III)–peroxo complex modulates the reactivity of the [(TMC)FeIII(O2)]+‐Mn + complexes in electron‐transfer, electrophilic, and nucleophilic reactions.  相似文献   

9.
The cryptate electrode (Ag/Ag+222), prepared by immersing silver wire in a solution of silver(I) salt and the cryptand 222 (4,7,13,16,21,24‐hexaoxa‐1,10‐diazabicyclo[8.8.8]hexacosane) in ionic liquids have been studied. The potential of the electrode is stabilized by the equilibrium of the Ag+ ion complexation by the cryptand, similarly to the potential stabilization by the ionic product of slightly soluble salts, used in aqueous electrodes of the second kind. The Ag/Ag+222 cryptate electrode (concentration of the cryptate was much higher than the silver(I) cation concentration, [222]>[Ag+]) may be used as a reference electrode in room temperature ionic liquids. The potential of the Ag/Ag+222 electrode is less sensitive to the presence of impurities, such as halides or water, in comparison to the Ag/Ag+ electrode. After anodic or cathodic polarization, the potential of the Ag/Ag+222 electrode comes back to the initial open circuit potential quickly. Preparation of the Ag/Ag+222 reference electrode is very easy: a silver wire is immersed in a solution of Ag+ salt and cryptand 222 (both available commercially) in the ionic liquid under study.  相似文献   

10.
Cation fluxes from binary mixtures of either Na+, Cs+ or Sr2+ with other alkali metal cations, alkaline earth metal cations, and Pb2+ through a H2OCHCl3H2O bulk liquid membrane system containing one of several macrocyclic carriers have been determined Nitrate salts were used in all cases. The most selective transport of Na+ over all other cations studied was found with the carrier cryptand [2.2.1]. Selective transport of Na+ relative to Li+, Cs+ and the alkaline earth cations was found with cryptand [2.2.2B] and cryptand [2.2.2D]. The ligands 21-crown-7 and dibenzo-24-crown-8 showed selective transport of Cs+ over the second cation in all cases. Several macrocycles showed selectivity for Sr2+ over the second cation with the macrocycle 1,10-diaza-18-crown-6 showing the highest selectivity for this cation of all ligands studied. Relative fluxes from binary cation mixtures are rationalized in terms of macrocycle cavity size, donor atom type and ring substituents.  相似文献   

11.
本文利用超微铂电极和循环伏安法研究了在碱金属碘化物与冠醚或穴醚配合物的3-甲氧基丙腈(MePN)溶液中I3-和I-的氧化还原行为。发现I3-和I-在其中的表观扩散系数与阳离子有关,且I3-的表观扩散系数符合以下规律:1,2-二甲基-3-丙基咪唑阳离子(DMPI+)> [Na(¯¯15-C-5]+ > [K(¯¯18-C-6]+ > [Na(¯¯2.2.1-cryptand]+,I-的表观扩散系数则为:[Na(¯¯2.2.1-cryptand]+> [Na(¯¯15-C-5]+ ≈[K(¯¯18-C-6]+> DMPI+。比较了由上述配合物和1,2-二甲基-3-丙基咪唑碘(DMPII)组成的染料敏化纳米薄膜太阳电池(DSC)的光伏性能,结果表明由上述配合物组成的DSC,其短路电流略高于DMPII,填充因子略低于DMPII,这与I-和I3-在其中的表观扩散系数的大小是相一致的。此外,电解质溶液中的溶剂对DSC的光电转换效率也有较大影响,以MePN为溶剂,含DMPII的DSC的光电转换效率要高于[K(¯¯18-C-6]I,而以乙腈为溶剂,两者的光电转换效率并没有明显的差别。  相似文献   

12.
For the calculation of the atomic or ionic volumes the Quantum Theory of Atoms In Molecules method was applied. The regions (basins) around the nuclei confined by the zero‐flux surfaces in the electron density gradient are called QTAIM atoms. They are non‐overlapping and completely fill the space. The volume of the basins gives volumes of atoms or ions. The integration of the electron density within the volumina yields effective charges, defining neutral or ionic character of the given QTAIM species. Present investigations refer to metal hydrides, metal nitrides and to intermetallic compounds of the system Al‐Pt. A linear relation between the ionic volumina of hydrogen or nitrogen established according to QTAIM and after Biltz has been found with (homodesmic) binary metal hydrides and binary metal nitrides, but has been observed merely as a trend with stronger deviations for heterodesmic compounds, such as ternary hydrido‐ and nitridometallates Aa[MmXx] (A – alkali or alkaline earth metal, M – transition metal and X – H or N). The deviation from linearity for heterodesmic compounds is caused by the different kinds of chemical bonds being present within the [MmXx] anions on the one hand and between the anions and the cations on the other hand reflected by the calculated volumes and the QTAIM charges of M and X components. Concerning the intermetallic compounds of the system Al‐Pt, the quantum chemical calculations reveal negative charges for the platinum atoms and positive ones for the aluminium atoms in accordance with their electronegativities. Introducing the variation of the atomic volume with the composition extends the Vegard's approach and gives a non‐linear slope for the concentration dependence of mean atomic volume which explains qualitatively the experimental results.  相似文献   

13.
Two new highly selective triiodide electrodes have been prepared using charge‐transfer complex of iodine with cryptand 222 as an electroactive ionophore and nitrophenyl octyl ether as a plasticizing agent. The electrodes showed Nernstian response to triiodide ions over a concentration range from 1.0 × 10?;2 — 7.9 × 10?;7 M and from 1.0 × 10?;2 — 1 × 10?;6 M with detection limits of 6.3 × 10?;7 and 7.9 × 10?;7 M for cryptand and its charge‐transfer complex with iodine, respectively. The response times (t95%) of the sensors were 10 and 5 s. The membrane could be used for more than 1 month without any divergence in potentials. The proposed sensors exhibited very high selectivity for triiodide ion over other anions, and could be used in a wide pH range ?2–10. These electrodes were successfully applied as an indicator electrode in potentiometric titration of copper in ore samples.  相似文献   

14.
Due to disorder, [K(crypt‐222)]+ ( 1 ) is crystallographically indistinguishable from its isomer, a 1:1 fluoroform clathrate of deprotonated [K(crypt‐222)]+, [K(crypt‐222{‐H+})]·CHF3 ( 2 ). In our preceding publications, 1 was characterized on the basis of combined X‐ray diffraction, NMR, reactivity, labeling, acid‐base, and DFT studies. Herein we report that neither incorporation of deuterium nor any other transformation of [K(crypt‐222)]+ is observed in the presence of CD3OD/CD3OK at 70 °C or CD3S(O)CD2K/(D6)DMSO at 23 °C over hours and even days. Since fluoroform is easily and quickly deprotonated under such conditions, the demonstrated greater acidity of CHF3 relative to [K(crypt‐222)]+ provides additional evidence in favor of 1 and against 2 . Likewise, crystal packing analysis and DFT calculations support 1 , which has now been refined in the ultimately correct space group R32. Our previously drawn conclusions regarding [K(crypt‐222)]+ and the existence of a ‘naked’ trifluoromethyl anion in a condensed phase remain valid. The recent alternative refinement (Becker and Müller, Chem. Eur. J. 2017 , 23, 7081 – 7086) of our raw diffraction data has been improperly performed in the wrong space group to yield a non‐charge‐balanced model [K(crypt‐222)]+·CHF3, which is hereby repudiated. Becker and Müller's attempts to resolve the charge balance issue have produced models that are inadequate from the perspective of both crystallography and chemistry.  相似文献   

15.
Fullerence C60‐cryptand 22 was prepared and successfully applied as the electric carrier in the PVC electrode membrane of a bifunctional ion‐selective electrode for cations, e.g., Ag+ ions as well as anions, e.g., I? ions. The bifunctional ion‐selective electrode based on C60‐cryptand 22 can be applied as a Silver (Ag+) ion selective electrode with an internal electrode solution of 10?3 M AgNO3 in water (pH = 6.3), or as an Iodide (I?) ion selective electrode with an acidic internal electrode solution of 10?4 M KI(aq) (pH = 2) in which the cryptand 22 is protonated, and the C60‐cryptand 22 is changed to C60‐Cryptand22–H+ and becomes an anionic electro‐carrier to absorb the I? ion. The Ag+ ion selective electrode based on C60‐cryptand 22 gave a linear response with a near‐Nernstian slope (59.5 mV decade?1) within the concentration range 10?1‐10?3 M Ag+(aq). The Ag+ ion electrode exhibited comparatively good selectivity for silver ions, over other transition‐metal ions, alkali and alkaline earth metal ions. The Ag+ ion selective electrode with good stability and reproducibility was successfully used for the titration of Ag+(aq) with Cl? ions. The Iodide (I?) Ion selective electrode based on protonated C60–cryptand22‐H+ also showed a linear response with a nearly Nernstian slope (58.5 mV decade?1) within 10?1 ‐ 10?3 M I? (aq) and exhibited good selectivity for I? ions and had small selectivity coefficients (10?2–10?3) for most of other anions, e.g., F? , OH?, CH3COO?, SO42?, CO32?, CrO42?, Cr2O72? and PO43? ions.  相似文献   

16.
The analytical capabilities associated with the use of silylation reactions have been extended to a new class of organic molecules, nitroaromatic compounds (NACs). These compounds are a possible contributor to urban particulate matter of secondary origin which would make them important analytes due to their (1) detrimental health effects, (2) potential to affect aerosol optical properties, and (3) and usefulness for identifying PM2.5 from biomass burning. The technique is based on derivatization of the parent NACs by using N,O‐bis‐(trimethylsilyl)‐trifluoro acetamide, one of the most prevalent derivatization reagent for analyzing hydroxylated molecules, followed by gas chromatography‐mass spectrometry using electron ionization (EI) and methane chemical ionization (CI). This method is evaluated for 32 NACs including nitrophenols, methyl‐/methoxy‐nitrophenols, nitrobenzoic acids, and nitrobenzyl alcohols. Electron ionization spectra were characterized by a high abundance of ions corresponding to [M+] or [M+ − 15]. Chemical ionization spectra exhibited high abundance for [M+ + 1], [M+ − 15], and [M+ + 29] ions. Both EI and CI spectra exhibit ions specific to nitro group(s) for [M+ − 31], [M+ − 45], and [M+ − 60]. The strong abundance observed for [M+] (EI), [M+ − 15] (EI/CI), or [M+ + 1] (CI) ions is consistent with the high charge stabilizing ability associated with aromatic compounds. The combination of EI and CI ionization offers strong capabilities for detection and identification of NACs. Spectra associated with NACs, containing hydrogen, carbon, oxygen, and nitrogen atoms only, as silylated derivatives show fragment/adduct ions at either (a) odd or (b) even masses that indicate either (a) odd or (b) even number of nitro groups, respectively. Mass spectra associated with silylated NACs exhibited 3 distinct regions where characteristic fragmentation with a specific pattern associated with (1) ─OH and/or ─COOH groups, (2) ─NO2 group(s), and (3) benzene ring(s). These findings were confirmed with applications to chamber aerosol and ambient PM2.5.  相似文献   

17.
A new ratiometric fluorescent sensor ( 1 ) for Cu2+ based on 4,4‐difluoro‐4‐bora‐3a,4a‐diaza‐s‐indacene (BODIPY) with di(2‐picolyl)amine (DPA) as ion recognition subunit has been synthesized and investigated in this work. The binding abilities of 1 towards different metal ions such as alkali and alkaline earth metal ions (Na+, K+, Mg2+, Ca2+) and other metal ions ( Ba2+, Zn2+, Cd2+, Fe2+, Fe3+, Pb2+, Ni2+, Co2+, Hg2+, Ag+) have been examined by UV‐vis and fluorescence spectroscopies. 1 displays high selectivity for Cu2+ among all test metal ions and a ~10‐fold fluorescence enhancement in I582/I558 upon excitation at visible excitation wavelength. The binding mode of 1 and Cu2+ is a 1:1 stoichiometry determined via studies of Job plot, the nonlinear fitting of the fluorometric titration and ESI mass.  相似文献   

18.
An investigation of competing metastable transitions in the mass spectra of ethylene ketals RSRLC(OCH2)2 (where RL is a larger n-alkyl group than RS) has established that in most cases RS is lost with a lower activation energy than RL. This technique has also been applied to ketones RSRLC?O, to show again that RS is usually lost with the lower activation energy (thus supporting earlier data based on relative daughter ion abundances at the threshold). In the classes of compounds so far investigated, although [M+ ? RS] ions are formed with lower activation energies than [M+ ? RL] ions, the ion yield of [M+ ? RS] ions is anomalously low from ions of high internal energy. Factors which may influence the [M+ ? RS]/[M+ ? RL] ratio of daughter ion intensities are examined. It is suggested that at the threshold [M+ ? RS] and [M+ ? RL] ions may be formed with rearrangement, or from an electronic state that cannot be effectively populated from molecular ions of high internal energies.  相似文献   

19.
Free energies and entropies of transfer from water to methanol have been obtained for [M+18C6] complexes, where M+ = Na+, K+, Rb+, Cs+, and Ag+. The variation of ΔGt° and ΔSt° with the central metal cation is smaller than with the [M+222] complexes and it is concluded that 18-crown-6 shields the metal cation from the solvent more effectively than crystal structure determinations would suggest.  相似文献   

20.
The formation process of M2+ molecular ions sputtered from elementary target materials is investigated. In a previous article it was shown that these molecules can be used to quantitate major elements [1]. The quantitation method was based on the assumption that the M2+ molecular ions are formed by the atomic combination of independently sputtered M and M + particles above the surface. In this paper this assumption will be investigated using a Monte Carlo model to simulate the formation mechanism. The model is used to calculate the velocity distribution of the M2+ dimers sputtered from three different elementary target materials (Fe, Ge, and Ni). The results are compared with experimental data. Good agreement exists between theory and experiment that supports the Monte Carlo model and hence also the assumed formation mechanism.  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