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.
Mixed Dicyanamido (thio) cyanato-cobaltates(II) Preparation and properties of mixed anionic pseudohalide-complexes of cobalt(II) [CoX2Y2]2? and [CoX3Y]2? (X = NCS, NCO, Y = N(CN)2) are reported. The structures of the complexes are discussed using the results of infrared and electronic spectroscopy and of magnetic measurements.  相似文献   

2.
The synthesis, structure, and magnetic properties of three clathrate derivatives of the spin‐crossover porous coordination polymer {Fe(pyrazine)[Pt(CN)4]} ( 1 ) with five‐membered aromatic molecules furan, pyrrole, and thiophene is reported. The three derivatives have a cooperative spin‐crossover transition with hysteresis loops 14–29 K wide and average critical temperatures Tc=201 K ( 1?fur ), 167 K ( 1?pyr ), and 114.6 K ( 1?thio ) well below that of the parent compound 1 (Tc=295 K), confirming stabilization of the HS state. The transition is complete and takes place in two steps for 1?fur , while 1?pyr and 1?thio show 50 % spin transition. For 1?fur the transformation between the HS and IS (middle of the plateau) phases occurs concomitantly with a crystallographic phase transition between the tetragonal space groups P4/mmm and I4/mmm, respectively. The latter space group is retained in the subsequent transformation involving the IS and the LS phases. 1?pyr and 1?thio display the tetragonal P4/mmm and orthorhombic Fmmm space groups, respectively, in both HS and IM phases. Periodic calculations using density functional methods for 1?fur , 1?pyr , 1?thio , and previously reported derivatives 1?CS2 , 1?I, 1?bz (benzene), and 1?pz (pyrazine) have been carried out to investigate the electronic structure and nature of the host–guest interactions as well as their relationship with the changes in the LS–HS transition temperatures of 1?Guest . Geometry‐optimized lattice parameters and bond distances in the empty host 1 and 1?Guest clathrates are in general agreement with the X‐ray diffraction data. The concordance between the theoretical results and the experimental data also comprises the guest molecule orientation inside the host and intermolecular distances. Furthermore, a general correlation between experimental Tc and calculated LS–HS electronic energy gap was observed. Finally, specific host–guest interactions were studied through interaction energy calculations and crystal orbital displacement (COD) curve analysis.  相似文献   

3.
Ruthenium(II) Phthalocyaninates(2–): Synthesis and Properties of (Acido)(carbonyl)phthalocyaninato(2–)ruthenate(II), [Ru(X)(CO)Pc2?]? (X = Cl, Br, I, NCO, NCS, N3) (nBu4N)[Ru(OH)2Pc2?] is reduced in acetone with carbonmonoxid to blue-violet [Ru(H2O)(CO)Pc2?], which yields in tetrahydrofurane with excess (nBu4N)X acido(carbonyl)phthalocyaninato(2–)ruthenate(II), [Ru(X)(CO)Pc2?]? (X = Cl, Br, I, NCO, NCS, N3) isolated as red-violet, diamagnetic (nBu4N) complex salt. The UV-Vis spectra are dominated by the typical π-π* transitions of the Pc2? ligand at approximately 15100 (B), 28300 (Q1) und 33500 cm?1 (Q2), only fairly dependent of the axial ligands. v(C? O) is observed at 1927 (X = I), 1930 (Cl, Br), 1936 (N3, NCO) 1948 cm?1 (NCS), v(C? N) at 2208 cm?1 (NCO), 2093 cm?1 (NCS) and v(N? N) at 2030 cm?1 only in the MIR spectrum. v(Ru? C) coincides in the FIR spectrum with a deformation vibration of the Pc ligand, but is detected in the resonance Raman(RR) spectrum at 516 (X = Cl), 512 (Br), 510 (N3), 504 (I), 499 (NCO), 498 cm?1 (NCS). v(Ru? X) is observed in the FIR spectrum at 257 (X = Cl), 191 (Br), 166 (I), 349 (N3), 336 (NCO) and 224 cm?1 (NCS). Only v(Ru? I) is RR-enhanced.  相似文献   

4.
Quantum-chemical calculations were performed on the mechanisms of reaction of NCN with NO and NS. Possible mechanisms were classified according to four pathways yielding products in the following four possible groups: N2O/N2S + CN, N2 + NCO/NCS, N2 + CNO/CNS, and CNN + NO/NS, labeled in order from p1/p1s to p4/p4s. The local structures, transition structures, and potential-energy surfaces with respect to the reaction coordinates are calculated, and the barriers are compared. In the NCN + NO reaction, out of several adduct structures, only the nitroso adduct NCNNO lies lower in energy than the reactants, by 21.89 kcal/mol; that adduct undergoes rapid transformation into the products, in agreement with experimental observation. For the NS counterpart, both thionitroso NCNNS and thiazyl NCNSN adducts have energies much lower than those of the reactants, by 43 and 29 kcal/mol, respectively, and a five-membered-ring NCNNS (having an energy lower than those of the reactants by 36 kcal/mol) acts as a bridge in connecting these two adducts. The net energy barriers leading to product channels other than p4s are negative for the NS reaction, whereas those for the NO analogue are all positive. The channel leading to p1 (N2O + CN) has the lowest energy (3.81 kcal/mol), whereas the channels leading to p2 (N2 + NCO) and p2s (N2 + NCS) are the most exothermic (100.94 and 107.38 kcal/mol, respectively).  相似文献   

5.
Direct nucleophilic displacement of iodine to give (Me3Si)3 CSiMe2 Y, where Y = F, NCO, NCS, CN or N3, takes place when (Me3Si)3 CSiMe2I is treated with solutions of CsF, KOCN, KSCN, KCN, or NaN3 in MeOH or CH3 CN. The order of effectiveness of the nucleophiles appears to be N3 > F > CN > NCS > NCO in MeOH and NCS > NCO > CN, F in CH3 CN.  相似文献   

6.
The preparation of the new silanes HSi(NCO)3 and HSi(NCS)3 from trihalosilanes and silver halogenoides is described. Based on vibrational and nmr spectroscopic arguments the inductive effect of the NCO- and NCS-groups is discussed. The mixed compounds HSiX 2(NCO) (X=Cl, Br, I), HSiX(NCO)2 (X=Cl, Br, I), H2SiCl(NCO), HSiX 2(NCS) (X=Cl, Br) and HSiCl(NCS)2 are prepared by halogen/halogenoid exchange reactions and are characterized spectroscopically.  相似文献   

7.
The hitherto unknown germanium(II) pseudohalides: Ge(CN)2, Ge(NCO)2 and Ge(NCS)2, have been prepared by reactions of germanium(II) halides with corresponding silver or potassium salts; they are stable in tetrahydrofuran or acetone solution in which they are extremely sensitive to moisture, and they undergo cycloaddition, insertion, and Lewis acid-base reactions characteristic of highly reactive germylenes. Infrared spectra indicate that in tetrahydrofuran solution Ge(CN)2 is the normal cyanide, whereas Ge(NCO)2 and Ge(NCS)2 are the isocyanate and isothiocyanate, respectively.  相似文献   

8.
Three Ni(II) complexes of cresol-based Schiff-base ligands, namely [Ni2(L1)(NCS)3(H2O)2], (1) [Ni2(L2)(CH3COO)(NCS)2(H2O)] (2) and [Ni2(L3)(NCS)3] (3), (where L1 = 2,6-bis(N-ethylpyrrolidineiminomethyl)-4-methylphenolato, L2 = 2,6-bis(N-ethylpiperidineiminomethyl)-4-methylphenolato and L3 = 2,6-bis{N-ethyl-N-(3-hydroxypropyl iminomethyl)}-4-methylphenolato), have been synthesized and structurally characterized by X-ray single-crystal diffraction in addition to routine physicochemical techniques. Density functional theory calculations have been performed to understand the nature of the electronic spectra of the complexes. Complexes 1?C3 when reacted with 4-nitrophenyl phosphate in 50:50 acetonitrile?Cwater medium promote the cleavage of the O?CP bond to form p-nitrophenol and smoothly convert 3,5-di-tert-butylcatechol (3,5-DTBC) to 3,5-di-tert-butylquinone (3,5-DTBQ) either in MeOH or in MeCN medium. Phosphatase- and catecholase-like activities were monitored by UV?Cvis spectrophotometry and the Michaelis?CMenten equation was applied to rationalize all the kinetic parameters. Upon treatment with urea, complexes 1 and 2 give rise to [Ni2(L1)(NCS)2(NCO)(H2O)2] (1??) and [Ni2(L2)(CH3COO)(NCO)(NCS)(H2O)] (2??) derivatives, respectively, whereas 3 remains unaltered under same reaction conditions.  相似文献   

9.
High Spin Manganese(II) Phthalocyanines: Preparation and Spectroscopical Properties of Acidophthalocyaninatomanganate(II) Acidophthalocyaninatomanaganese(III) is reduced by boranate, thioacetate or hydrogensulfide to yield acidophthalo-cyaninatomanganate(II) ([Mn(X)Pc2?]?; X = Cl, Br, NCO, NCS) being isolated as tetra(n-butyl)ammonium salt. In the cyclovoltammogram of [Mn(NCO)Pc2?]? the halv-wave potential for the redoxcouple MnII/MnIII is at ?0.13 V, that of the first ring reduction at ?0.99 V. The magnetic moments are indicative of high-spin 6A1 ground states: μMn = 5.84 (NCO), 5.78(Cl), 5.65 (Br), 5.68 μB (NCS). A Curie-like temperature dependence of μMn is observed in the region 300–30 K. Below 30 K an increase in μMn occurs due to weak intermolecular ferromagnetic coupling. The ESR spectra confirm the S = 5/2 ground state with a strong g = 6 resonance observed (AMn = 80 G) as expected for an axially distorted ligand-field. Besides the typical π-π* transitions of the Pc2?-ligand several weak bands are observed in the Uv-vis-n.i.r. spectra at ca. 7.5, 9.1, 14.0 and 19.0 kK that are assigned to trip-multiplet transitions. In resonance with the band at 19.0 kK the Mn? X stretching vibration (v(MnX)) is resonance Raman enhanced: X = NCO: 319, Cl: 286, SCN: 238, Br: 202 cm?1. These vibrational frequencies are confirmed by the f.i.r. spectra. In the case of the thiocyanato-complex probably both forms of bonding of the ambident NCS-ligand are present (v(Mn? NCS): 274 cm?1). The frequencies of the vibrations of the inner (CN)8 ring are reduced by up to 20 cm?1 as compared with those of low spin MnII phthalocyanines.  相似文献   

10.
The substituent effect of electron‐withdrawing groups on electron affinity and gas‐phase basicity has been investigated for substituted propynl radicals and their corresponding anions. It is shown that when a hydrogen of the α‐CH3 group in the propynyl system is substituted by an electron‐withdrawing substituent, electron affinity increases, whereas gas‐phase basicity decreases. These results can be explained in terms of the natural atomic charge of the terminal acetylene carbon of the systems. The calculated electron affinities are 3.28 eV (?C?C? CH2F), 3.59 eV (?C?C? CH2Cl) and 3.73 eV (?C?C? CH2Br), and the gas‐phase basicities of their anions are 359.5 kcal/mol (?:C?C? CH2F), 354.8 kcal/mol (:C?C? CH2Cl) and 351.3 kcal/mol (?:C?C? CH2Br). It is concluded that the larger the magnitude of electron‐withdrawing, the greater is the electron affinity of radical and the smaller is the gas‐phase basicity of its anion. © 2009 Wiley Periodicals, Inc. J Comput Chem, 2009  相似文献   

11.
《Chemical physics》1987,117(2):315-324
Photoabsorption cross sections and fluorescence excitation spectra of CH3NCO, CH3NCS and CH3SCN vapor were measured in the vacuum ultraviolet using synchrotron radiation. Many sharp structures observed from CH3NCO and CH3SCN in the 120–180 nm region are classified into three Rydberg series and their vibrational progressions, whereas for CH3NCS six broad bands exhibit no fine structure. The emission which starts to appear at 172.8 ± 1.0 nm excitation of CH3NCO is attributed to the NCO(A2Σ+-X2Π) band. The emissions from CH3NCS and CH3SCN are assigned to the A2Π-X2Π and B2Σ+-X2Π bands of NCS; the CN(B2Σ+-X2Σ+) band is also observed at 125 nm excitation of CH3SCN. The photodissociation processes are discussed in accord with the emission observed.  相似文献   

12.
The λ6Si‐silicate [K(18‐crown‐6)]2[Si(NCO)6] ( 10 ) was synthesized by treatment of Si(NCO)4 with KNCO in the presence of 18‐crown‐6. Compound 10 (SiN6 skeleton) is the first example of a hexa(cyanato‐N)silicate. It was characterized by solid‐state and solution NMR spectroscopy, and the acetonitrile solvate 10· 2CH3CN was studied by single‐crystal X‐ray diffraction. To differentiate between the two isomeric [Si(NCO)6]2? and [Si(OCN)6]2? dianions, computational studies were performed.  相似文献   

13.
Several stable new compounds of transition metals with C6Cl5 as ligand, of the type [MX(C6Cl5)(PPh3)2] (M  Pd and X  Cl, Br, I, NCS, NCO, N3; M  Ni and X  NCS, NCO, N3) are described. Their preparation in some cases required in situ reaction of [MX2(PPh3)2], Mg and C6Cl6. The action of gaseous HCl and Cl2 on the solutions of these compounds has been examined. The IR spectra of the cyanato and thiocyanato complexes indicates coordination through the nitrogen atom.  相似文献   

14.
A series of NCO/NCS pincer precursors, 3‐(Ar2OCH2)‐2‐Br‐(Ar1N?CH)C6H3 ((Ar1NCOAr2)Br, 3a , 3b , 3c , 3d ) and 3‐(2,6‐Me2C6H3SCH2)‐2‐Br‐(Ar1N?CH)C6H3 ((Ar1NCSMe)Br, 4a and 4b ) were synthesized and characterized. The reactions of [Ar1NCOAr2]Br/ [Ar1NCSMe]Br with nBuLi and the subsequent addition of the rare‐earth‐metal chlorides afforded their corresponding rare‐earth‐metal–pincer complexes, that is, [(Ar1NCOAr2)YCl2(thf)2] ( 5a , 5b , 5c , 5d ), [(Ar1NCOAr2)LuCl2(thf)2] ( 6a , 6d ), [(Ar1NCOAr2)GdCl2(thf)2] ( 7 ), [{(Ar1NCSMe)Y(μ‐Cl)}2{(μ‐Cl)Li(thf)2(μ‐Cl)}2] ( 8 , 9 ), and [{(Ar1NCSMe)Gd(μ‐Cl)}2{(μ‐Cl)Li(thf)2(μ‐Cl)}2] ( 10 , 11 ). These diamagnetic complexes were characterized by 1H and 13C NMR spectroscopy and the molecular structures of compounds 5a , 6a , 7 , and 10 were well‐established by X‐ray diffraction analysis. In compounds 5a , 6a , and 7 , all of the metal centers adopted distorted pentagonal bipyramidal geometries with the NCO donors and two oxygen atoms from the coordinated THF molecules in equatorial positions and the two chlorine atoms in apical positions. Complex 10 is a dimer in which the two equal moieties are linked by two chlorine atoms and two Cl? Li? Cl bridges. In each part, the gadolinium atom adopts a distorted pentagonal bipyramidal geometry. Activated with alkylaluminum and borate, the gadolinium and yttrium complexes showed various activities towards the polymerization of isoprene, thereby affording highly cis‐1,4‐selective polyisoprene, whilst the NCO? lutetium complexes were inert under the same conditions.  相似文献   

15.
The new diiodine basicity scale pKBI2 is quasi‐orthogonal to most known Lewis basicity scales (hydrogen‐bond, dative‐bond and cation basicity scales). The diiodine basicity falls in the sequence N>P≈Se>S>I≈O>Br>Cl>F for the iodine‐bond acceptor atomic site and SbO≈NO≈AsO>SeO>PO>SO>C?O>? O? >SO2 or PS?? S? >C?S?N?C?S for the functionality of oxygen or sulfur bases. Substituent effects are quantified through linear free energy relationships, which allow the calculation of individual complexation constants for each site of polybases and thus the classification of aromatic ethers as carbon π bases and of aromatic amines, thioethers and selenoethers as N, S and Se bases, respectively. The pKBI2 values of nBu3N+‐N?C≡N, 2‐aminopyridine and 1,10‐phenanthroline reveal a superbasic nitrile, a hydrogen‐bond‐assisted iodine bond and a two‐centre iodine bond, respectively. The diiodine basicity scale is a general inorganic but family‐dependent organic halogen‐bond basicity scale because organic halogen‐bond donors such as IC≡N and ICF3 have a stronger electrostatic character than I2. The family independence can be restored by the addition of an electrostatic parameter, either the experimental pKBHX hydrogen‐bond basicity scale or the computed minimum electrostatic potential.  相似文献   

16.
Mean amplitudes and shrinkages are calculated from spectroscopic data for SiH3NCO, SiD3NCO, SiH3NCS and SiD3NCS. The lowest frequency (ν10) in each of the molecules is uncertain, and its influence on the calculated amplitude parameters is investigated. The results are compared with electron diffraction studies.  相似文献   

17.
We have already shown that the in-vacuum gas-phase Meerwein reaction of (thio)acylium ions is general in nature and useful for class-selective screening of cyclic (thio)epoxides. Herein we report that this gas-phase reaction can also be performed efficiently at atmospheric pressure under both electrospray ionization (ESI) and atmospheric pressure chemical ionization (APCI) conditions. This alternative expands the range of molecules that can be reacted by gas-phase Meerwein reaction. Phenyl epoxide, thiirane, 3-methoxy-2,2-dimethyloxirane, propylene oxide, 2,2'-bioxirane, trans-1,3-diphenyl-2,3-epoxypropan-1-one, epichloridrine and propylene oxide are shown to react efficiently in both ESI and APCI conditions. Tetramethylurea (TMU) and (thio)TMU were both used as dopants, being co-injected with either toluene, acetonitrile or methanol solutions of the (thio)epoxides, with similar results. In both ESI and APCI, (thio)TMU is protonated preferentially, and these labile species dissociate promptly to yield (CH3)2N-C+=O and (CH3)2NCS+, which are the least acidic and most reactive (thio)acylium ions so far tested in the gas-phase Meerwein reaction. Under the low-energy ESI conditions set to favor both the formation of the (thio)acylium ion and ion/molecule reactions, (CH3)2NCO(S)+ react competitively with (thio)TMU to form acylated (thio)TMU and with the (thio)epoxide to form the characteristic Meerwein products. Enhanced selectivity in structural characterization or for the screening of (thio)epoxides is achieved by performing on-line collision-induced dissociation of Meerwein products, particularly for the more structurally complex (thio)epoxides.  相似文献   

18.
The addition of water to 1-(4-Z-2,6-dinitrophenyl)-3-methylimidazolium chloride (Z = CN, NO2, and CF3) is catalyzed by carboxylate bases. There is a decrease in ρ with increasing basicity of the catalyst with ?ρ/?pKBH = ?0.090. The results indicate that the mechanism of the reaction involves the addition of water to the aromatic substrates catalyzed by general bases. The changes in the structure-reactivity parameters with changing pK of the catalysts or Z in the substrate are rationalized in terms of changing bond lengths in the transition state. The activation parameters and the kinetic solvent isotopic effect were determined for the water, acetate, and OH? catalyzed reactions and are consistent with the proposed mechanism.  相似文献   

19.
The NCO + C2H4 reaction is simple and prototype for reaction of the NCO radical with unsaturated hydrocarbons, and it is considered to be important in fuel‐rich combustion. In this article, we for the first time perform detailed theoretical investigations for its reaction mechanism based on Gaussian‐3//B3LYP scheme covering various entrance and decomposition channels. The most favorable channel is firstly the NCO and C2H4 approach each other, forming a weakly‐bound complex L1 OCN···C2H4, followed by the formation of isomer L2 OCNCH2CH2 via a small barrier of 1.3 kcal/mol. Transition states of any decomposable or isomeric channels for L2 in energy are much higher than reactants, which indicate that adduct L2 has stabilization effect in this NCO + C2H4 reaction. The direct H‐abstraction channel leading to P1 HNCO + C2H3, might have an important contribution to the eventual products in high temperature. These results can well explain available kinetic experiment. Moreover, reaction mechanism for the title reaction is significantly different from the NCO + C2H2 reaction which proceeds on most favorably to generate the products HCN + HCCO and OCCHCN + H via a four‐membered ring intermediate. © 2008 Wiley Periodicals, Inc. Int J Quantum Chem, 2009  相似文献   

20.
A UiO‐66‐NCS MOF was formed by postsynthetic modification of UiO‐66‐NH2. The UiO‐66‐NCS MOFs displays a circa 20‐fold increase in activity against the chemical warfare agent simulant dimethyl‐4‐nitrophenyl phosphate (DMNP) compared to UiO‐66‐NH2, making it the most active MOF materials using a validated high‐throughput screening. The ?NCS functional groups provide reactive handles for postsynthetic polymerization of the MOFs into functional materials. These MOFs can be tethered to amine‐terminated polypropylene polymers (Jeffamines) through a facile room‐temperature synthesis with no byproducts. The MOFs are then crosslinked into a MOF–polythiourea (MOF–PTU) composite material, maintaining the catalytic properties of the MOF and the flexibility of the polymer. This MOF–PTU hybrid material was spray‐coated onto Nyco textile fibers, displaying excellent adhesion to the fiber surface. The spray‐coated fibers were screened for the degradation of DMNP and showed durable catalytic reactivity.  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