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.
Effective core potential (ECP) and full-electron (FE) calculations for MoS4?2, MoO4?2, and MoOCl4 compounds were analyzed. Geometry parameters, binding energies, charge distributions, and topological properties of the electronic density were studied for Mo? L bonds (L = S, O, Cl). Results clearly indicate that those approaches that include valence plus 4s and 4p electrons (ECP2 methods) are able to reproduce the topological properties of Mo? L bonds, charge distributions, and geometries with respect to those obtained by FE methods. ECP methods that consider only the 4d and 5s valence electrons (ECP1) fail in the calculation of molecular properties. The use of 5p functions in ECP1 approaches produces a negative Mulliken charge on Mo. Bader's charges give more consistent results than Mulliken's ones. A new parameter for measuring the degree of ionicity is proposed. © 1994 by John Wiley & Sons, Inc.  相似文献   

2.
2‐Phenylethanol, racemic 1‐phenyl‐2‐propanol, and 2‐methyl‐1‐phenyl‐2‐propanol have been pyrolyzed in a static system over the temperature range 449.3–490.6°C and pressure range 65–198 torr. The decomposition reactions of these alcohols in seasoned vessels are homogeneous, unimolecular, and follow a first‐order rate law. The Arrhenius equations for the overall decomposition and partial rates of products formation were found as follows: for 2‐phenylethanol, overall rate log k1(s−1)=12.43−228.1 kJ mol−1 (2.303 RT)−1, toluene formation log k1(s−1)=12.97−249.2 kJ mol−1 (2.303 RT)−1, styrene formation log k1(s−1)=12.40−229.2 kJ mol−1(2.303 RT)−1, ethylbenzene formation log k1(s−1)=12.96−253.2 kJ mol−1(2.303 RT)−1; for 1‐phenyl‐2‐propanol, overall rate log k1(s−1)=13.03−233.5 kJ mol−1(2.303 RT)−1, toluene formation log k1(s−1)=13.04−240.1 kJ mol−1(2.303 RT)−1, unsaturated hydrocarbons+indene formation log k1(s−1)=12.19−224.3 kJ mol−1(2.303 RT)−1; for 2‐methyl‐1‐phenyl‐2‐propanol, overall rate log k1(s−1)=12.68−222.1 kJ mol−1(2.303 RT)−1, toluene formation log k1(s−1)=12.65−222.9 kJ mol−1(2.303 RT)−1, phenylpropenes formation log k1(s−1)=12.27−226.2 kJ mol−1(2.303 RT)−1. The overall decomposition rates of the 2‐hydroxyalkylbenzenes show a small but significant increase from primary to tertiary alcohol reactant. Two competitive eliminations are shown by each of the substrates: the dehydration process tends to decrease in relative importance from the primary to the tertiary alcohol substrate, while toluene formation increases. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 401–407, 1999  相似文献   

3.
N-chloralkyl-nitridochloro Complexes of Molybdenum (VI). [Cl5MoN-R] with R = CCl3, C2Cl5. Crystal Structure of (AsPh4)2[(MoOCl4)2CH3CN] In the reaction of tetraphenyl arsonium chloride with the complexes Cl3PO (Cl4)-MoN R (R=CCl3, C2Cl5) the POCl3 is displaced by chloride and yields [Cl5MoN R] . From the i.r. spectra a structure with six-coordinated molybdenum and a MoN triple bond can be deduced. By reaction with water in acetonitrile the molybdenum is reduced to Mo(V) and the nitride ligand is removed yielding (AsPh4)2[(MoOCl4)2CH3CN]. The crystal structure of this compund was determined with X-ray diffraction data. In the tetragonal structure (space group P4/n) AsPh4+ cations and two different anions were found: square pyramidal [MoOCl4] and [MoOCl4 · NCCH3] in which the nitrile is bonded in trans position to the oxygen. The short Mo O distances of 165 pm indicate a strong π-bonding.  相似文献   

4.
Synthesis, Crystal Structure, and Phase Transition of Se4(MoOCl4)2 Dark green, very air sensitive crystals of Se4(MoOCl4)2 are formed from selenium and MoOCl4 at 190°C in a sealed, evacuated glass ampoule in quantitative yield. The structure is built of nearly square planar Se42+ ions and centrosymmetric dimeric MoOCl4? ions which are linked by bridging Cl atoms. At ?21°C Se4(MoOCl4)2 undergoes a reversible solid state phase transition of first order. Structure determinations at ?70°C and 23°C show that during the phase transition the structures of the ions remain unchanged, while the orientations of the ions with respect to each other change in such a way that in the low temperature form the Se42+ ions obtain a higher coordination number by Cl and O atoms of neighboring MoOCl4? ions.  相似文献   

5.
6.
Molybdenum(VI) bis(imido) complexes [Mo(NtBu)2(LR)2] (R=H 1 a ; R=CF3 1 b ) combined with B(C6F5)3 ( 1 a /B(C6F5)3, 1 b /B(C6F5)3) exhibit a frustrated Lewis pair (FLP) character that can heterolytically split H−H, Si−H and O−H bonds. Cleavage of H2 and Et3SiH affords ion pairs [Mo(NtBu)(NHtBu)(LR)2][HB(C6F5)3] (R=H 2 a ; R=CF3 2 b ) composed of a Mo(VI) amido imido cation and a hydridoborate anion, while reaction with H2O leads to [Mo(NtBu)(NHtBu)(LR)2][(HO)B(C6F5)3] (R=H 3 a ; R=CF3 3 b ). Ion pairs 2 a and 2 b are catalysts for the hydrosilylation of aldehydes with triethylsilane, with 2 b being more active than 2 a . Mechanistic elucidation revealed insertion of the aldehyde into the B−H bond of [HB(C6F5)3]. We were able to isolate and fully characterize, including by single-crystal X-ray diffraction analysis, the inserted products Mo(NtBu)(NHtBu)(LR)2][{PhCH2O}B(C6F5)3] (R=H 4 a ; R=CF3 4 b ). Catalysis occurs at [HB(C6F5)3] while [Mo(NtBu)(NHtBu)(LR)2]+ (R=H or CF3) act as the cationic counterions. However, the striking difference in reactivity gives ample evidence that molybdenum cations behave as weakly coordinating cations (WCC).  相似文献   

7.
Oxo-Mo(VI) imido-chloride, [MoOCl2(NH)(Et 2O)] n and nitrido-chloride, [Mo2O2Cl2(N)2(Et 2O)] n have been synthesized by equimolar reactions of MoOCl4 with HN(SiMe 3)2 and LiN(SiMe 3)2, respectively. Higher molar reactions of HN(SiMe 3)2 lead to imido-silylamido derivatives, [Mo2OCl3(NH)3(NHSiMe 3)] n , whereas those of LiN(SiMe 3)2 give silylimido bridged compounds, Mo4O4Cl4(NSiMe 3)6 and Mo4O4(NSiMe 3)8. Elemental analyses, redox titration, magnetic moment, molecular weight, molar conductance, infrared,1H-NMR and TG-DTG-DTA studies are reported.
Reaktionen von Bis(trimethylsilyl)amin und -amid mit MoOCl4
Zusammenfassung Durch equimolare Reaktionen von MoOCl4 mit HN(SiMe 3)2 und LiN(SiMe 3)2 wurden die Oxo-Mo(VI) Imido-chloride [MoOCl2(NH)(Et 2O)] n und die Nitrido-chloride [Mo2O2Cl2(N)2(Et 2O)] n dargestellt. Höhermolekulare Reaktionen von HN(SiMe 3)2 führen zu Imido-silylamido Derivaten [Mo2OCl3(NH)3(NHSiMe 3)] n , währenddessen die von LiN(SiMe 3)2 silylimidoüberbrückte Verbindungen ergeben: Mo4O4Cl4(NSiMe 3)6 und Mo4O4(NSiMe 3)8. Die Strukturen sind mit Elementaranalysen, Redoxtitrationen, Messung der magnetischen Momente, Molekulargewichten, molarer Leitfähigkeit, Infrarot,1H-NMR und TG-DTG-DTA-Untersuchungen charakterisiert.
  相似文献   

8.
孙婷婷 《高分子科学》2011,29(5):520-531
The effect of channel-protein interaction on the translocation of a protein-like chain through a finite channel under certain electric field was studied by using dynamical Monte Carlo simulations.The interior behavior of chain conformation under different interactions was investigated,such as the number of monomers outside of channel nout,monomers inside of channel nm,mean-square radius of gyration〈S2〉and the average energy〈U〉.It shows that with strong attractive interaction,the translocation is more difficult than moderate interaction.At the same time,the dependence of translocation time with different interactions shows that moderate repulsive interaction(εcp= 0.5) accelerates the translocation.Although the waiting time for successful translocation ofεcp = 1.0 is the longest,the average translocation time is not very large.It is far smaller than that ofεcp=-1.0.The probability distributions of translocation time p(t’) and the probability distributions of three duration times p(t1’),p(t2’) and p(t3’) were all discussed.Log-normal distributions are found.All these findings will strengthen the understanding of protein translocation.  相似文献   

9.
《化学:亚洲杂志》2017,12(8):910-919
Reduction of aluminum(III), gallium(III), and indium(III) phthalocyanine chlorides by sodium fluorenone ketyl in the presence of tetrabutylammonium cations yielded crystalline salts of the type (Bu4N+)2[MIII(HFl−O)(Pc.3−)].−(Br) ⋅ 1.5 C6H4Cl2 [M=Al ( 1 ), Ga ( 2 ); HFl−O=fluoren‐9‐olato anion; Pc=phthalocyanine] and (Bu4N+) [InIIIBr(Pc.3−)].− ⋅ 0.875 C6H4Cl2 ⋅ 0.125 C6H14 ( 3 ). The salts were found to contain Pc.3− radical anions with negatively charged phthalocyanine macrocycles, as evidenced by the presence of intense bands of Pc.3− in the near‐IR region and a noticeable blueshift in both the Q and Soret bands of phthalocyanine. The metal(III) atoms coordinate HFl−O anions in 1 and 2 with short Al−O and Ga−O bond lengths of 1.749(2) and 1.836(6) Å, respectively. The C−O bonds [1.402(3) and 1.391(11) Å in 1 and 2 , respectively] in the HFl−O anions are longer than the same bond in the fluorenone ketyl (1.27–1.31 Å). Salts 1 – 3 show effective magnetic moments of 1.72, 1.66, and 1.79 μB at 300 K, respectively, owing to the presence of unpaired S= 1/2 spins on Pc.3−. These spins are coupled antiferromagnetically with Weiss temperatures of −22, −14, and −30 K for 1 – 3 , respectively. Coupling can occur in the corrugated two‐dimensional phthalocyanine layers of 1 and 2 with an exchange interaction of J /k B=−0.9 and −1.1 K, respectively, and in the π‐stacking {[InIIIBr(Pc.3−)].−}2 dimers of 3 with an exchange interaction of J /k B=−10.8 K. The salts show intense electron paramagnetic resonance (EPR) signals attributed to Pc.3−. It was found that increasing the size of the central metal atom strongly broadened these EPR signals.  相似文献   

10.
Kinetic and thermodynamic investigations were performed for a mixed aqueous-organic, 1:1 (v/v) water–1,4-dioxane medium, which was found to be an efficient solvent for the interaction of a neutral dichlorotris(triphenylphosphine) ruthenium(II), RuCl2(PPh3)3 complex with carbon monoxide at atmospheric pressure. During the interaction, RuCl2(PPh3)3 dissociates to a neutral complex dichlorobis(triphenylphosphine) ruthenium(II), RuCl2(PPh3)2, by losing a coordinated PPh3 ligand and RuCl2(PPh3)2 coordinates with CO to form an in situ carbonyl complex RuCl2(CO)(PPh3)2. The in situ formed carbonyl complex RuCl2(CO)(PPh3)2 was thoroughly characterized by equilibrium, spectrophotometric, IR, and electrochemical techniques. Under equilibrium conditions, the rate and dissociation constants for the dissociation of PPh3 from RuCl2(PPh3)3 were found to be favorable for the formation of the carbonyl complex RuCl2(CO)(PPh3)2. The rates of complexation for the formation of RuCl2(CO)(PPh3)2 were found to follow an overall second-order kinetics being first order in terms of the concentrations of both carbon monoxide and RuCl2(PPh3)2. The determined activation parameters corresponding to the rate constant (ΔH# = 35.9 ± 2.5 kJ mol−1 and ΔS# = −122 ± 6 J K−1 mol−1) and thermodynamic parameters corresponding to the formation constant (ΔH° = −33.5 ± 4.5 kJ mol−1, ΔS° = −25 ± 8 J K−1 mol−1, and ΔG° = −25.7 ± 2.0 kJ mol−1) were found to be highly favorable for the formation of the complex RuCl2(CO)(PPh3)2. © 2008 Wiley Periodicals, Inc. Int J Chem Kinet 40: 359–369, 2008  相似文献   

11.
Molybdenum-95 NMR chemical shifts are reported for a series of Mo(O) compounds of the type Mo(CO)4(pip)2−nLn(n = 1,2; L = substituted pyridine ligands). The σ(95Mo) values correlate well with the pKa values for the substituted pyridines; for the n = 1 series, σ (95Mo) ranges from − 1053 ppm (pKa = 1.86 for 4-CN) to − 1120 ppm (pKa = 9.61 for 4-NMe2). The effects of solvent polarity and some in situ reactivity studies are described and the nature of the MoL bond compared to that with piperidine and some other ligands is discussed.  相似文献   

12.
The data on temperature, solvent, and high hydrostatic pressure influence on the rate of the ene reactions of 4‐phenyl‐1,2,4‐triazoline‐3,5‐dione ( 1 ) with 2‐carene ( 2 ), and β‐pinene ( 4 ) have been obtained. Ene reactions 1 + 2 and 1 + 4 have high heat effects: ∆Hrn ( 1 + 2 ) −158.4, ∆Hrn( 1 + 4 ) −159.2 kJ mol−1, 25°C, 1,2‐dichloroethane. The comparison of the activation volume (∆V( 1 + 2 ) −29.9 cm3 mol−1, toluene; ∆V( 1 + 4 ) −36.0 cm3 mol−1, ethyl acetate) and reaction volume values (∆Vr‐n( 1 + 2 ) −24.0 cm3 mol−1, toluene; ∆Vr‐n( 1 + 4 ) −30.4 cm3 mol−1, ethyl acetate) reveals more compact cyclic transition states in comparison with the acyclic reaction products 3 and 5 . In the series of nine solvents, the reaction rate of 1+2 increases 260‐fold and 1+4 increases 200‐fold, respectively, but not due to the solvent polarity.  相似文献   

13.
Two structural isomers containing five second-row element atoms with 24 valence electrons were generated and identified by matrix-isolation IR spectroscopy and quantum chemical calculations. The OCBNO complex, which is produced by the reaction of boron atoms with mixtures of carbon monoxide and nitric oxide in solid neon, rearranges to the more stable OBNCO isomer on UV excitation. Bonding analysis indicates that the OCBNO complex is best described by the bonding interactions between a triplet-state boron cation with an electron configuration of (2s)0(2pσ)0(2pπ)2 and the CO/NO ligands in the triplet state forming two degenerate electron-sharing π bonds and two ligand-to-boron dative σ bonds.  相似文献   

14.
The performance of effective core potentials (ECP) for the main group elements of group IV has been studied by calculating the geometries and reaction energies of isodesmic reactions for the molecules M(CH3)nCl4 ? n (M = C, Si, Ge, Sn, Pb; n = 0–4) at the Hartree–Fock level of theory. The results are compared with data from all electron calculations and experimental results as far as available. The all electron calculations were performed with a 3-21G(d) and a 6-31G(d) basis set for Si, a (43321/4321/41) basis set for Ge, and a (433321/43321/431) basis set for Sn. For the ECP calculations the potentials developed by Hay and Wadt with a configuration (n)sa(n)pb and the valence basis set (21/21), extended by a set of d functions, are employed. © 1992 by John Wiley & Sons, Inc.  相似文献   

15.
The change in the valence state of nanocluster can induce remarkable changes in the properties and structure. However, achieving the valence state changes in nanoclusters is still a challenge. In this work, we use Cu2+ as dopant to “oxidize” [Ag62S12(SBut)32]2+ (4 free electrons) to obtain the new nanocluster: [Ag62−xCuxS12(SBut)32]4+ with 2 free electrons. As revealed by its structure, the [Ag62−xCuxS12(SBut)32]4+ (x=10∼21) has a similar structure to that of [Ag62S12(SBut)32]2+ precursor and all the Cu atoms occupy the surface site of nanocluster. It′s worth noting that with the Cu atoms doping, the [Ag62−xCuxS12(SBut)32]4+ nanocluster is more stable than [Ag62S12(SBut)32]2+ at higher temperature and in electrochemical cycle. This result has laid a foundation for the subsequent application and exploration. Overall, this work reveals crystals structure of a new Ag−Cu nanocluster and offers a new insight into the electron reduction/oxidation of nanocluster.  相似文献   

16.
The results of ab initio molecular orbital calculations for [CrOfs]2? and polarised single crystal electronic spectra of [MoOCl3(Op(NMe2)3)2] and Ph4As[MoOCl4(H2O)] are presented. These data are consistent with the electronic transitions of the MO3+ moieties, O2pπ → Mdxy and Mdxy → Mddxy,dyz being the lowest energy transitions in the spectra of their respective complexes, both these transitions being of low intensity.  相似文献   

17.
This contribution focuses on complex [Mo2(H)2(μ-AdDipp2)2] ( 1 ) and tetrahydrofuran and pyridine adducts [Mo2(H)2(μ-AdDipp2)2(L)2] ( 1⋅thf and 1⋅py ), which contain a trans-(H)Mo≣Mo(H) core (AdDipp2=HC(NDipp2)2; Dipp=2,6-iPr2C6H3). Computational studies provide insights into the coordination and electronic characteristics of the central trans-Mo2H2 unit of 1 , with four-coordinate, fourteen-electron Mo atoms and ϵ-agostic interactions with Dipp methyl groups. Small size C- and N-donors give rise to related complexes 1⋅L but only one molecule of P-donors, for example, PMe3, can bind to 1 , causing one of the hydrides to form a three-centered, two-electron (3c-2e) Mo-H→Mo bond ( 2⋅PMe3 ). A DFT analysis of the terminal and bridging hydride coordination to the Mo≣Mo bond is also reported, along with reactivity studies of the Mo−H bonds of these complexes. Reactions investigated include oxidation of 1⋅thf by silver triflimidate, AgNTf2, to afford a monohydride [Mo2(μ-H)(μ-NTf2)(μ-AdDipp2)2] ( 4 ), with an O,O’-bridging triflimidate ligand.  相似文献   

18.
Group V Nb-polyoxometalate (Nb-POM) chemistry generally lacks the elegant pH-controlled speciation exhibited by group VI (Mo, W) POM chemistry. Here three Nb-POM clusters were isolated and structurally characterized; [Nb14O40(O2)2H3]14−, [((UO2)(H2O))3Nb46(UO2)2O136H8(H2O)4]24−, and [(Nb7O22H2)4(UO2)7(H2O)6]22−, that effectively capture the aqueous Nb-POM species from pH 7 to pH 10. These Nb-POMs illustrate a reaction pathway for control over speciation that is driven by counter-cations (Li+) rather than pH. The two reported heterometallic POMs (with UO22+ moieties) are stabilized by replacing labile H2O/HO−Nb=O with very stable O=U=O. The third isolated Nb-POM features cis-yl-oxos, prior observed only in group VI POM chemistry. Moreover, with these actinide-heterometal contributions to the burgeoning Nb-POM family, it now transects all major metal groups of the periodic table.  相似文献   

19.
Summary Oxomolybdenum(V) complexes of the type (LH4) [MoOCl5] (where LH2 = dimethylene bis-2-benzimidazole or tetramethylene bis-2-benzimidazole), [MoOCl3(LH2)] (where LH2 = tetramethylene bis-2-benzimidazole), [(Mo2O4Cl2-(H2O)3)2(LH2)] (where LH2 = dimethylene bis-2-benzimidazole, tetramethylene bis-2-benzimidazole or hexamethylene bis-2-benzimidazole) and [Mo2O3Cl4(LH2)2] (where LH2 = tetramethylene bis-5-nitro-2-benzimidazole) were prepared and characterised. The mononuclear complexes show u.v.-vis. absorptions characteristic of octahedral molybdenum(V). The dinuclear complexes do not absorb in the visible region, possibly due to the presence of an Mo2O 4 2 +} core, which is also indicated by their diamagnetic behaviour. The biological activities of the free ligands and their complexes have been studied.  相似文献   

20.
Heterometal‐doped gold clusters are poorly accessible through wet‐chemical approaches and main‐group‐metal‐ or early‐transition‐metal‐doped gold clusters are rare. Compounds [M(AuPMe3)11(AuCl)]3+ (M=Pt, Pd, Ni) ( 1 – 3 ), [Ni(AuPPh3)(8?2n)(AuCl)3(AlCp*)n] (n=1, 2) ( 4, 5 ), and [Mo(AuPMe3)8 (GaCl2)3(GaCl)]+ ( 6 ) were selectively obtained by the transmetalation of [M(M′Cp*)n] (M=Mo, E=Ga, n=6; M=Pt, Pd, Ni, M′=Ga, Al, n=4) with [ClAuPR3] (R=Me, Ph) and characterized by single‐crystal X‐ray diffraction and ESI mass spectrometry. DFT calculations were used to analyze the bonding situation. The transmetalation proved to be a powerful tool for the synthesis of heterometal‐doped gold clusters with a design rule based on the 18 valence electron count for the central metal atom M and in agreement with the unified superatom concept based on the jellium model.  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