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 710 毫秒
1.
The reactons of (CO2)2+ and (CO)2+ with various additives have been investigated using the NBS high-pressure photoionization mass spectrometer at total pressures of 0.4–1.0 torr of either CO2 or CO. The additives include CH4, CD4, C2H2, O2, H2O, 15,14N2O, and CO in both CO2 and 13CO2. Second- and third-order rate coefficients based on an ambipolar diffusion model are reported for 25 separate reaction pairs at 295°K, as well as sequential cationic reaction mechanisms. An approximate value of 225 ± 3 kcal/mol (941 ± 13 kJ/mol) was derived for ΔHf (CO)2+ based on the kinetics observed in various CO-additive mixtures. Some projections regarding the utility of the data under other conditions are also included.  相似文献   

2.
Ab initio molecular orbital calculations with split-valence plus polarization basis sets and incorporating valence-electron correlation have been performed to determine the equilibrium structure of ethyloxonium ([CH3CH2OH2]+) and examine its modes of unimolecular dissociation. An asymmetric structure (1) is predicted to be the most stable form of ethyloxonium, but a second conformational isomer of Cs symmetry lies only 1.4 kJ mol?1 higher in energy than 1. Four unimolecular decomposition pathways for 1 have been examined involving loss of H2, CH4, H2O or C2H4. The most stable fragmentation products, lying 65 kJ mol?1 above 1, are associated with the H2 elimination reaction. However, large barriers of 257 and 223 kJ mol?1 have to be surmounted for H2 and CH4 loss, respectively. On the other hand, elimination of either C2H4 or H2O from ethyloxonium can proceed without a barrier to the reverse associations and, with total endothermicities of 130 and 160 kJ mol?1, respectively, these reactions are expected to dominate at lower energies. A second important equilibrium structure on the surface is a hydrogen-bridged complex, lying 53 kJ mol?1 above 1. This complex is involved in the C2H4 elimination reaction, acts as an intermediate in the proton-transfer reaction connecting [C2H5]+ +H2O and C2H4 + [H3O]+ and plays an important role in the isotopic scrambling that has been observed experimentally in the elimination of either H2O or C2H4 from ethyloxonium. The proton affinity of ethanol was calculated as 799 kJ mol?1, in close agreement with the experimental value of 794 kJ mol?1.  相似文献   

3.
The hydrolysis of (C2H5)2Sn2+, (C2H5)3Sn+ and (n‐C3H7)3Sn+ has been studied, by potentiometric measurements ([H+]‐glass electrode), in NaNO3, NaCl, NaCl/Na2SO4 mixtures and in a synthetic seawater (SSWE), as an ionic medium simulating the major composition of natural seawater, at different ionic strengths (0 ≤ I ≤ 5 mol dm?3) and salinities (15 ≤ S ≤ 45), and at t = 25 °C. Five hydrolytic species for (C2H5)2Sn2+, three for (C2H5)3Sn+ and two for (C3H7)3Sn+ are found. Interactions with the anion components of SSWE, considered as single‐salt seawater, are determined by means of a complex formation model. A predictive equation for the calculation of unknown hydrolysis constants of trialkyltin(IV) cations, such as tributyltin(IV), in NaNO3, NaCl, and SSWE media at different ionic strengths is proposed. Equilibrium constants obtained are also used to determine the interaction parameters of Pitzer equations. Copyright © 2001 John Wiley & Sons, Ltd.  相似文献   

4.
High-energy collisional activation mass spectrometry of HFe(CO)5+ ions shows that Fe(CO)5 is protonated on the iron atom rather than on one of the ligands. This finding is supported by ab initio quantum chemical calculations. The value of the proton affinity of Fe(CO)5 was measured by high-pressure mass spectrometry to be 857 kJ mol?1. The Fe? CO bond dissociation energies for HFe(CO)n+ (n = 1–5) were measured by energy-variable low-energy collisional activation mass spectrometry. The Fe? H bond dissociation energies in HFe(CO)n+ ions were also determined. A synergistic effect on the strengths of the Fe? H and Fe? CO bonds in HFe(CO)+ is noticed. It is demonstrated that the electronically unsaturated species HFe(CO)n+ (n = 3, 4) formed in exothermic proton-transfer reactions with Fe(CO)5 form adducts with CH4. Adducts between C2H5+ or C3H5+ and Fe(CO)n are observed. These adducts are probably formed in direct reactions between the respective carbocations and Fe(CO)5.  相似文献   

5.
李强国  叶丽娟  首梦娟 《中国化学》2003,21(12):1580-1585
IntroductionBothrareearthions1and 8 hydroxyquinolineareofantibacterialfunction ,2 andtheircomplexeshavemorepowerfuldisinfection .Theirbinarycomplexeswerereport edasearlyasin 196 3.Atthesametime ,theresearchontheirternarycomplexeshavebecomeveryactiveinrecentyears,andtheyarewidelyappliedinmanyfields .3 6Dong6 reportedthesynthesisandcharacterizationofthecomplexesofrareearthtrichloroaceticacidsaltswith 8 hy droxyquinoline.Itsapplicationinleathermouldyproofshowedthatthecomplexeshavepowerfuldisinfe…  相似文献   

6.
The mass spectra of a series of β-ketosilanes, p-Y? C6H4Me2SiCH2C(O)Me and their isomeric silyl enol ethers, p-Y? C6H4Me2SiOC(CH3)?CH2, where Y = H, Me, MeO, Cl, F and CF3, have been recorded. The fragmentation patterns for the β-ketosilanes are very similar to those of their silyl enol ether counterparts. The seven major primary fragment ions are [M? Me·]+, [M? C6H4Y·]+, [M? Me2SiO]+˙, [M? C3H4]+˙, [M? HC?CCF3]+˙, [Me2SiOH]+˙ and [C3H6O]+˙ Apparently, upon electron bombardment the β-ketosilanes must undergo rearrangement to an ion structure very similar to that of the ionized silyl enol ethers followed by unimolecular ion decompositions. Substitutions on the benzene ring show a significant effect on the formation of the ions [M? Me2SiO]+˙ and [Me2SiOH]+˙, electron donating groups favoring the former and electron withdrawing groups favoring the latter. The mass spectral fragmentation pathways were identified by observing metastable peaks, metastable ion mass spectra and ion kinetic energy spectra.  相似文献   

7.
The electron impact induced mass spectra of [CF3SMn(CO)4]2, [CF3SeMn(CO)4]2, [CF3SFe(CO)3]2, [CF3SeFe(CO)3]2, CF3SeFe(CO)2C5H5 and CF3SCr(NO)2C5H5 are reported. These compounds exhibit weak molecular ion peaks and undergo preferential loss of CO or NO groups. The CO or NO free fragments suffer typical loss of ECF2(E = S, Se) with the simultaneous shift of F from carbon to metal. The ions [FFeC5H5]+ and [FCrC5H5]+ in the spectra of the cyclopentadienyl compounds prefer expulsion of π-cyclopentadienyls. The pyrolysis effects on the spectra of the compounds have been studied. An increase in temperature eases the expulsion of ECF2 groups from all the compounds and favors the formation of [Fe(C5H5)2]+ and [Cr(C5H5)2]+ in the cyclopentadienyl compounds.  相似文献   

8.
Cathinones belong to a group of compounds of great interest in the new psychoactive substances (NPS) market. Constant changes to the chemical structure made by the producers of these compounds require a quick reaction from analytical laboratories in ascertaining their characteristics. In this article, three cathinone derivatives were characterized by X-ray crystallography. The investigated compounds were confirmed as: 1-[1-(4-methylphenyl)-1-oxohexan-2-yl]pyrrolidin-1-ium chloride ( 1 , C17H26NO+·Cl?, the hydrochloride of 4-MPHP), 1-(4-methyl-1-oxo-1-phenylpentan-2-yl)pyrrolidin-1-ium chloride ( 2 ; C16H24NO+·Cl?, the hydrochloride of α-PiHP) and methyl[1-(4-methylphenyl)-1-oxopentan-2-yl]azanium chloride ( 3 ; C13H20NO+·Cl?, the hydrochloride of 4-MPD). All the salts crystallize in a monoclinic space group: 1 and 2 in P21/c, and 3 in P21/n. To the best of our knowledge, this study provides the first detailed and comprehensive crystallographic data on salts 1 – 3 .  相似文献   

9.
The density functional theory (DFT) and the complete active space self‐consistent‐field (CASSCF) method have been used for full geometry optimization of carbon chains C2nH+ (n = 1–5) in their ground states and selected excited states, respectively. Calculations show that C2nH+ (n = 1–5) have stable linear structures with the ground state of X3Π for C2H+ or X3Σ? for other species. The excited‐state properties of C2nH+ have been investigated by the multiconfigurational second‐order perturbation theory (CASPT2), and predicted vertical excitation energies show good agreement with the available experimental values. On the basis of our calculations, the unsolved observed bands in previous experiments have been interpreted. CASSCF/CASPT2 calculations also have been used to explore the vertical emission energy of selected low‐lying states in C2nH+ (n = 1–5). Present results indicate that the predicted vertical excitation and emission energies of C2nH+ have similar size dependences, and they gradually decrease as the chain size increases. © 2008 Wiley Periodicals, Inc. Int J Quantum Chem, 2009  相似文献   

10.
Oxidation of the cyclohexadienyl complex Fe(η5-C5H5)(1-5-75-6-exo-C5H5-C6H6) (2) by (Ph3C)PF6 (CH2Cl2, from −30 to +20 °C) occurs as two concurrent processes: elimination of an H atom from the cyclohexadienyl ligand and replacement of an H atom in the cyclopentadienyl ring by a CPh3 fragment. A mixture of cationic complexes [Fe(η5-C5H5) (η6-Ph-C5H5]+ (1+) and [Fe(η5-C5H4CPh3) (η6-Ph-C5H5]+ (4+) (4 +) with PF6 anions is obtained. Deprotonation of the mixture of 1+ and 4+ complexes under the action of Bu t OK inm-xylene followed by boiling of the reaction mixture gives phenylferrocene (7) as the product of η66 haptotropic rearrangement. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, NO. 5, pp. 1045–1047, May, 1997.  相似文献   

11.
The complex (η5-C5H4CH3)Mn(NO)(PPh3)I has been prepared by the reaction of NaI with [(η5-C5H4CH3)Mn(NO)(CO)(PPh3)]+ and also by the reaction of [(η5-C5H4CH3)Mn(NO)(CO)2]+ with NaI followed by PPh3. This iodide compound reacts with NaCN to yield (η5-C5H4CH3)Mn(NO)(PPh3)CN which is ethylated by [(C2H5)3O]BF4 to yield [(η5-C5H4CH3)Mn(NO)(PPh3)(CNC2H5)]+. Both [(η5-C5H4CH3)Mn(NO)(CO)2]+ and [(η5-C5H4CH3)Mn(NO)(PPh3)(CO)]+ react with NaCN to yield [(η5-C5H4CH3)Mn(NO)(CN)2]?. This anion reacts with Ph3SnCl to yield cis-(η5-C5H4CH3)Mn(NO)(CN)2SnPh3 and with [(C2-H5)3O]BF4 to yield [(η5-C5H4CH3)Mn(NO)(CNC2H5)2]+. The reaction of (η5-C5-H4CH3)Mn(NO)(PPh3)I with AgBF4 in acetonitrile yields [(η5-C5H4CH3)Mn-(NO)(PPh3)(NCCH3)]+. The complex (η5-C5H4CH3)Mn(NO)(CO)I, produced in the reaction of [(η5-C5H4CH3)Mn(NO)(CO)2]+ with NaI, is not stable and decomposes to the dimeric complex (η5-C5H4CH3)2Mn2(NO)3I for which a reasonable structure is proposed. Similar dimers can be prepared from the other halide salts. The reaction of (η7-C7H7)Mo(CO)(PPh3)I with NaCN yields (η7-C7-H7)Mo(CO)(PPh3)CN which is ethylated by [(C2H5)3O]BF4 to yield [(η7-C7H7)-Mo(CO)(PPh3)(CNC2H5)]+. The interaction of this molybdenum halide complex with AgBF4 in acetonitrile and pyridine yields [(η7-C7H7)Mo(CO)(PPh3)-(NCCH3)]+ and [(η7-C7H7)Mo(CO)(PPh3)(NC5H5)]+, respectively. Both (η5-C5-H4CH3)Mn(NO)(PPh3)I and (η7-C7H7)Mo(CO)(PPh3)I are oxidized by NOPF6 to the respective 17-electron cations in acetonitrile at ?78°C but revert to the neutral halide complex at room temperature. This result is supported by electrochemical data.  相似文献   

12.
The mechanism of ethylene insertion reactions catalyzed by cationic δ‐alkyl platinum complexes has been studied at the B3LYP level of density functional theory. The initial steps of the reactions proceed via the coordination of ethylene to the reactants L2Pt(II)R+, where L2=none, (NH3)2, (CHNH)2; R=H, CH3, C2H5 in which ethylene coordinates strongly to the complexes PtCH+3 and PtC2H+5 (coordination energies (CE) are 296.52 and 229.28 kJ/mol, respectively), while nitrogen‐containing ligands decrease the energies: Pt(NH3)2CH+3 (CE: 180.04 kJ/mol), Pt(NH3)2C2H+5 (CE: 97.86 kJ/mol), Pt(CHNH)2CH+3 (CE : 176.31 kJ/mol) and Pt(CHNH)2C2H+5 (CE: 91.00 kJ/mol). Moreover, ethylene insertion into the Pt‐alkyl bond, which is the rate‐determining step, is endothermic with barrier heights for L2PtCH3(C2H4)+ decreasing in the order: PtCH+3 (164.18 kJ/mol)>(NH3)2 PtCH+3 (129.95 kJ/mol)>(CHNH)2 PtCH+3 (115.27 kJ/mol), which has the same tendency for the ethyl case. The insertion product will continually undergo β‐hydride elimination, which is exothermic. On the other hand, the effects of solvent (dichloromethane, THF and benzene) are investigated with PCM method, but the inclusion of the effects in the computations only slightly affects the results. Beside that, a complete catalytic cycle for ethylene dimerization is studied in detail and the calculations agree well with known energetic and recognized tendencies.  相似文献   

13.
The novel ternary solid complex Gd(C5H8NS2)3(C12H8N2) has been obtained from the reaction of hydrous gadolinium chloride, ammonium pyrrolidinedithiocarbamate (APDC), and 1,10-phenanthroline (o-phen · H2O) in absolute ethanol. The complex was described by an elemental analysis, TG-DTG, and an IR spectrum. The enthalpy change of the complex formation reaction from a solution of the reagents, Δr H m ϑ (sol), and the molar heat capacity of the complex, c m , were determined as being − 15.174 ± 0.053 kJ/mol and 72.377 ± 0.636 J/(mol K) at 298.15 K by using an RD496-III heat conduction microcalorimeter. The enthalpy change of a complex formation from the reaction of the reagents in a solid phase, Δr H m ϑ (s), was calculated as being 52.703 ± 0.304 kJ/mol on the basis of an appropriate thermochemical cycle and other auxiliary thermodynamic data. The thermodynamics of the formation reaction of the complex was investigated by the reaction in solution. Fundamental parameters, the activation enthalpy (ΔH ϑ ), the activation entropy (ΔS ϑ ), the activation free energy (ΔG ϑ ), the apparent reaction rate constant (k), the apparent activation energy (E), the preexponential constant (A), and the reaction order (n), were obtained by the combination of the thermochemical data of the reaction and kinetic equations, with the data of thermokinetic experiments. The constant-volume combustion energy of the complex, Δc U, was determined as being −17588.79 ± 8.62 kJ/mol by an RBC-II type rotatingbomb calorimeter at 298.15 K. Its standard enthalpy of combustion, Δc H m ϑ , and standard enthalpy of formation, Δf H m ϑ , were calculated to be −17604.28 ± 8.62 and −282.43 ± 9.58 kJ/mol, respectively. The text was submitted by the authors in English.  相似文献   

14.
Comparative results on the reduction of 4,6,7,8-tetrahydro-7,7-dimethyl-2H-1-benzopyran-2,5(3H)-diones 1 are reported. Hydride reduction (LiAlH4 in Et2O or NaBH4 in i-PrOH) affords 2,3,4,6,7,8-hexahydro-5H-1-benzopyran-5-ones 5 in 30-60% isolated yield. Photochemical reduction of 1b and 1d (direct irradiation at λ = 300 or 254 nm in i-PrOH, or sensitized irradiation in acetone/i-PrOH or benzene/i-PrOH) gives 3-(6-oxo-1-cyclohexenyl)alkanoic acids 6 in 50–80%, while 1c affords the isomeric 3-(4,4-dimethyl-6-oxo-1-cyclohexenyl)-4-methyl-4-pentenoic acid ( 9 ) in 73% isolated yield. Electrochemical reduction (Hg, CH3CN, Bu4N+ClO, ?2.6 V vs. Ag/Ag+) requires more than 4 Farad/mol for the consumption of 1 without any major product being detected.  相似文献   

15.
The o‐substituted hybrid phenylphosphines, PPh2(o‐C6H4NH2) and PPh2(o‐C6H4OH), could be deprotonated with LDA or n‐BuLi to yield PPh2(o‐C6H4NHLi) and PPh2(o‐C6H4OLi), respectively. When added to a solution of (η5‐C5H5)Fe(CO)2I at room temperature, these two lithiated reagents produce a chelated neutral complex 1 (η5‐C5H5)Fe(CO)[C(O)NH(o‐C6H4)PPh2C,P‐η2] for the former and mainly a zwitterionic complex 2 , (η5‐C5H5)Fe+(CO)2[PPh2(o‐C6H4O?)] for the latter. Complex 1 could easily be protonated and then decarbonylated to give 4 [(η5‐C5H5)Fe(CO){NH2(o‐C6H4)PPh2N,P‐η2}+]. Complexes 1 and 4‐I have been crystallographically characterized with X‐ray diffraction.  相似文献   

16.
Tetranitratogold(III) Acid, (H5O2)[Au(NO3)4]·H2O: Synthesis, Crystal Structure, and Thermal Behaviour of the First Acidic Nitrate of Gold Yellow single crystals of (H5O2)[Au(NO3)4]·H2O grow upon cooling of a solution of Au(OH)3 in conc. nitric acid. The crystal structure contains (monoclinic, C2/c, Z = 4, a = 1214.5(2), b = 854.4(1), c = 1225.7(2) pm, β = 117.75(1)°, Rall = 0.0331) the Au3+ ion in coordination of four monodentate NO3 ligands. The [Au(NO3)4] units are linked by H5O2+‐ions. Significant hydrogen bonding is observed in the crystal structure between the H5O2+ ions and the H2O molecules. The thermal analysis reveals a five step decomposition leading to elemental gold.  相似文献   

17.
In the two title complexes, (C24H20P)[Au(C3S5)2]·C3H6O, (I), and (C20H20P)[Au(C3S5)2], (II), the AuIII atoms exhibit square‐planar coordinations involving four S atoms from two 2‐thioxo‐1,3‐dithiole‐4,5‐dithiolate (dmit) ligands. The Au—S bond lengths, ranging from 2.3057 (8) to 2.3233 (7) Å in (I) and from 2.3119 (8) to 2.3291 (10) Å in (II), are slightly smaller than the sum of the single‐bond covalent radii. In (I), there are two halves of independent Ph4P+ cations, in which the two P atoms lie on twofold rotation axis sites. The Ph4P+ cations and [Au(C3S5)2] anions are interspersed as columns in the packing. Layers composed of Ph4P+ and [Au(C3S5)2] are separated by layers of acetone molecules. In (II), the [Au(C3S5)2] anions and EtPh3P+ counter‐cations form a layered arrangement, and the [Au(C3S5)2] anions form discrete pairs with a long intermolecular Au...S interaction for each Au atom in the crystal structure.  相似文献   

18.
The synthesis and structural characterization by 1H NMR and 197Au Mössbauer spectroscopy as well as by chiral labelling of the built-in ligands of three different types of arylgold(I) compounds is described.197Au Mössbauer data revealed that the benzyl- and arylgold(I) triphenylphosphine complexes which bear potential coordinating substituents at an ortho position still contain linearly coordinated AuI with 2c-2e gold(I)carbon bonds. The observation of isochronous NME resonances in (S)-2-Me2NCH(Me)C6H4AuPPh3 confirms that no additional intramolecular AuN coordination occurs in solution. Preliminary results of an X-ray diffraction study of 2,6-(MeO)2C6H3AuPPh3 are reported (R = 0.040, PAuC1 angle 172.6°. Unsymmetrical AuC1C2 and AuC1C6 angles of 126.4 and 117.4°, respectively).Pure, uncomplexed arylgold(I) compounds have been isolated from the reaction of diarylgoldlithium compounds (arylaurates) with trimethyltin bromide. (S)-2-Me2NCHMeC6H4Au has a dimeric structure which most likely consists of two monomeric units associated by intermolecular AuN coordination thus forming a ten-membered chelate ring. The structure of insoluble 2-Me2NCH2C6H4Au and 2-Me2NC6H4Au are less clear. The former compound probably has a structure similar to (S)-2-Me2NCHMeC6H4Au (IS/QS values for two-coordinate AuI centers). However, the strongly deviating IS and QS values of 2-Me2NC6H4Au indicate that a polynuclear structure for this compound similar to that proposed for 2-Me2NC6H4Cu cannot be excluded (a polymeric structure containing 2-Me2NC6H4 groups which span three Au atoms by 3c-2e Au2C bonds and AuN coordination).The mixed Au/Cu cluster (2-Me2NCH2C6H4)4Au2Cu2 is accessible via the 12 reaction of (2-Me2NCH2C6H4)4Au2Li2 with CuI. Molecular weight and 1H NMR studies point to a tetranuclear structure in solution, while mass spectrometry shows fragment ions with m/e corresponding to (2-Me2NCH2C6H4)3Au2Cu2+, (2-Me2NCH2C6H4)3Cu2Au+, (2-Me2NCH2C6H4)2CuAu2+ and of (2-Me2NCH2C6H4)2Au+.  相似文献   

19.
The reaction of stoichiometric MeLi with the 1:1 mixture of (?5‐C5H5)Fe(CO)2I/P(OR)3 (R = Me, Et, and Ph) at ?78°C changes the bonding mode between metal and ring from (?5‐C5H5) to (?4exo‐MeC5H5) and the oxidation state of metal from Fe(II) to Fe(O), the novel complexes (?4exo‐MeC5H5)Fe(CO)2P(C)R)3 being obtained in 45‐57% yields. The reaction of trace MeLi with the 1:1 mixture of (?5‐C5H5)Fe(CO)2I/P(OMe)3 at ?78°C results in 70% yield of the phosphonate complex (?5‐C5H5)Fe(CO)2P(O)(OMe)2 which is an Arbuzov‐like dealkylation product from the cationic intermediate [(?5‐C5H5)Fe(CO)2P(OMe)3+] and the iodide. The amines could assist the Arbuzov‐like dealkylation of [(?5‐C5H5)Fe(CO)2P(OMe)3+] [PF6?] where iron‐carbamoyl intermediates are likely involved in the case of primary amines.  相似文献   

20.
The photodissociation of gaseous benzaldehyde (C6H5CHO) at 193, 248, and 266 nm using multimass ion imaging and step‐scan time‐resolved Fourier‐transform infrared emission techniques is investigated. We also characterize the potential energies with the CCSD(T)/6‐311+G(3df,2p) method and predict the branching ratios for various channels of dissociation. Upon photolysis at 248 and 266 nm, two major channels for formation of HCO and CO, with relative branching of 0.37:0.63 and 0.20:0.80, respectively, are observed. The C6H5+HCO channel has two components with large and small recoil velocities; the rapid component with average translational energy of approximately 25 kJ mol?1 dominates. The C6H6+CO channel has a similar distribution of translational energy for these two components. IR emission from internally excited C6H5CHO, ν3 (v=1) of HCO, and levels v≤2, J≤43 of CO are observed; the latter has an average rotational energy of approximately 13 kJ mol?1 and vibrational energy of approximately 6 kJ mol?1. Upon photolysis at 193 nm, similar distributions of energy are observed, except that the C6H5+HCO channel becomes the only major channel with a branching ratio of 0.82±0.10 and an increased proportion of the slow component; IR emission from levels ν1 (v=1) and ν3 (v=1 and 2) of HCO and v≤2, J≤43 of CO are observed; the latter has an average energy similar to that observed in photolysis at 248 nm. The observed product yields at different dissociation energies are compared to statistical‐theory predicted results based on the computed singlet and triplet potential‐energy surfaces.  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