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 914 毫秒
1.
The specific decomposition rates of chemically activated methylcyclobutane produced from CH2(1A1) reaction with cyclobutane have been determined. CH2(1A1)was produced from ketene photolyses at 3340 and 3130 Å and from diazomethane photolyses at 4358 and 3660 Å. Comparisons of the excitation energies of the methylcyclobutane, determined by RRKM theory calculations, and the experimental results for the ketene systems, with thermochemically predicted maximum excitation energies, favor an Arrhenius A factor in the range of 5 × 1015 to 1 × 1016 sec?1 for methylcyclobutane. This result is consistent with (1) the comparison of RRKM theory calculations and the experimental unimolecular falloff for methylcyclobutane, (2) the comparison of experimental A factors for cyclobutane and other alkylcyclobutane decompositions, and (3) two out of three reported experimental A factors for methylcyclobutane. An analysis of these and previous results leads to a value of the CH2(1A1) ? CH2(3B1) energy splitting of 9±3 kcal/mole.  相似文献   

2.
Approximate vibrational energy distribution for CH*2(1A1) from diazomethane photolyses at 4358 and 3660 Å have been determined to be reasonably broad. These distributions apply to CH*2(1A1) at the time of reaction with cyclobutane and were deduced from the internal energy distribution of the formed chemically activated methylcyclobutane. An apparent anomaly in the pressure dependence of the decompositions of CH2(1A1) generated chemically molecules is explained. The anomaly pertains to the relative behavior of systems utilizing ketene and diazomethane photolyses as CH2(1A1) sources. The explanation offered is that the vibrational energy distributions for CH*2(1A1) are narrow for ketene photolyses at 3340 or 3130 Å and broad for diazomethane photolyses at 4358 or 3660 Å.  相似文献   

3.
A kinetic study has been made of the 3130-Å photolysis of CH2O (8 torr) in O2-containing mixtures (0.02–8 torr) and in the presence of added CO2 (0–300 torr) at 25°C. Quantum yields of formation of H2, CO, and CO2 and the loss of O2 were measured. Φ and ΦCO were much above unity. In an explanation of these unexpected results, a new H-atom-forming chain mechanism was postulated involving HO2 and HO addition to CH2O: CH2O + hν → H + HCO (1) H + CH2O → H2 + HCO (3) H + O2 + M → HO2 + M (6) HCO + O2 → HO2 + CO (8) HO2 + CH2O → (HO2CH2O) → HO + HCO2H (15) HO + CH2O → H2O + HCO? (16); HCO? → H + CO (19) HO + CH2O → H2O + HCO (17) and HO + CH2O → HCO2H + H (18). When the results are rationalized in terms of this mechanism, the data suggest k16 ? k17 and k16/k18 ? 0.5. The data require that a reassessment of the relative rates of reactions (7) and (8) be made, since in the previous work HCO2H formation was used as a monitor of the rate of reaction (7) HCO + O2 + M → HCOO2 + M (7). The present data from experiments at P = 8 torr and P = 1–4 torr give k7[M]/(k7[M] + k8) ≥ 0.049 ± 0.017. These data coupled with the k8 estimates of Washida and coworkers give k7 ≥ (4.4 ± 1.6) × 1011 l2/mol2·sec for M = CH2O. The reaction sequence proposed here is consistent with the observed deterimental effect of O2 addition on the laser-induced isotope enrichment in HDCO. In additional studies of CH2O-O2-isobutene mixtures it was found that Φ was equal to ?2 as estimated in O2-free CH2O-isobutene mixtures. These results suggest that the increase in CO (ν = 1) product observed with O2 addition in CH2O photolysis does not result from perturbations in the fragmentation pattern of the excited CH2O, but it is likely that it originates in the occurrence of the exothermic reaction HCO + O2 → HO2 + CO (ν = 1).  相似文献   

4.
In the title compound, C4H12N22+·2C8H7O3?·2CH4O, the cations lie across centres of inversion and are disordered over two orientations with equal occupancy; there are equal numbers of (R)‐ and (S)‐mandelate anions present (mandelate is α‐hydroxy­benzene­acetate). The anions and the neutral water mol­ecules are linked by O—H?O hydrogen bonds [O?O 2.658 (3) and 2.682 (3) Å, and O—H?O 176 and 166°] into deeply folded zigzag chains. Each orientation of the cation forms two symmetry‐related two‐centre N—H?O hydrogen bonds [N?O 2.588 (4) and 2.678 (4) Å, and N—H?O 177 and 171°] and two asymmetric, but planar, three‐centre N—H?(O)2 hydrogen bonds [N?O 2.686 (4)–3.137 (4) Å and N—H?O 137–147°], and by means of these the cations link the anion/water chains into bilayers.  相似文献   

5.
Deep blue-violet single crystals of hitherto unknown chromous orthophosphate have been obtained reducing CrPO4 by elemental Cr at temperatures above 1050°C in evacuated silica ampoules (NH4I or I2 as mineraliser). The complex structure of Cr3(PO4)2 (P212121, Z = 8, a = 8.4849(10) Å, b = 10.3317(10) Å, c = 14.206(2) Å) contains six crystallographically independent Cr2+ per unit cell. Five of them are coordinated by four oxygen atoms which form a distorted (roof shaped) square plane as first coordination sphere at interatomic distances 1.96 Å ? d(Cr? O) ? 2.15 Å. Their coordination is completed by additional oxygen atoms (2 or 3) at distances 2.32 Å ? d(Cr? O) ? 3.21 Å. The sixth Cr2+ shows six-fold octahedral coordination with strong radial distortion (d(Cr? O): 1.97, 2.04, 2.15, 2.28, 2.29, 2.53 Å). The four different [PO4] groups exhibit only minor deviations from ideal tetrahedral geometry (1.51 Å ? d(P? O) ? 1.57 Å, 104.3° ? ∠(O? P? O) ? 114.4°). An unusually low magnetic moment μexp = 4.28(2) μBP = ?54.8(5) K) has been observed for Cr2+.  相似文献   

6.
The mechanism of the photolysis of formaldehyde was studied in experiments at 3130 Å and in the pressure range of 1–12 torr at 25°C. The experiments were designed to establish the quantum yields of the primary decomposition steps (1) and (2), CH2O + hν → H + HCO (1): CH2O + hν → H2 + CO (2), through the effects of added isobutene, trimethylsilane, and nitric oxide on ΦCO and Φ. The ratio ΦCO/Φ was found to be 1.01 ± 0.09(2σ) and (Φ + ΦCO)/2 = 1.10 ± 0.08 over the range of pressures and a 12-fold change in incident light intensity. Isobutene and nitric oxide additions reduced Φ to about the same limiting value, 0.32 ± 0.03 and 0.34 ± 0.04, respectively, but these added gases differed in their effects on ΦCO. With isobutene addition ΦCO/Φ reached a limiting value of 2.3; with NO addition ΦCO exceeded unity. The addition of small amounts of Me3SiH reduced Φ to 1.02 ± 0.08 and lowered ΦCO to 0.7. These findings were rationalized in terms of a mechanism in which the “nonscavengeable,” molecular hydrogen is formed in reaction (2) with ?2 = 0.32 ± 0.03, while the “free radical” hydrogen is formed in reaction (1) with ?1 = 0.68 ± 0.03. In the pure formaldehyde system these reactions are followed by (3)–(5): H + CH2O → H2 + HCO (3); 2HCO → CH2O + CO (4); 2HCO → H2 + 2CO (5). The data suggest k4/k5 ? 5.8. Isobutene reduced Φ by the reaction H + iso-C4H8 → C4H9 (20), and the results give k20/k3 ? 43 ± 4, in good agreement with the ratio of the reported values of the individual constants k3 and k20.  相似文献   

7.
Ab initio SCF and CI calculations on the cationic and neutral complexes of formaldehyde and lithium are reported. For the cationic complex CH2O/Li+, the stabilization energy of 41.7 kcal/mol obtained from the SCF calculation increases to 51.6 kcal/mol if a configuration interaction is introduced. For the neutral complex CH2O?/Li+, the C2v-conformer of the 2A1-state with the equilibrium bond distances of d(C? O) = 1.23 Å and d (O? Li) = 1.90 Å is calculated to be more stable than the 2B1-state with d (C? O) = 1.34 Å, and d (O? Li) = 1.65 Å. Charge transfer and polarization effects upon complex formation are discussed.  相似文献   

8.
The compound Ru2Cl(4-Cl-C6H4CONH)4 was prepared by reaction of Ru2Cl(O2CCH3)4 with 4-Cl-C6H4CONH2 at 180°C. Crystals of the composition Ru2Cl(4-Cl-C6H4CONH)4CH3OH were obtained by slow diffusion of CH3OH containing Et4NCl into a Me2SO solution of the compound. The structure of the crystalline product, which loses solvent of crystallization on removal from the mother liquor, was solved by X-ray crystallography by mounting a single crystal in a capillary containing the mother liquor. The crystals belong to the space group P1? (triclinic crystal system) with a = 12.731(3) Å, b = 14.389(3) Å, c = 12.604(3) Å, α = 103.41(2)°, β = 106.43(2)°, γ = 64.90(2)°, V = 1988.6(8) Å3 and Z = 2. There are two half ruthenium dimers linked by a Cl atom and an uncoordinated solvent CH3OH molecule per asymmetric unit. The ruthenium dimers lie on two centers of inversion at 0, 0, 0 and 1/2, 0, 0. The chloride ions bridge dinuclear cations in the crystal, forming infinite zigzag chains. The average Ru-Ru distance is 2.296[1] Å and each ruthenium atom has a RuClN2O2 coordination sphere where the average Ru′-Ru-Cl angle is virtually linear (175.68[6]°). The metal oxidation states in the complex are + 2 and + 3, giving an average value of + 2.5. The arrangement of four bridging 4-Cl-benzamidato ligands is of the 2 : 2 type. The average Ru-N, Ru-O, Ru-Cl distances and Ru(1)-Cl(1)-Ru(2) angle are 2.036[6] Å, 2.044[5] Å, 2.583[2] Å and 117.26(8)°, respectively. The IR spectrum of the compound shows two N-H stretches at 3380 and 3340 cm?1. The electronic spectrum of the compound in Me2SO exhibits bands at 558 nm (ε = 340 M?1 cm?1), 425 nm (1000) and 320 nm (22,700).  相似文献   

9.
In the title compound, C6H16N22+·2C2H4O5P?, the cations lie across centres of inversion; in the anions, two of the H‐atom sites have 0.50 occupancy. The anions are linked by short O—H?O hydrogen bonds [O?O 2.465 (3)–2.612 (3) Å and O—H?O 165–171°] into sheets of alternating R(12) and R(28) rings, both of which are centrosymmetric; the cations lie at the centres of the larger rings linked to the anion sheet by N—H?O hydrogen bonds [N?O 2.642 (2) Å and N—H?O 176°].  相似文献   

10.
The reaction of the meso-diol, Δ,Λ-[(en)2Rh(OH)2Rh(en)2]4+, with aqueous H2O2 and 1 equiv. of NaOH at 90° forms the μ-peroxo-μ-hydroxo-bridged species Δ,Λ-[(en)2Rh(O2,OH)Rh(en)2]3+ in a yield of ca. 50%. The compound was crystallized as perchlorate and trifluoromethanesulfonate salts. The structure of the latter salt was determined by single-crystal X-ray diffraction. The crystals are triclinic with space group P1 and lattice constants a = 11.895(5), b = 12.491(4), c = 13.053(5) Å, α = 103.98(3), β = 92.59(3), γ = 119.52(6)°. The distances of the metal centres to the bridging peroxo ligand are 1.999(8) and 1.983(6) Å. The O? O distance in the peroxo group is 1.521(14) Å, and the dihedral angle of the Rh? O? O? Rh unit deviates 65° from planarity. The peroxo complex reacts reversibly with acid, and spectrophotometric studies suggest that the reaction involves protonation of the peroxo bridge, with pKa = 2.70(2) at 25° in 1M NaClO4.  相似文献   

11.
The structural parameters of the completely relaxed 4–21G ab initio geometries of more than 30 basic organic compounds are compared to experimental results. Some ranges for systematic empirical corrections, which relate 4–21G bond distances to experimental parameters, are associated with total energy increments. In general, for the currently feasible comparisons, the following corrections can be given which relate calculated distances to experimental rg parameters and calculated angles to rs-structures For CC single bond distances, deviations between calculated and observed parameters (rg) are in the ranges of ?0.006(2) to ?0.010(2) Å for normal or unstrained hydrocarbons; ?0.011(3) to ?0.016(3) Å for cyclobutane type compounds; and +0.001(5) to +0.004(4) Å for CH3 conjugated with CO. For CO single bonds the ranges are ?0.006(9) to +0.002(3) Å for CO conjugated with CO; and ?0.019(3) to ?0.027(9) Å for aliphatic and ether compounds. A very large and exceptional discrepancy exists for the highly strained ethylene oxide, rsre = ?0.049(5) Å and in CH3OCH3 and C2H5OCH3 the rsre differences are ?0.029(5), ?0.040(10) and ?0.025(10) Å. Some of these discrepancies may also be due to deficiencies of the microwave substitution method caused by atomic coordinates close to inertial planes. For CN bonds, two types of NCH3 corrections are from +0.005(6) to ?0.006(6) and from ?0.009(2) to ?0.014(6) Å; and the range for NCO is +0.012(3) to +0.028(4) Å. For isolated CC double bonds the range is + 0.025(2) to +0.028(2) Å. For conjugated CC double bonds the correction is less positive (+0.014(1) Å for benzene). For CO double bonds the corrections are ?0.004(3) to +0.003(3) Å. For bond angles of type HCH, CCH, CCC, CCO, CCO, OCO, NCO and CCC the corrections are of the order of magnitude about 1–2° (or better). Angles centered at heteroatoms are less accurate than that, when hydrogen atoms are involved. Differences in HOC and NHC angles were found in a range of ?2.3(5)° to ?6.2(4)°.  相似文献   

12.
The adduct 1,6‐di­amino­hexane–1,1,1‐tris(4‐hydroxy­phenyl)­ethane (1/2) is a salt {hexane‐1,6‐diyldiammonium–4‐[1,1‐bis(4‐hydroxyphenyl)ethyl]phenolate (1/2)}, C6H18N22+·2C20H17O3?, in which the cation lies across a centre of inversion in space group P. The anions are linked by two short O—H?O hydrogen bonds [H?O 1.74 and 1.76 Å, O?O 2.5702 (12) and 2.5855 (12) Å, and O—H?O 168 and 169°] into a chain containing two types of R(24) ring. Each cation is linked to four different anion chains by three N—H?O hydrogen bonds [H?O 1.76–2.06 Å, N?O 2.6749 (14)–2.9159 (14) Å and N—H?O 156–172°]. In the adduct 2,2′‐bipyridyl–1,1,1‐tris(4‐hydroxy­phenyl)­ethane (1/2), C10H8N2·2C20H18O3, the neutral di­amine lies across a centre of inversion in space group P21/n. The tris­(phenol) mol­ecules are linked by two O—H?O hydrogen bonds [H?O both 1.90 Å, O?O 2.7303 (14) and 2.7415 (15) Å, and O—H?O 173 and 176°] into sheets built from R(38) rings. Pairs of tris­(phenol) sheets are linked via the di­amine by means of a single O—H?N hydrogen bond [H?N 1.97 Å, O?N 2.7833 (16) Å and O—H?N 163°].  相似文献   

13.
A 2-D coordination polymer, (C7N4H16)2{NH(CH3)3}[{K(H2O)}4Na(H2O)5{Co4(H2O)2(B-α-PW9O34)2}]·2H2O (1), was hydrothermally synthesized and structurally characterized by IR spectroscopy, elemental analysis, X-ray powder diffraction, and X-ray single-crystal crystallography. Crystal structure analysis shows a triclinic space group Pī with a?=?12.4677(8)?Å, b?=?12.5054(8)?Å, c?=?18.5745(1)?Å, α?=?73.3220(1)°, β?=?87.1890(1)°, γ?=?62.2710(1)°, and V?=?2443.4(3)?Å3. Sandwich-type tetra-cobalt(II)-substituted [Co4(H2O)2(B-α-PW9O34)2]10? of 1 consists of two trivacant Keggin [B-α-PW9O34]9? moieties and a rhomb-like Co4O16 unit. Each sandwich-type polyoxotungstate subunit connects 12 K(1) and K(2) centers from two adjacent 1-D K-chain units resulting in an interesting 2-D layer framework. Magnetic properties of 1 have been investigated.  相似文献   

14.
The synthesis and X-ray single crystal study of two mixed-ligand Cu(II) complexes are performed: (CH3C(NCH3)CHC(O)CH3)(CF3C(O)CHC(O)CF3)Cu (1) (space group P21/c, a = 7.0848(12) Å, b = 17.854(3) Å, c = 11.837(2) Å, β = 100.495(6)°, V = 1472.4(4) Å3, Z = 4), (CH3C(NC6H5)CHC(O)CH3)· (CF3C(O)CHC(O)CF3)Cu (2) (space group P-1, a = 9.1119(4) Å, b = 9.6954(4) Å, c = 11.1447(6) Å, α = 113.784(2)°, β = 92.383(2)°, γ = 95.402(2)°, V = 893.52(7) Å3, Z = 2). The structures are molecular, formed from neutral mixed-ligand copper complexes. The central copper atom has the (3O+N) coordination environment with average Cu-O distances of 1.948 Å and Cu-N of 1.932 Å; the chelate O-Cu-N angle (average) is 94.0°. In the structures, the complexes are linked into dimeric associates with Cu…Cu distances of 3.197 Å (for 1) and 3.246 Å (for 2). The volatility of mixed-ligand complexes 1 and 2 is in between of that of the starting homo-ligand complexes.  相似文献   

15.
In the title compounds, C6H8N3O2+·NO3? and C5­H6­N3­O2+·­CH3SO3?, respectively, the cations are almost planar; the twist of the nitr­amino group about the C—N and N—N bonds does not exceed 10°. The deviations from coplanarity are accounted for by intermolecular N—H?O interactions. The coplanarity of the NHNO2 group and the phenyl ring leads to the deformation of the nitr­amino group. The C—N—N angle and one C—C—N angle at the junction of the phenyl ring and the nitr­amino group are increased from 120° by ca 6°, whereas the other junction C—C—N angle is decreased by ca 5°. Within the nitro group, the O—N—O angle is increased by ca 5° and one O—N—N angle is decreased by ca 5°, whereas the other O—N—N angle remains almost unchanged. The cations are connected to the anions by relatively strong N—H?O hydrogen bonds [shortest H?O separations 1.77 (2)–1.81 (3) Å] and much weaker C—H?O hydrogen bonds [H?O separations 2.30 (2)–2.63 (3) Å].  相似文献   

16.
Studies on the equilibrium in the system MCl2? CH3OH? H2O at 25°C and 50°C (M = Sr2+, Ba2+) show that the dehydration in the water-methanol systems proceeds stepwise and all possible lower crystal hydrates may be obtained at 25°C depending on the molar ratio for the mixed solvent. The dehydration and solvation processes in the three-component system MCl2? CH3OH? H2O (M = Mg2+, Ca2+, Sr2+, Ba2+) have been considered in general and compared with those in the bromide system.  相似文献   

17.
The quantum yields of phosphorescence (Φp) of biacetyl have been determined in pure biacetyl, biacetyl-SO2, and biacetyl-c-C6H12 mixtures in experiments using bands of radiation centered at 3450, 3650, 3880, and 4348 Å. It has been shown that the unexpected effect of gas concentration on the quantum yields of the sulfur dioxide triplet-sensitized phosphorescence of biacetyl resulted largely from the significant destruction of biacetyl triplets at the wall of the cell. The kinetics of the variation of Φp with [Ac2], wavelength of the absorbed light, and added gases provide new estimates of the energy relations and the rate constants for the decomposition reaction of vibrationally rich biacetyl molecules in the first excited singlet state (1Ac2?): 1Ac2? → products (1), 1Ac2? + Ac21Ac2 + Ac2 (2); the minimum energy necessary in 1Ac2? for reaction (1) to occur is estimated to be about 72.8 kcal/mole above the ground state of biacetyl: k1/k2 = (4.3 ± 0.1) × 10?3M at 3450 Å, (4.07 ± 0.04) × 10?4M at 3650 Å, and (5.6 ± 0.4) × 10?5M at about 3800 Å. The variation of the rate constant ratio is shown to be consistent with the expectations of the simple theory of excited molecule decomposition. Biacetyl triplet (3Ac2) rate constants were determined by measurements of Φp in O2 and NO-containing mixtures: 3Ac2 + S → (Ac2–S, products) (8); for O2 = S, k8 = (5.76 ± 0.40) × 108 (3650 Å experiments), (5.76 ± 0.27) × 108 (4358 Å); for NO = S, k8 = (3.34 ± 0.20) × 109 (3650 Å), (3.33 ± 0.18) × 109 1./mole-sec (4358 Å). A comparison between these and previous findings of the SO2 triplet (3SO2)-sensitized excitation of biacetyl [5,6] show that the decomposition of the initial 3Ac2 product of the exothermic energy transfer reaction 3SO2 + Ac2 → SO2 + 3Ac2 is unimportant.  相似文献   

18.
Crystal and Molecular Structure of Dicaesium-μ-Oxodecafluorodiarsenate, Cs2(As2F10O) The crystal structure of Cs2(As2F10O) has been determined from three-dimensional data. The compound crystallizes in the monoclinic space group P21/m, the lattice constants being a = 9.175(4), b = 10.690(5), c = 5.619(3)(Å); β = 105,50(5)°. The anion (As2F10O)2– with the point symmetry Cs contains a As? O? As-bridge, whose partial π-bonding is to be discussed. The bond lengths and angles are: As? O: 1.77(2) and 1.68(2) Å, resp., As? O? As: 139(1)°; As…As: 3.225(4) Å, the numbers in parantheses being the standard deviation of the last figure.  相似文献   

19.
Three salts containing different iodobismuthate anions have been synthesized. [(CH3)2NH2]3[BiI6] was prepared by oxidation of bismuth by iodine in N,N‐dimethylformamide. [(CH3)2NH2]3[BiI6] crystallizes in the space group with a = 30.760(3) Å and c = 8.8039(5) Å and contains monomeric [BiI6]3? anions. The hydrate Na4[Bi2I10] · 14H2O was prepared by the oxidation of bismuth using iodine in acetonitrile in the presence of NaI. Na4[Bi2I10] · 14H2O crystallizes in the space group C2/m with a = 12.875(2) Å, b = 16.177(2) Å, c = 9.904(2) Å and β = 106.57(6)°. The structure contains dimeric [Bi2I10]4? anions and rows of sodium ions, with bridging water molecules. The hydrate [Na{(CH3)2NCHO}2(H2O)]3[Bi2I9] was prepared by dissolution of Na4[Bi2I10] · 14H2O in N,N‐dimethylformamide and crystallizes in the space group with a = 13.2309(13) Å, b = 13.3791(14) Å, c = 18.722(2) Å, α = 70.338(9)°, β = 72.651(9)° and γ = 62.183(5). The structure contains dimeric [Bi2I9]3? anions and cationic polymers, equation/tex2gif-stack-1.gif[Na{(CH3)2NCHO}2(H2O)]+.  相似文献   

20.
Methyliminodiacetic acid (H2Mida) and imidazole react with copper(II) to form crystals of the square pyramidal complex [Cu(Mida)Im]. One N and two O atoms of the Mida ligand (Cu-N 2.010(1) Å, Cu-O 1.955(1) Å, and 1.978(1) Å) and the imidazole N atom (1.950(1) Å) lie at the base of the pyramid. The carboxyl O atom of the neighboring complex lies at the apical position (2.411(1) Å); in this way the individual complexes are linked into infinite zigzag chains. Substitution of imidazole by 1,10-phenanthroline gave [Cu2(Mida)2(Phen)H2O]·2H2O crystals with two nonequivalent centrosymmetric octahedral anions [Cu(Mida)2]2? of face type (Cu-N 2.023 Å and 2.028(2) Å, Cu-Oax 2.579 Å and 2.530(2) Å, Cu-Obas 1.952 Å and 1.936(2) Å). The anions serve as bridges in chains between the [Cu(Phen)H2O]2+ cation fragments to which they are bonded by their axial carboxyl groups. The Cu atom of the cation has a [4+1] environment (with the H2O molecule lying on the axis of the pyramid, and with two N atoms of the ligand and two O atoms of the anions lying at the base).  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