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 46 毫秒
1.
Density functional theory (DFT) calculations are used to investigate the reaction mechanism of V3O8+C2H4. The reaction of V3O8 with C2H4 produces V3O7CH2+HCHO or V3O7+CH2OCH2 overall barrierlessly at room temperature, whereas formation of hydrogen‐transfer products V3O7+CH3CHO is subject to a tiny overall free energy barrier (0.03 eV), although the formation of the last‐named pair of products is thermodynamically more favorable than that of the first two. These DFT results are in agreement with recent experimental observations. The (Ob)2V(OtOt). (b=bridging, t=terminal) moiety containing the oxygen radical in V3O8 is the active site in the reaction with C2H4. Similarities and differences between the reactivities of (Ob)2V(OtOt). in V3O8 and the small VO3 cluster [(Ot)2VOt.] are discussed. Moreover, the effect of the support on the reactivity of the (Ob)2V(OtOt). active site is evaluated by investigating the reactivity of the cluster VX2O8, which is obtained by replacing the V atoms in the (Ob)3VOt support moieties of V3O8 with X atoms (X=P, As, Sb, Nb, Ta, Si, and Ti). Support X atoms with different electronegativities influence the oxidative reactivity of the (Ob)2V(OtOt). active site through changing the net charge of the active site. These theoretical predictions of the mechanism of V3O8+C2H4 and the effect of the support on the active site may be helpful for understanding the reactivity and selectivity of reactive O. species over condensed‐phase catalysts.  相似文献   

2.
[CnH2n?3]+ and [CnH2n?4]+·(n = 7, 8) ions have been generated in the mass spectrometer from CnH2n?3 Br (n = 7, 8) precursors and from two steroids. The relative abundances of competing ‘metastable transitionss’ indicate (partial) isomerization to a common structure (or mixture of structures) prior to decomposition in most examples of all four types of ions. In contrast, [C8H10O]+· and [C8H12O]+· ions, generated from different sources as molecular ions and by fragmentation of steroids, do not decompose through common-intermediates.  相似文献   

3.
The oxidation of D ‐glucitol and D ‐mannitol by CrVI yields the aldonic acid (and/or the aldonolactone) and CrIII as final products when an excess of alditol over CrVI is used. The redox reaction occurs through a CrVI→CrV→CrIII path, the CrVI→CrV reduction being the slow redox step. The complete rate laws for the redox reactions are expressed by: a) −d[CrVI]/dt {kM2 H [H+]2+kMH [H+]}[mannitol][CrVI], where kM2 H (6.7±0.3)⋅10 M s−1 and kMH (9±2)⋅10 M s−1; b) −d[CrVI]/dt {kG2 H [H+]2+kGH [H+]}[glucitol][CrVI], where kG2 H (8.5±0.2)⋅10 M s−1 and kGH (1.8±0.1)⋅10 M s−1, at 33°. The slow redox steps are preceded by the formation of a CrVI oxy ester with λmax 371 nm, at pH 4.5. In acid medium, intermediate CrV reacts with the substrate faster than CrVI does. The EPR spectra show that five‐ and six‐coordinate oxo‐CrV intermediates are formed, with the alditol or the aldonic acid acting as bidentate ligands. Pentacoordinate oxo‐CrV species are present at any [H+], whereas hexacoordinate ones are observed only at pH<2 and become the dominant species under stronger acidic conditions where rapid decomposition to the redox products occurs. At higher pH, where hexacoordinate oxo‐CrV species are not observed, CrV complexes are stable enough to remain in solution for several days to months.  相似文献   

4.
The reaction of [Pt(CH2COMe)(Ph)(cod)] (cod=1,5‐cyclooctadiene) with (ArCH2NH2CH2‐C6H4COOH)+(PF6)? (Ar=4‐tBuC6H4 or 9‐anthryl) in the presence of cyclic oligoethers such as dibenzo[24]crown‐8 (DB24C8) and dicyclohexano[24]crown‐8 (DC24C8) produces {(ce)[ArCH2NH2CH2C6H4COOPt(Ph)(cod)]}+(PF6)? (ce=DB24C8 or DC24C8, Ar=4‐tBuC6H4 or 9‐anthryl) with interlocked structures. FABMS and NMR spectra of a solution of these compounds indicate that the Pt complexes with a secondary ammonium group and DB24C8 (or DC24C8) make up the axis and cyclic components, respectively. Temperature‐dependent 1H NMR spectra of a solution of {(DB24C8)[4‐tBuC6H4CH2NH2CH2‐C6H4COOPt(Ph)(cod)]}+(PF6)? ({(DB24C8)[ 4 ‐H]}+(PF6)?) show equilibration with free DB24C8 and the axis component. The addition of DB24C8 to a solution of {(DC24C8)[ 4 ‐H]}+(PF6)? causes partial exchange of the macrocyclic component of the interlocked molecules, giving a mixture of {(DC24C8)[ 4 ‐H]}+(PF6)?, {(DB24C8)[ 4 ‐H]}+(PF6)?, and free macrocyclic compounds. The reaction of 3,5‐Me2C6H3COCl with {(DB24C8)[ 4 ‐H]}+(PF6)? affords the organic rotaxane {(DB24C8)(4‐tBuC6H4CH2NH2CH2‐C6H4COOCOC6H3Me2‐3,5)}+(PF6)? through C? O bond formation between the aroyl group and the carboxylate ligand of the axis component. The addition of 2,2′‐bipyridine (bpy) to a solution of {(DB24C8)[ 4 ‐H]}+(PF6)? induces the degradation of the interlocked structure to form a complex with trigonal bipyramidal coordination, [Pt(Ph)(bpy)(cod)]+(PF6)?, whereas the reaction of bpy with [Pt(OCOC6H4Me‐4)(Ph)(cod)] produces the square‐planar complex [Pt(OCOC6H4Me‐4)(Ph)(bpy)].  相似文献   

5.
The treatment of N,C,N‐chelated antimony(III) and bismuth(III) chlorides [C6H3‐2,6‐(CH=NR)2]MCl2 [R = tBu and M = Sb ( 1 ) or Bi ( 2 ); R = Dmp and M = Sb ( 3 ) or Bi ( 4 )] (Dmp = 2,6‐Me2C6H3) with one molar equivalent of Ag[CB11H12] led to a smooth formation of corresponding ionic pairs {[C6H3‐2,6‐(CH=NR)2]MCl}+[CB11H12] [R = tBu and M = Sb ( 7 ) or Bi ( 8 ), R = Dmp and M = Sb ( 9 ) or Bi ( 10 )]. Similarly, the reaction of C,N‐chelated analogues [C6H2‐2‐(CH=NDip)‐4,6‐(tBu)2]MCl2 [M = Sb ( 5 ) or Bi ( 6 ), Dip = 2′,6′‐iPr2C6H3] gave compounds {[C6H2‐2‐(CH=NDip)‐4,6‐(tBu)2]MCl}+[CB11H12] [M = Sb ( 11 ) or Bi ( 12 )]. All compounds 7 – 12 were characterized with 1H, 11B and 13C{1H} NMR spectroscopy, ESI‐mass spectrometry, IR spectroscopy, and molecular structures of 7 – 9 and 12 were determined by the help of single‐crystal X‐ray diffraction analysis. In contrast, all attempts to cleave also the second M–Cl bond in 7 – 12 using another molar equivalent Ag[CB11H12] remained unsuccessful. Nevertheless, the reaction between 7 (or 8 ) and Ag[CB11H12] produced unprecedented adducts of both reagents namely {[C6H3‐2,6‐(CH=NtBu)2]SbCl}22+[Ag2(CB11H12)4]2– ( 13 ) and {[C6H3‐2,6‐(CH=NtBu)2]BiCl}+[Ag(CB11H12)2] ( 14 ) in a reproducible manner. The molecular structures of these sparingly soluble compounds were determined by single‐crystal X‐ray diffraction analysis.  相似文献   

6.
[C4H5N] ions have been generated from eleven neutral species. From a study of their metastable transitions and the translational energy released in the fragmentation in which C2H2 is lost, it is concluded that [C4H5N] ions with sufficient energy to decompose do so from a common structure or mixture of structures when they are generated from crotonitrile, allyl cyanide, cyclopropyl cyanide, methacrylontrile, pyrrole, 2-, 3- and 4-hydroxypyridines and 2-aminopyridine. The [C4H5N] ions formed from allyl isocyanide decompose from a different structure and those given by cyclopropyl isocyanide appear to decompose from a mixture of the two structures. Non-decomposing [C4H5N] ions were investigated by means of their collision induced decomposition spectra using a B/E linked scan. Six different structures or mixtures of structures are suggested to explain these observations.  相似文献   

7.
The complexes [Pt(tBu3tpy){C?C(C6H4C?C)n?1R}]+ (n=1: R=alkyl and aryl (Ar); n=1–3: R=phenyl (Ph) or Ph‐N(CH3)2‐4; n=1 and 2, R=Ph‐NH2‐4; tBu3tpy=4,4’,4’’‐tri‐tert‐butyl‐2,2’:6’,2’’‐terpyridine) and [Pt(Cl3tpy)(C?CR)]+ (R=tert‐butyl (tBu), Ph, 9,9’‐dibutylfluorene, 9,9’‐dibutyl‐7‐dimethyl‐amine‐fluorene; Cl3tpy=4,4’,4’’‐trichloro‐2,2’:6’,2’’‐terpyridine) were prepared. The effects of substituent(s) on the terpyridine (tpy) and acetylide ligands and chain length of arylacetylide ligands on the absorption and emission spectra were examined. Resonance Raman (RR) spectra of [Pt(tBu3tpy)(C?CR)]+ (R=n‐butyl, Ph, and C6H4‐OCH3‐4) obtained in acetonitrile at 298 K reveal that the structural distortion of the C?C bond in the electronic excited state obtained by 502.9 nm excitation is substantially larger than that obtained by 416 nm excitation. Density functional theory (DFT) and time‐dependent DFT (TDDFT) calculations on [Pt(H3tpy)(C?CR)]+ (R= n‐propyl (nPr), 2‐pyridyl (Py)), [Pt(H3tpy){C?C(C6H4C?C)n?1Ph}]+ (n=1–3), and [Pt(H3tpy){C?C(C6H4C?C)n?1C6H4‐N(CH3)2‐4}]+/+H+ (n=1–3; H3tpy=nonsubstituted terpyridine) at two different conformations were performed, namely, with the phenyl rings of the arylacetylide ligands coplanar (“cop”) with and perpendicular (“per”) to the H3tpy ligand. Combining the experimental data and calculated results, the two lowest energy absorption peak maxima, λ1 and λ2, of [Pt(Y3tpy)(C?CR)]+ (Y=tBu or Cl, R=aryl) are attributed to 1[π(C?CR)→π*(Y3tpy)] in the “cop” conformation and mixed 1[dπ(Pt)→π*(Y3tpy)]/1[π(C?CR)→π*(Y3tpy)] transitions in the “per” conformation. The lowest energy absorption peak λ1 for [Pt(tBu3tpy){C?C(C6H4C?C)n?1C6H4‐H‐4}]+ (n=1–3) shows a redshift with increasing chain length. However, for [Pt(tBu3tpy){C?C(C6H4C?C)n?1C6H4‐N(CH3)2‐4}]+ (n=1–3), λ1 shows a blueshift with increasing chain length n, but shows a redshift after the addition of acid. The emissions of [Pt(Y3tpy)(C?CR)]+ (Y=tBu or Cl) at 524–642 nm measured in dichloromethane at 298 K are assigned to the 3[π(C?CAr)→π*(Y3tpy)] excited states and mixed 3[dπ(Pt)→π*(Y3tpy)]/3[π(C?C)→π*(Y3tpy)] excited states for R=aryl and alkyl groups, respectively. [Pt(tBu3tpy){C?C(C6H4C?C)n?1C6H4‐N(CH3)2‐4}]+ (n=1 and 2) are nonemissive, and this is attributed to the small energy gap between the singlet ground state (S0) and the lowest triplet excited state (T1).  相似文献   

8.
Two types of sandwich complexes (η5‐MeOCH2CH2C9H6) Ln (η8‐C8H8) (THF)n [Ln=La (1), Nd(2), n=0; Sm(3), Dy (4) and Er (5). n = l] and (η5‐C4H7OCH2C9H6)Ln(η8‐C8H8) (THF) [Ln = La (6), Nd(7). Sm(8). Dy (9) and Er (10)] were synthesized by the reactions of LnCl3 with equivalent mole of K2C8H8, followed by treatment with corresponding potassium salt of ether‐substituted indenide. The molecular structures of 3 and 8 were determined by single crystal X‐ray diffraction. (η5 ‐MeOCH2CH2C9H6) Sm (η8‐C8H8) (THF) (3) monoclinic. Pt1/c, a = 1.4793(3) nm, b = 0.8716 (2) nm, c = 1.6149 (3) nm, β = 98. 17(3), V = 2.0612(7) nm3, Z = 4, R(F)=0.0362. (η5‐C4H7OCH2C9H6)Sm(η8‐C8H8)(THF) (8) orthorhombic. p212121. a = 0.8754(2) nm, b = 1.1000(2) nm, c = 2.3117 (5) nm, V = 2.2260(8) nm3, Z=4, R(F) =0.0497.  相似文献   

9.
Synthesis and Molecular Structure of the Binuclear tert-Butyliminovanadium(IV) Complexes [(μ-NtC4H9)2V2(CH2CMe3)2X2] (X = OtC4H9, CH2CMe3) Syntheses of the neopentylvanadium(V) compounds tC4H9N?V(CH2CMe3)3?n(OtC4H9)n (n = 0 ( 7 ), 1 ( 6 ), 2) are described. 6 and 7 decompose by irradiation splitting off neopentane and yielding the binuclear diamagnetic neopentylvanadium(IV) complexes [(μ-NtC4H9)2V2(CH2CMe3)2X2] [X = OtC4H9 ( 8 ), CH2CMe3 ( 11 )]. All compounds obtained are characterized by 1H and 51V NMR spectroscopy. 8 has been found by X-ray diffraction analysis to be a binuclear complex with bridging tert-butylimino ligands and a vanadium—vanadium single bond. The complexes tC4H9N?V(CH2C6H5)(OtC4H9)2 and [(μ-NtC4H9)2V2(CH2SiMe3)2(OtC4H9)2] ( 10 ) have been also prepared; the crystal structure of 8 and 10 are nearly identical.  相似文献   

10.
The kinetics of the silver(I) catalysed autoxidation of aqueous sulphur(IV) an acetate buffered medium obey the rate law: –d[SIV]/dt = D[AgI][SIV]2[H+]–1/(B+C[SIV]). The rate is independent of [O2] but strongly inhibited by EtOH. A free radical mechanism is proposed.  相似文献   

11.
A major fragmentation route of N-t-butyl-N-isopropyl-and N-ethylphenoxyacetamide is the formation of [C7H8O]+. Deuterium labeling implicates hydrogen transfer from nitrogen in all three species and from the t-butyl groups of the t-butylamide. The anisole structure is assigned to the [C7H8O]+· ion on the basis of its secondary fragmentation.  相似文献   

12.
Abstract

The reaction of Ar[sbnd]P[dbnd]C[dbnd]P[sbnd]Ar (Ar=2.4.6-tBu3C6H2) with electrophiles (H+, S8) proceeds at the phosphorus atom with subsequent cyclisation of an o-tbutyl group.  相似文献   

13.
A study of the thermal decomposition of an acetylene–ethane-d6 mixture indicates that the rate constant for hydrogen abstraction from acetylene by methyl is more than 20 times less than for abstraction from ethane. Isotopic exchange is initiated by a rapid reaction between product D atoms and C2H2. A series of experiments involving the reactions of a D2–acetylene mixture indicated that a molecular exchange process was also occurring, and it was shown that d[C2HD]/dt = k[D2]0.7[C2H2]0.3, effective activation energy = 15.8 kcal/mol. This mechanism made an insignificant contribution to isotope exchange in C2H2–C2D6 mixtures.  相似文献   

14.
The natural compound gossypol forms a stable clathrate with acridine. The composition of the clathrate is C30H30O8·0.5C13H9N. The unit cell is monoclinic, C2/c space group, a = 11.3213(3) ?, b = 30.5957(13) ?, c = 17.0824(4) ?, γ = 94.153(2)°, V = 5901.5(3) ?3, M = 1153.24, Z = 8, d x = 1.369 g/cm3, and R = 0.0413 for 4726 reflections. The structure of the clathrate allows one to refer this compound to the ethyl acetate isomorphic host-guest group of gossypol complexes.  相似文献   

15.
This study reports the use of the trisalkylgallium GaR3 (R=CH2SiMe3), containing sterically demanding monosilyl groups, as an effective Lewis‐acid component for frustrated Lewis pair activation of carbonyl compounds, when combined with the bulky N‐heterocyclic carbene 1,3‐bis(tert‐butyl)imidazol‐2‐ylidene (ItBu) or 1,3‐bis(tert‐butyl)imidazolin‐2‐ylidene (SItBu). The reduction of aldehydes can be achieved by insertion into the C=O functionality at the C2 (so‐called normal) position of the carbene affording zwitterionic products [ItBuCH2OGaR3] ( 1 ) or [ItBuCH(p‐Br‐C6H4)OGaR3] ( 2 ), or alternatively, at its abnormal (C4) site yielding [aItBuCH(p‐Br‐C6H4)OGaR3] ( 3 ). As evidence of the cooperative behaviour of both components, ItBu and GaR3, neither of them alone are able to activate any of the carbonyl‐containing substrates included in this study NMR spectroscopic studies of the new compounds point to complex equilibria involving the formation of kinetic and thermodynamic species as implicated through DFT calculations. Extension to ketones proved successful for electrophilic α,α,α‐trifluoroacetophenone, yielding [aItBuC(Ph)(CF3)OGaR3] ( 7 ). However, in the case of ketones and nitriles bearing acidic hydrogen atoms, C?H bond activation takes place preferentially, affording novel imidazolium gallate salts such as [{ItBuH}+{(p‐I‐C6H4)C(CH2)OGaR3}?] ( 8 ) or [{ItBuH}+{Ph2C=C=NGaR3}?] ( 12 ).  相似文献   

16.
Energy differences, ΔX s−t (X = E, H, and G) (ΔX s−t = X(singlet) − X(triplet)) between singlet (s) and triplet (t) states of C12H8M were calculated at B3LYP/6-311+G*. The DFT calculations indicated that the ΔG s−t between singlet (s) and triplet (t) states of C12H8M were increased from M = C to M = Pb. The ΔG s−t of C12H8M was compared with its analogue C4H4M through replacement of heavy atoms from M = C to M = Pb. Configurations of the electrons in orbitals (σ2 or π2) for the singlet state of C12H8M were discussed.  相似文献   

17.
Substituents have been found to have a marked influence on the metastable ion decompositions and collisionally activated (CA) fragmentations of the M+˙ ion of a number of 1,2,3-triarylpropen-1-ones. An attempt has been made to confirm the structures of the rearrangement ions, [C14H10]+˙, [C13H11]+˙, [C13H9]+ and [C12H8]+˙ by comparison of their CA spectra with those of the corresponding ions produced from reference compounds. The results imply that [C14H10]+˙ and the M+˙ ions of phenanthrene and diphenylacetylene have a common structure, [C13H9]+ and the fluorenyl cation have a common structure and [C12H8]+˙ and biphenylene molecular ion have a common structure. The available data indicate that the ion at m/z 167 consists of a mixture of structures, likely possibilities being diphenylmethyl, phenyltropylium and dihydrofluorenyl cations.  相似文献   

18.
Bivalent germanium was polarographically studied in sodium hydroxide solution at various concentrations. A well-defined reduction wave with half-wave potential varying from ?0.90 to ?0.98 volt vs. S.C.E. was observed for 1×10?4 M Ge(II) in concentration range of 0.2 to 2.0 F with respect to NaOH, and from ?0.70 to ?0.88 volt vs. S.C.E. in the pH range 9.0–12.1 at 25°. The value of id/C is 5.43 μα/mM and that of id/C mfor t1/8 is 5.21. Dependence of E1/2 upon pH is expressed by —E1/2 =0.18+0.058 pH. Experimental result suggests that the reaction proceeds in two-steps involving an irreversible two-electron reduction: Ge(OH)2+OH?→HGeO2?+H2O and HGeO2?+H2O+2e→Ge0+3OH?.  相似文献   

19.
Alkylation of spiro[fluorene-9,3’-indazole] at N(1) and N(2) with tBuCl affords the nitrenium cations [C6H4N2(tBu)C(C12H8)][BF4], 1 and 2 , respectively. Compound 1 converts to 2 over the temperature range 303–323 K with a free energy barrier of 28±5 kcal mol−1. Reaction of 1 with PMe3 afforded the N-bound phosphine adduct [C6H4N(tBu)N(PMe3)C(C12H8)]BF4] 3 . However, phosphines attack 2 at the para-carbon atom of the aryl group with concurrent cleavage of N(2)−C(1) bond and proton migration to C(1) affording [(R3P)C6H3NN(tBu)CH(C12H8)][BF4] (R=Me 4 , nBu 5 ). Analogous reactions of 1 and 2 with the carbene SIMes prompt attack at the para-carbon with concurrent loss of H. affording the radical cation salts [(SIMes)C6H3N(tBu)NC(C12H8).][BF4] 6 and [(SIMes)C6H3NN(tBu)C(C12H8).][BF4] 7 , whereas reaction of 2 with BAC gives the Lewis acid-base adduct, [C6H4N(BAC)N(tBu)C(C12H8)][BF4] 8 . Finally, reactions of 1 and 2 with KPPh2 result in electron transfer affording (PPh2)2 and the persistent radicals C6H4N(tBu)NC(C12H8). and C6H4NN(tBu)C(C12H8).. The detailed reaction mechanisms are also explored by extensive DFT calculations.  相似文献   

20.
Coordination polymers [Ag(Me4Pyz)] PF6(I) and [Ag2(Me4Pyz)3](BF4)2·H2O (II) have been synthesized, and their structures have been determined. The crystals of I are monoclinic, space group C2/c, a = 9.440(2) ?, b = 10.587(2) ?, c = 13.165(3) ?, β= 107.19(3)°, V = 1257.0(5) ?3, d = 2.056 g/cm3, Z = 4. The crystals of II are monoclinic, space group P21/n, a = 13.062(3) ?, b = 12.259(2) ?, c = 18.996(4) ?, β = 97.73(3)°, V = 3014.1(11)?3, ρ = 1.798 g/cm3, Z = 4. The structure of I is built of linear polymeric cations [Ag(C8H12N2)] + and octahedral anions [PF6]. Upon the interaction of tetramethylpyrazine molecule with Ag+ ions, intersecting polymeric chains [Ag(C8H12)] + (1D polymer) are formed extending in mutually perpendicular diagonal directions. The structure of II consists of layers (2D polymers) formed by fused sixmembered rings. These rings consist of Ag+ ions linked by bridging ligands Me4Pyz. Original Russian Text ? Yu.V. Kokunov, Yu.E. Gorbunova, 2007, published in Zhurnal Neorganicheskoi Khimii, 2007, Vol. 52, No. 5, pp. 743–750.  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