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 515 毫秒
1.
New bis[N-(2,6-di-t-butyl-1-hydroxyphenyl)salicylideneminato]copper(II) complexes bearing HO and CH3O substituents on the salicyaldehyde moiety were prepared, and their spectroscopic properties, as well as redox reactivity towards PbO2 and PPh3, examined by ESR and UV spectroscopy. In the process of synthesis of HO complexes unlike CH3O the oxidative C-C coupling of coordinated salicylaldimine ligands does not takes place. The powder ESR spectra of CH3O substituted complexes unlike of HO analogues are typical of a triplet state Cu(II) dimers with a half-field forbidden (deltaM = +/- 2) transition and the allowed transitions (AM = +/- 1) dimeric form of the complexes at 300 and 113 K. The one-electron oxidation of 3-CH3O and all of the OH complexes with PbO2 to give indophenoxyl type secondary radicals which are significantly different from those observed for analogues Cl, Br and NO2 substituted chelates. The presented complexes unlike their electron-withdrawing analogues are readily reduced by PPh3 via intramolecular electron transfer from ligand to copper(II) to give various radical intermediates as well as Cu(I) radical ligand compounds. The analysis of ESR spectra all of the complexes and radical intermediates are presented.  相似文献   

2.
An ab initio computational study of the dual functions of C?S group in the M2C?S ··· HCN (M = H, F, Cl, Br, HO, H3C, H2N) complex has been performed at the MP2(Full)/aug‐cc‐pVTZ level. The C?S group can act as both the electron donor and acceptor, thus two minima complexes were found for each molecular pairs. The interaction energy of hydrogen bond in the F, Cl, or Br substituted complexes is less negative than that in the corresponding H2CS one, while the interaction energy of the σ‐hole interaction is more negative. The OH substitution weakens the hydrogen bond, whereas the H3C and H2N substitution strengthens it. The σ‐hole interaction in the HO, H3C, and H2N complexes is very weak. The substitution effect has been understood with electrostatic induction and conjugation effects. The energy decomposition analysis has been performed for the halogen‐substituted complexes. © 2011 Wiley Periodicals, Inc. Int J Quantum Chem, 2012.  相似文献   

3.
Inter‐residue H‐bonds of oligosaccharides in (D6)DMSO have been assigned on the basis of a combined interpretation of the chemical shift (δ(OH)), coupling constant (J(H,OH)), and temperature dependence (Δδ(OH)/ΔT) of OH signals. Cellobiose, lactose, and N,N′‐diacetylchitobiose possess a completely persistent C(3)OH⋅⋅⋅OC(5′) H‐bond. Maltose is characterised by flip‐flop H‐bonds between HO−C(3) and HO−C(2′), and agarose by two weakly persistent inter‐residue H‐bonds. Sucrose forms an equilibrium of differently H‐bonded species, and hyaluronates possess four strong inter‐residue H‐bonds.  相似文献   

4.
Secondary Hydroxyalkylphosphanes: Synthesis and Characterization of Mono‐, Bis‐ and Trisalkoxyphosphane‐substituted Zirconium Complexes and the Heterobimetallic Trinuclear Complex [Cp2Zr{O(CH2)3PHMes(AuCl)}2] The secondary hydroxyalkylphosphanes RPHCH2OH [R = 2,4,6‐Me3C6H2 (Mes) ( 1 ), 2,4,6‐iPr3C6H2 (Tipp) ( 2 )], 1‐AdPH‐2‐OH‐cyclo‐C6H10 ( 3 ) and RPH(CH2)3OH [R = Ph ( 4 ), Mes ( 5 ), Tipp ( 6 ), Cy ( 7 ), tBu ( 8 )] were obtained from primary phosphanes RPH2 and formaldehyde ( 1 , 2 ) or from LiPHR and cyclohexene oxide ( 3 ) or trimethylene oxide ( 4 ‐ 8 ). Starting from 5 or 7 and [CpR2ZrMe2] [CpR = C5EtMe4 (Cp°), C5H5 (Cp), C5MeH4 (Cp′)], the monoalkoxyphosphane‐substituted zirconocene complexes [CpR2Zr(Me){O(CH2)3PHMes}] [CpR = Cp° ( 9 ), Cp ( 10 )] were prepared. With [CpR2ZrCl2], the bisalkoxyphosphane‐substituted complexes [Cp′2Zr{O(CH2)3PHMes}2] ( 11 ) and [Cp2Zr{O(CH2)3PHCy}2] ( 12 ) are obtained, and with [TpRZrCl3], the trisalkoxyphosphane‐substituted zirconium complexes [TpRZr{O(CH2)3PHMes}3] [TpR = trispyrazolylborato (Tp) ( 13 ), TpR = tris(3,5‐dimethyl)pyrazolylborato (Tp*) ( 14 )] are prepared. The reaction of 5 with [AuCl(tht)] (tht = tetrahydrothiophene) yielded the mononuclear complex [AuCl{PHMes(CH2)3OH}] ( 15 ). The trinuclear complex [Cp2Zr{O(CH2)3PHMes(AuCl)}2] ( 16 ) was obtained from [Cp2ZrCl2] and 15 . Compounds 1 ‐ 16 were characterized spectroscopically (1H‐, 31P‐, 13C‐NMR; IR; MS) and compound 2 also by crystal structure determination. The bis‐ and trisalkoxyphosphane‐substituted complexes 11‐14 and 16 were obtained as mixtures of two diastereomers which could not be separated.  相似文献   

5.
Several novel organotin(IV) complexes with formula SnCl2(CH3)2(X)2, X = C6H5C(O)NHP(O)(NC4H8)2 (1), C6H5C(O)NHP(O)(NC5H10)2 (2), C6H5C(O)NHP(O)[N(CH3)(C6H11)]2 (3), C6H5C(O)NHP(O)[NH-C(CH3)3]2 (4) were synthesized and characterized by 1H, 13C, 31P NMR, IR spectroscopy and elemental analysis. The structures have been determined for each of the four compounds. Compound 1 exists in the form of two symmetrically independent molecules in the crystalline state due to differences in their similar torsion angles. In all of the four structures there are intramolecular -Sn-Cl?H-N- hydrogen bonds, in addition to weak C-H?O and C-H?Cl hydrogen bonds. Both 1H and 13C NMR spectra show the coupling of 119/117Sn nuclei with methyl proton and carbon atoms. The δ(31P) of these complexes are in upfields with respect to their corresponding reported ligands. The spectroscopic and structural properties of these complexes were compared with those corresponding ligands.  相似文献   

6.
Complexes of diacetyl salicylaldehyde oxalic acid dihydrazone, CH3COC(CH3)= NNHCOCONHN=CHC6H4(OH),(dsodh) and diacetyl salicylaldehyde malonic acid dihydrazone CH3COC(CH3)=NNHCOCH2CONHN=CHC6H4(OH), (dsmdh) of general compositions [M(L)]Cl, [M′(L)Cl], [M(L′)]Cl and [M′(L′)Cl] (where M?=?Co(II), Cu(II), Zn(II), Cd(II) and M′?=?Ni(II); HL?=?dsodh and HL′?=?dsmdh) were prepared and characterized by elemental analyses, molar conductance, magnetic moments, electronic, ESR and infrared spectra and X-ray diffraction data. The magnetic moments and electronic spectra indicate six-coordinate octahedral geometry for Co(II) and square planar geometry for Ni(II) complexes. The ESR spectral data of Cu(II) complexes in DMF solution reveal a tetragonally distorted octahedral geometry. Both ligands bond through >C=O, >C=N and deprotonated phenolate groups in all octahedral complexes and through >C=N and deprotonated phenolate groups in Ni(II) square planar complexes. The lattice parameters for Cu(dsodh) and Co(dsmdh) correspond to an orthorhombic and Ni(dsodh) corresponds to a tetragonal crystal lattice.

The complexes show significant antifungal activity against a number of pathogenic fungi viz. Stemphylium, Myrothecium and Alternaria. The antibacterial activity was studied against Pseudomonas fluorescence (gram ?ve) and Clostridium thermocellum (gram +ve).  相似文献   

7.
Five monophosphine‐substituted diiron propane‐1,2‐dithiolate complexes as the active site models of [FeFe]‐hydrogenases have been synthesized and characterized. Reactions of complex [Fe2(CO)6{μ‐SCH2CH(CH3)S}] ( 1 ) with a monophosphine ligand tris(4‐methylphenyl)phosphine, diphenyl‐2‐pyridylphosphine, tris(4‐chlorophenyl)phosphine, triphenylphosphine, or tris(4‐fluorophenyl)phosphine in the presence of the oxidative agent Me3NO·2H2O gave the monophosphine‐substituted diiron complexes [Fe2(CO)5(L){μ‐SCH2CH(CH3)S}] [L = P(4‐C6H4CH3)3, 2 ; Ph2P(2‐C5H4N), 3 ; P(4‐C6H4Cl)3, 4 ; PPh3, 5 ; P(4‐C6H4F)3, 6 ] in 81%–94% yields. Complexes 2 – 6 have been characterized by elemental analysis, spectroscopy, and X‐ray crystallography. In addition, electrochemical studies revealed that these complexes can catalyze the reduction of protons to H2 in the presence of HOAc.  相似文献   

8.
Tri(1‐cyclohepta‐2, 4, 6‐trienyl)phosphane, P(C7H7)3 ([P] when coordinated to a metal atom), was used to stabilize complexes of platinum(II) and palladium(II) with chelating dichalcogenolato ligands as [P]M(E∩E) [E = S, ∩ = CH2CH2, M = Pt ( 3a ); E = S, ∩ = 1, 2‐C6H4, M = Pt ( 5a ), Pd ( 6a ); E = S, ∩ = C(O)C(O), M = Pt ( 7a ), Pd ( 8a ); E = S, Se, ∩ = 1, 2‐C2(B10H10), M = Pt ( 9a, 9b ), Pd ( 10a, 10b ); E = S, ∩ = Fe2(CO)6, M = Pt ( 11a ), Pd ( 12a )]. Starting materials in all reactions were [P]MCl2 with M = Pt ( 1 ) and Pd ( 2 ). Attempts at the synthesis of [P]M(ER)2 with non‐chelating chalcogenolato ligands were not successful. All new complexes were characterized by multinuclear magnetic resonance spectroscopy in solution (1H, 13C, 31P, 77Se and 195Pt NMR), and the molecular structures of 5a and 12a were determined by X‐ray analysis. Both in the solid state and in solution the ligand [P] is linked to the metal atom by the P‐M bond and by η2‐C=C coordination of the central C=C bond of one of the C7H7 rings. In solution, intramolecular exchange between coordinated and non‐coordinated C7H7 rings is observed, the exchange process being markedly faster in the case of M = Pd than for M = Pt.  相似文献   

9.
Some new N‐4‐Fluorobenzoyl phosphoric triamides with formula 4‐F‐C6H4C(O)N(H)P(O)X2, X = NH‐C(CH3)3 ( 1 ), NH‐CH2‐CH=CH2 ( 2 ), NH‐CH2C6H5 ( 3 ), N(CH3)(C6H5) ( 4 ), NH‐CH(CH3)(C6H5) ( 5 ) were synthesized and characterized by 1H, 13C, 31P NMR, IR and Mass spectroscopy and elemental analysis. The structures of compounds 1 , 3 and 4 were investigated by X‐ray crystallography. The P=O and C=O bonds in these compounds are anti. Compounds 1 and 3 form one dimensional polymeric chain produced by intra‐ and intermolecular ‐P=O···H‐N‐ hydrogen bonds. Compound 4 forms only a centrosymmetric dimer in the crystalline lattice via two equal ‐P=O···H‐N‐ hydrogen bonds. 1H and 13C NMR spectra show two series of signals for the two amine groups in compound 1 . This is also observed for the two α‐methylbenzylamine groups in 5 due to the presence of chiral carbon atom in molecule. 13C NMR spectrum of compound 4 shows that 2J(P,Caliphatic) coupling constant for CH2 group is greater than for CH3 in agreement with our previous study. Mass spectra of compounds 1 ‐ 3 (containing 4‐F‐C6H4C(O)N(H)P(O) moiety) indicate the fragments of amidophosphoric acid and 4‐F‐C6H4CN+ that formed in a pseudo McLafferty rearrangement pathway. Also, the fragments of aliphatic amines have high intensity in mass spectra.  相似文献   

10.
The interpretation of 1H‐NMR chemical shifts, coupling constants, and coefficients of temperature dependence (δ(OH), J(H,OH), and Δδ(OH)/ΔT values) evidences that, in (D6)DMSO solution, the signal of an OH group involved as donor in an intramolecular H‐bond to a hydroxy or alkoxy group is shifted upfield, whereas the signal of an OH group acting as acceptor of an intramolecular H‐bond and as donor in an intermolecular H‐bond to (D6)DMSO is shifted downfield. The relative strength of the intramolecular H‐bond depends on co‐operativity and on the acidity of OH groups. The acidity of OH groups is enhanced when they are in an antiparallel orientation to a C−O bond. A comparison of the 1H‐NMR spectra of alcohols in CDCl3 and (D6)DMSO allows discrimination between weak and strong intramolecular H‐bonds. Consideration of IR spectra (CHCl3 or CH2Cl2) shows that the rule according to which the downfield shift of δ(OH) for H‐bonded alcohols in CDCl3 parallels the strength of the H‐bond is valid only for alcohols forming strong intramolecular H‐bonds. The combined analysis of J(H,OH) and δ(OH) values is illustrated by the interpretation of the spectra of the epoxyalcohols 14 and 15 (Fig. 3). H‐Bonding of hexopyranoses, hexulopyranoses, alkyl hexopyranosides, alkyl 4,6‐O‐benzylidenehexopyranosides, levoglucosans, and inositols in (D6)DMSO was investigated. Fully solvated non‐anomeric equatorial OH groups lacking a vicinal axial OR group (R=H or alkyl, or (alkoxy)alkyl) show characteristic J(H,OH) values of 4.5 – 5.5 Hz and fully solvated non‐anomeric axial OH groups lacking an axial OR group in β‐position are characterized by J(H,OH) values of 4.2 – 4.4 Hz (Figs. 4 – 6). Non‐anomeric equatorial OH groups vicinal to an axial OR group are involved in a partial intramolecular H‐bond (J(H,OH)=5.4 – 7.4 Hz), whereas non‐anomeric equatorial OH groups vicinal to two axial OR form partial bifurcated H‐bonds (J(H,OH)=5.8 – 9.5 Hz). Non‐anomeric axial OH groups form partial intramolecular H‐bonds to a cis‐1.3‐diaxial alkoxy group (as in 29 and 41 : J(H,OH)=4.8 – 5.0 Hz). The persistence of such a H‐bond is enhanced when there is an additional H‐bond acceptor, such as the ring O‐atom ( 43 – 47 : J(H,OH)=5.6 – 7.6 Hz; 32 and 33 : 10.5 – 11.3 Hz). The (partial) intramolecular H‐bonds lead to an upfield shift (relative to the signal of a fully solvated OH in a similar surrounding) for the signal of the H‐donor. The shift may also be related to the signal of the fully solvated, equatorial HO−C(2), HO−C(3), and HO−C(4) of β‐D ‐glucopyranose ( 16 : 4.81 ppm) by using the following increments: −0.3 ppm for an axial OH group, 0.2 – 0.25 ppm for replacing a vicinal OH by an OR group, ca. 0.1 ppm for replacing another OH by an OR group, 0.2 ppm for an antiperiplanar C−O bond, −0.3 ppm if a vicinal OH group is (partially) H‐bonded to another OR group, and −0.4 to −0.6 for both OH groups of a vicinal diol moiety involved in (partial) divergent H‐bonds. Flip‐flop H‐bonds are observed between the diaxial HO−C(2) and HO−C(4) of the inositol 40 (J(H,OH)=6.4 Hz, δ(OH)=5.45 ppm) and levoglucosan ( 42 ; J(H,OH)=6.7 – 7.1 Hz, δ(OH)=4.76 – 4.83 ppm; bifurcated H‐bond); the former is completely persistent and the latter to ca. 40%. A persistent, unidirectional H‐bond C(1)−OH⋅⋅⋅O−C(10) is present in ginkgolide B and C, as evidenced by strongly different δ(OH) and Δδ(OH)/ΔT values for HO−C(1) and HO−C(10) (Fig. 9). In the absence of this H‐bond, HO−C(1) of 52 resonates 1.1 – 1.2 ppm downfield, while HO−C(10) of ginkgolide A and of 48 – 50 resonates 0.5 – 0.9 ppm upfield.  相似文献   

11.
Diethylenetriamine‐N,N,N′,N′′,N′′‐pentaacetic acid (DTPA) and 1,4,7,10‐tetraazacyclododecane‐1,4,7,10‐tetraacetic acid (DOTA) scandium(III) complexes were investigated in the solution and solid state. Three 45Sc NMR spectroscopic references suitable for aqueous solutions were suggested: 0.1 M Sc(ClO4)3 in 1 M aq. HClO4 (δSc=0.0 ppm), 0.1 M ScCl3 in 1 M aq. HCl (δSc=1.75 ppm) and 0.01 M [Sc(ox)4]5? (ox2?=oxalato) in 1 M aq. K2C2O4 (δSc=8.31 ppm). In solution, [Sc(dtpa)]2? complex (δSc=83 ppm, ?ν=770 Hz) has a rather symmetric ligand field unlike highly unsymmetrical donor atom arrangement in [Sc(dota)]? anion (δSc=100 ppm, ?ν=4300 Hz). The solid‐state structure of K8[Sc2(ox)7] ? 13 H2O contains two [Sc(ox)3]3? units bridged by twice “side‐on” coordinated oxalate anion with Sc3+ ion in a dodecahedral O8 arrangement. Structures of [Sc(dtpa)]2? and [Sc(dota)]? in [(Hguanidine)]2[Sc(dtpa)] ? 3 H2O and K[Sc(dota)][H6dota]Cl2 ? 4 H2O, respectively, are analogous to those of trivalent lanthanide complexes with the same ligands. The [Sc(dota)]? unit exhibits twisted square‐antiprismatic arrangement without an axial ligand (TSA′ isomer) and [Sc(dota)]? and (H6dota)2+ units are bridged by a K+ cation. A surprisingly high value of the last DOTA dissociation constant (pKa=12.9) was determined by potentiometry and confirmed by using NMR spectroscopy. Stability constants of scandium(III) complexes (log KScL 27.43 and 30.79 for DTPA and DOTA, respectively) were determined from potentiometric and 45Sc NMR spectroscopic data. Both complexes are fully formed even below pH 2. Complexation of DOTA with the Sc3+ ion is much faster than with trivalent lanthanides. Proton‐assisted decomplexation of the [Sc(dota)]? complex (τ1/2=45 h; 1 M aq. HCl, 25 °C) is much slower than that for [Ln(dota)]? complexes. Therefore, DOTA and its derivatives seem to be very suitable ligands for scandium radioisotopes.  相似文献   

12.
本文研究了羰基钼(钨)与双(1-甲基咪唑-2-基)甲酮和双(1-甲基咪唑-2-基)甲烷以及双(1-甲基咪唑-2-基)乙烯的反应,获得了6个双齿螯合的双(1-甲基咪唑-2-基)甲酮,双(1-甲基咪唑-2-基)甲烷和双(1-甲基咪唑-2-基)乙烯四羰基金属衍生物,以及1个单齿配位的双(1-甲基咪唑-2-基)乙烯五羰基钨化合物。它们的结构通过红外,核磁以及X-射线单晶衍射分析得到确证。所有这些新化合物的电化学测试表明,它们只存在一个不可逆的氧化过程。  相似文献   

13.
The aluminum complexes containing two iminophenolate ligands of the type (p‐XC6H4NCHC6H4O‐o)2AlR' (R′=Me ( 3, 4 ) or R′=O(CH2)4OCH=CH2 ( 5, 6 ), X=H ( 3, 5 ), F( 4, 6 )) were synthesized and characterized by 1H, 13C NMR spectroscopy, and X‐ray crystallography. The reaction of AlMe3 with two equivalents of substituted iminophenols gave five‐coordinated {ONR}2AlMe ( 3, 4 ) complexes. Subsequent reaction of these methyl complexes with unsaturated alcohol, HO(CH2)4OCH=CH2, resulted in target compounds 5 and 6 in a good yield. It was shown that the complexes ( 3 ‐ 6 ) are monomeric in solution (NMR) and in solid state (X‐ray analysis). The catalytic activity of the complexes 5 and 6 towards ring‐opening polymerization (ROP) of ?‐caprolactone and d,l ‐lactide was assessed. Complex 5 showed higher activity as compared with 6 , while both of these catalysts induced controlled homo‐ and copolymerization to afford the macromonomers with high content of vinyl ether end groups (Fn > 80%) in a broad range of molecular weights (Mn = 4000–30,000 g mol?1) with relatively narrow MWD (Mw/Mn = 1.1–1.5). © 2014 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2014 , 52, 1237–1250  相似文献   

14.
The regioselective effects of tert‐butyl or bromine as the position‐protecting group of feruloytyamide on the oxidative coupling reactions for the synthesis of natural (±)‐canabisin D were investigated in detail. The coupling yield of 8‐8‐coupled aryldihydronaphthalene product of 5‐Br‐feruloytyamide was higher than that of tert‐butyl substituted precursor under FeCl3·6H2O‐acetone‐water oxidative condition.  相似文献   

15.
Surface Compounds of Transition Metals. XLI [1] Preparation and Properties of Organochromium Compounds by Reaction of Phillips Catalysts with Ethylene Reaction of reduced Phillips catalysts with ethylene at 300 °C deactivates the catalyst; supported organochromium compounds are formed. These can be cleaved from the silica support by HCl and other acids, and transferred into solution by extraction with CH3OH. Chromatography yields fractions of organochromium compounds which differ by CH2 moieties. XPS, 1/2H NMR, and mass spectra as well as magnetic measurements prove that an ensemble of (RnCp)CrCl2(CH3OH) (RnCp = alkylated cyclopentadienyl) has been formed. The RnCp ligand results from a chromium‐assisted oxidative coupling of the olefin with or without CC‐cleavage. According to UV/Vis and mass spectroscopy Cl and CH3OH can be substituted for other anions and donor molecules. Without a donor dinuclear, Cl‐bridged molecules are obtained, of which [(1,2,3‐Me3Cp)CrCl2]2 was established by crystal structure analysis. Reaction with O2 reversibly leads to chromium(V) compounds of the type (R2Cp)Cr(O)Cl2.  相似文献   

16.
Rh‐containing metallacycles, [(TPA)RhIII2‐(C,N)‐CH2CH2(NR)2‐]Cl; TPA=N,N,N,N‐tris(2‐pyridylmethyl)amine have been accessed through treatment of the RhI ethylene complex, [(TPA)Rh(η2CH2CH2)]Cl ([ 1 ]Cl) with substituted diazenes. We show this methodology to be tolerant of electron‐deficient azo compounds including azo diesters (RCO2N?NCO2R; R=Et [ 3 ]Cl, R=iPr [ 4 ]Cl, R=tBu [ 5 ]Cl, and R=Bn [ 6 ]Cl) and a cyclic azo diamide: 4‐phenyl‐1,2,4‐triazole‐3,5‐dione (PTAD), [ 7 ]Cl. The latter complex features two ortho‐fused ring systems and constitutes the first 3‐rhoda‐1,2‐diazabicyclo[3.3.0]octane. Preliminary evidence suggests that these complexes result from N–N coordination followed by insertion of ethylene into a [Rh]?N bond. In terms of reactivity, [ 3 ]Cl and [ 4 ]Cl successfully undergo ring‐opening using p‐toluenesulfonic acid, affording the Rh chlorides, [(TPA)RhIII(Cl)(κ1‐(C)‐CH2CH2(NCO2R)(NHCO2R)]OTs; [ 13 ]OTs and [ 14 ]OTs. Deprotection of [ 5 ]Cl using trifluoroacetic acid was also found to give an ethyl substituted, end‐on coordinated diazene [(TPA)RhIII2‐(C,N)‐CH2CH2(NH)2‐]+ [ 16 ]Cl, a hitherto unreported motif. Treatment of [ 16 ]Cl with acetyl chloride resulted in the bisacetylated adduct [(TPA)RhIII2‐(C,N)‐CH2CH2(NAc)2‐]+, [ 17 ]Cl. Treatment of [ 1 ]Cl with AcN?NAc did not give the Rh?N insertion product, but instead the N,O‐chelated complex [(TPA)RhI ( κ2‐(O,N)‐CH3(CO)(NH)(N?C(CH3)(OCH?CH2))]Cl [ 23 ]Cl, presumably through insertion of ethylene into a [Rh]?O bond.  相似文献   

17.
Carbenes are reactive molecules of the form R1? C?? R2 that play a role in topics ranging from organic synthesis to gas‐phase oxidation chemistry. We report the first experimental structure determination of dihydroxycarbene (HO? C?? OH), one of the smallest stable singlet carbenes, using a combination of microwave rotational spectroscopy and high‐level coupled‐cluster calculations. The semi‐experimental equilibrium structure derived from five isotopic variants of HO? C?? OH contains two very short CO single bonds (ca. 1.32 Å). Detection of HO? C?? OH in the gas phase firmly establishes that it is stable to isomerization, yet it has been underrepresented in discussions of the CH2O2 chemical system and its atmospherically relevant isomers: formic acid and the Criegee intermediate CH2OO.  相似文献   

18.
The gas‐phase pyrolytic and oxidative chemistry of furans has received much attention recently because of their potential as platform chemicals and biofuels. Typically these compounds exhibit very strong ring carbon to H or CH3 bonds. 2‐Methoxyfuran had been reported to be exceptionally unstable in comparison to related substituted heterocycles in pyrolytic experiments. The origins of its reactivity are shown to be due to the very weak O–CH3, which at 189.5 ± 1.9 kJ mol?1 is some 200 kJ mol?1 weaker than C–H bonds in the molecule. We show that the reported reactivity is somewhat overestimated but that does not alter the fact that 2‐methoxyfuran is exceptionally unstable. It may prove to be a useful alternative to azomethane as a thermal source of methyl radicals.  相似文献   

19.
A bimolecular rate constant,kDHO, of (29 ± 9) × 10?12 cm3 molecule?1 s?1 was measured using the relative rate technique for the reaction of the hydroxyl radical (OH) with 3,5‐dimethyl‐1‐hexyn‐3‐ol (DHO, HC?CC(OH)(CH3)CH2CH(CH3)2) at (297 ± 3) K and 1 atm total pressure. To more clearly define DHO's indoor environment degradation mechanism, the products of the DHO + OH reaction were also investigated. The positively identified DHO/OH reaction products were acetone ((CH3)2C?O), 3‐butyne‐2‐one (3B2O, HC?CC(?O)(CH3)), 2‐methyl‐propanal (2MP, H(O?)CCH(CH3)2), 4‐methyl‐2‐pentanone (MIBK, CH3C(?O)CH2CH(CH3)2), ethanedial (GLY, HC(?O)C(?O)H), 2‐oxopropanal (MGLY, CH3C(?O)C(?O)H), and 2,3‐butanedione (23BD, CH3C(?O)C(?O)CH3). The yields of 3B2O and MIBK from the DHO/OH reaction were (8.4 ± 0.3) and (26 ± 2)%, respectively. The use of derivatizing agents O‐(2,3,4,5,6‐pentalfluorobenzyl)hydroxylamine (PFBHA) and N,O‐bis(trimethylsilyl)trifluoroacetamide (BSTFA) clearly indicated that several other reaction products were formed. The elucidation of these other reaction products was facilitated by mass spectrometry of the derivatized reaction products coupled with plausible DHO/OH reaction mechanisms based on previously published volatile organic compound/OH gas‐phase reaction mechanisms. © 2004 Wiley Periodicals, Inc. Int J Chem Kinet 36: 534–544, 2004  相似文献   

20.
Two heterospin complexes [Cu(NIT3Py)(cda)H2O] · H2O ( 1 ) and [Cu(NIT2Py)(cda)H2O] · H2O · CH3OH ( 2 ) with CuII ions and pyridyl‐substituted nitronyl nitroxide radicals (NITxPy = 2‐(x′‐pyridyl)‐4,4,5,5‐tetramethyl‐imidazoline‐1‐oxyl‐3‐oxide, x = 3, 2; H2cda = 4‐hydroxy‐pyridine‐2,6‐dicarboxylic acid) were synthesized and characterized structurally and magnetically. The single crystal structures show that the two complexes are both two‐spin complexes, in which the different radicals make the two complexes have different hydrogen bonding interactions to form 2D and 1D supramolecular network for complexes 1 and 2 , respectively. The magnetic measurements indicate that complexes 1 and 2 both exhibit antiferromagnetic interactions between CuII and radicals.  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