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.
Zusammenfassung Es wurden die Verbindungen HYT *·4 H2O, Y4 T 3·14 H2O, LiYT·4 H2O, NaYT·5 H2O, KYT·3 H2O, RbYT·4 H2O, CsYT·4 H2O, NH4YT·3 H2O, K2YTOH·4 H2O, K3YT(OH)2·4 H2O, K4YT(OH)3·3 H2O, K5YT(OH)4·3 H2O, KYH4 T 2·3 H2O, K2YH3 T 2·5 H2O, K3YH2 T 2·4 H2O, KY2 T(OH)3·5 H2O, K2Y2 T(OH)4·5 H2O isoliert. Die Präparate wurden mit Hilfe von Thermoanalyse, IR-Absorptionsspektren und Röntgenstreuung näher charakterisiert und ihre Löslichkeit in Wasser untersucht.
Some complexes of Yttrium with tartrates were isolated and the compounds characterised by thermogravimetric analysis, IR-spectroscopy and X-ray diffraction. Solubility in water was examined.
  相似文献   

2.
The reaction C2H5 + O2 → C2H5O2 in glassy methanol-d4 and the H-atom abstraction by CH3, C2H5, and n-C4H9 radicals in C2H5OH + C2D5OH and CD3CH2OH + C2D5OH glassy mixtures have been studied by electron spin resonance. The analysis of the dependence of the reaction rates on the concentration of O2 (oxidation) and C2H5OH, CD3CH2OH (H-atom abstraction) has shown that the √t law is not conditioned by the existence of regions characterized by different rate constants.  相似文献   

3.
Hydrogen peroxide oxidation of platinum(II) compounds containing labile groups such as Cl, OH, and alkene moieties has been carried out and the products characterized. The reactions of [PtII (X)2 (N–N)] (X = Cl, OH, X2 = isopropylidenemalorate (ipm); N–N 2,2-dimethyl-1,3-propanediamine [(dmpda), N-isopropyl-1,3-propanediamine (ippda)] with hydrogen peroxide in an appropriate solvent at room temperature affords [PtIV (OH)(Y)(X)2(N–N)] (Y = OH, OCH3). The crystal structures of [PtIV(OH)(OCH3)(Cl)2(dmpda)]·2H2O (P-1 bar, a = 6.339(2) Å , b = 9.861(1) Å, c = 11.561(1) Å, a = 92.078(9)°, β = 104.78(1)°, γ=100.54(1)°, V = 684.3(2) Å3, Z = 2R = 0.0503) and [PtIV(OH)2(ipm)(ippda)]·3H2O (C 2/c, a = 27.275(6) Å, b=6.954(2) Å, c = 22.331(4) Å, β = 118.30(2)°, V = 3729(2) Å3, Z = 8, R = 0.0345) have been solved and refined. The local geometry around the platinum(IV) atom approximates to a typical octahedral arrangement with two added groups (OH and OCH3; OH and OH) in a transposition. The platinum(IV) compounds with potential labile moieties may be important intermediate species for further reactions.  相似文献   

4.
The syntheses of glycosides from the diazirine 1 and a range of alcohols under thermal and/or photolytic conditions are described. Yields and diastereoselectivities depend upon the pKHA values of the alcohols, the solvent, and the reaction temperature. The glycosidation of weakly acidic alcohols (MeOH, EtOH, i-PrOH, and t-BuOH, 1 equiv. each) in CH2Cl2 at room temperature leads to the glycosides 2–5 in yields between 60 and 34% (Scheme 1 and Table 1). At ?70 to ?60°, yields are markedly higher. In CH2Cl2, diastereoselectivities are very low. In THF, at ?70 to ?60°, however, glycosidation of i-PrOH leads to α-D -/β-D - 4 in a ratio of 8:92. More strongly acidic alcohols, such as CF3CH2OH, (CF3)2 CHOH, and (CF3)2C(Me)OH, and the highly fluorinated long-chain alcohols CF3(CF2)5(CH2)2OH ( 11 ) and CHF2(CF2)9CH2OH ( 13 ) react (CH2Cl2, r.t.) in yields between 73 and 85% and lead mainly to the β-D -glucosides β-D - 6 to β-D - 8 , β-D - 12 , and β-D - 14 (d.e. 14–68%). Yields and diastereoselectivities are markedly improved, when toluene, dioxane, 1,2-dimetoxyethane, or THF are used, as examined for the glycosidation of (CF3)2C(Me)OH, yielding (1,2-dimethoxyethane, 25°) 80% of α-D -/ β-D - 8 in a ratio of 2:98 (d.e. 96%; Table 4). In EtCN, (CF3)2C(Me)OH yields up to 55% of the imidate 10 . Glycosidation of di-O-isopropylideneglucose 15 leads to 16 (CH2Cl2, r.t.; 65%, α-D / β-D = 33:67). That glycosidation occurs by initial protonation of the intermediate glycosylidene carbene is evidenced, for strongly acidic alcohols, by the formation of 10 , derived from the attack of (CF3)2MeCO? on an intermediate nitrilium ion (Scheme 4), and for weakly acidic alcohols, by the formation of α-D - 9 and β-D - 9 , derived by attack of i-PrO? on intermediate tetrahydrofuranylium ions. A working hypothesis is presented (Scheme 3). The diastereoselectivities are rationalized on the basis of a protonation in the σ plane of the intermediate carbene, the stabilization of the thereby generated ion pair by interaction with the BnO? C(2) group, with the solvent, and/or with the alcohol, and the final nucleophilic attack by RO? in the π plane of the (solvated) oxonium ion.  相似文献   

5.
Abstract —On photoexcitation, hydroxyacetone undergoes a Norrish-type-1 fragmentation to yield CH3CO and CH2OH. CH2OH is identified by its EPR spectrum. The existence of CH3CO is inferred from the presence of diacetyl and acetaldehyde in irradiated solutions. Above pH 5, in addition to CH2OH, the cis and trans forms of the hydroxyacetone enediol radical anion, CH3C(O-)=C(O***)H, are detected. 1.3-Dihydroxyacetone is photodecomposed to HOCH2C?O and C?H2OH. The former radical decarbonylates to yield CH2OH and CO. At 254 nm the overall quantum yield of CO production is 0.75. Above pH 5, in addition to CH2OH, the cis and trans forms of the 1.3-dihydroxyacetone enediol radical anion, HOCH2C(O-)C(O***)H, are observed. Electronically excited hydroxyacetone and 1.3-dihydroxyacetone react exclusively by C-C fragmentation, and no H-abstraction from H-donors is observed. In contrast, electronically excited 1.3-dicarboxyacetone shows H-abstraction from H-donors in competition with C-C fragmentation. In the absence of H-donors, fragmentation resulting in CH2CO2- and -O2CCH2C?O occurs followed by decarbonylation of -O2H2C?O. At 254 nm the quantum yield of CO production is 0.02. In the presence of H-donors, H-abstraction, yielding HO2CCH2C(OH)CH2CO2, predominates.  相似文献   

6.
Summary Pulsed laser photolysis with resonance fluorescence monitoring of OH radicals was applied at T = 300±2 K to obtain the rate constants of k1= (3.38±0.60)x10-12, k2= (2.52±0.44)x10-13and k3 = (1.06±0.30)x10-13cm3molecule-1s-1with 2σprecision given for the overall reactions OH + CH3CH2OH (1), OH + CF2HCH2OH (2) and OH + CF3CH2OH (3), respectively. k2is the first direct kinetic data for the reaction of OH radicals with CF2HCH2OH reported in the literature.</o:p>  相似文献   

7.
Abstract— Kinetic studies were made of light production and 02 absorption elicited by treatment of dimethylbiacridylium hydroxide [D(OH)2] with reducing agents in alkaline aqueous solutions. D(OH)2 addition stimulated O2 uptake which proceeded with zero-order kinetics until nearly all of the O2 or of the D(OH)2 was converted to end products. At the termination of the reactions with fructose as reductant the D(OH)2 was converted to methylacridone and to dimethylbiacridene each compound accounting for approximately one-half the D(OH)2 consumed. O2 was reduced to H2O2. With hydroxylamine as the reducing agent the emitted light intensity was related to the first power of the D(OH)2 concentration and the rate of O2 uptake to the square root of the D(OH)2. The disappearance of D(OH)2 in these circumstances was by a first order or pseudo-first order rate. Fructose as a reducing agent by contrast resulted in an O2 uptake nearly independent of D(OH)2 concentration over a range from 1 × 10-5M-1 × 10-4M, while the light intensity and disappearance of D(OH)2 followed a one-half order rate. O2 uptake was by zero order kinetics and the oxidation of fructose proceeded at the same rate as was found with ferricyanide as oxidant. The kinetics, quantum yields and temperature dependence of the fructose reactions were compared with similar reactions employing H2O2 as the light eliciting reagent. The results are interpreted as indicating that D(OH)2 acts as a chain initiator in a manner analogous to better known, radical producing compounds found to accelerate hydrocarbon autooxidations.  相似文献   

8.

A suspension of sparingly soluble zinc (1-hydroxyethylidene)diphosphonate ZnH2L?2H2O (H4L = MeC(OH)[P(O)(OH)2]2) is rapidly transformed into an aqueous solution when treated with ammonia or aliphatic amines (hexamethylenediamine, triethylamine, tert-butylamine, di-n-butylamine) containing no hydrophilic groups ?OH and ?(OCH2CH2)n?. The dissolution effect is due to the decomposition of the coordination polymer giving ammonium derivatives. Dehydrated dry powders of the corresponding ammonium compounds based on triethylamine, tert-butylamine, or di-n-butylamine rapidly dissolve in water to form transparent colorless solutions, whereas hexamethylenediamine and ammonia derivatives are poorly soluble. (1-Hydroxyethylidene)diphosphonic acid forms ammonium salts with hexamethylenediamine, triethylamine, and tert-butylamine. The molecular structures of these compounds are considered.

  相似文献   

9.
A series of polymers, {Cr(OH)(OPRR′O)[OOC(CF2)nCF(CF3)2]}x has been prepared and studied. The polymers with R = R′ = C6H5 are soluble in CCl2FCClF2, whereas those with R = CH3 and R′ = C6H5 and with R = R′ = C8H17 are insoluble in all solvents. Attempts to prepare similar materials without hydroxyl groups gave the polymers {Cr(OH)r(OPRR′O)p[OOC(CF2)nCF(CF3)2]q}x with 0 < r < 1. The latter polymers are much more tractable than the former; however they are also less thermally stable. The perfluoro-carboxylate groups in these materials can either be chelating or bridging, depending on the other ligands present.  相似文献   

10.
The deactivation of I(2P½) by R-OH compounds (R = H, CnH2n+1) was studied using time-resolved atomic absorption at 206.2 nm. The second-order quenching rate constants determined for H2O, CH3OH, C2H5OH, n-C3H7OH, i-C3H7OH, n-C4H9OH, i-C4H9OH, s-C4H9OH, t-C4H9OH, are respectively, 2.4 ± 0.3 × 10−12, 5.5 ± 0.8 × 10−12, 8 ± 1 × 10−12, 10 ± 1 × 10−12, 10 ± 1 × 10−12, 11.1 ± 0.9 × 10−12, 9.8 ± 0.9 × 10−12, 7.1 ± 0.7 × 10−12, and 4.1 ± 0.4× 10−12 cm3 molec−1 s−1 at room temperature. It is believed that a quasi-resonant electronic to vibrational energy transfer mechanism accounts for most of the features of the quenching process. The influence of the alkyl group and its role in the total quenching rate is also discussed. © 1997 John Wiley & Sons, Inc.  相似文献   

11.
During an investigation of the SrO–CuO–P2O5–H2O system, single crystals of distrontium hexahydroxidocuprate(II), Sr2[Cu(OH)6], were obtained by the hydrothermal method. The blue prismatic crystals of Sr2[Cu(OH)6] adopt the same structure type as Ba2[Cu(OH)6], Sr2[Zn(OH)6] and Ba2[Zn(OH)6]. The Cu atoms, located at (0, 0, ) (site symmetry ), form mutually isolated and highly elongated Cu(OH)6 octahedra, which are interconnected to slightly distorted Sr(OH)6 trigonal prisms, forming a layered structure. The location of H atoms from difference Fourier maps and their refinement allowed the precise determination of a three‐dimensional hydrogen‐bonding network in which all hydroxide O atoms are involved. In addition, the hydrogen‐bonding topologies in Sr2[Cu(OH)6] and other similar hexahydroxidometallates with the general formulae M1[M2(OH)6], M12[M2(OH)6] and M13[M2(OH)6] were analysed in detail.  相似文献   

12.
The syntheses of m-azido-L-phenylalanine, p-azido-o-nitro-L-phenylalanine, Z · Ala-Ala-Xxx · OH, and Z · Ala-Ala(β-3H)-Xxx · OH, with Xxx – Phe, Phe(p-NO2), Phe(p-N3), Phe(m-N3), and Phe(p-N3, o-NO2) are described. The tripeptides with substituted phenylalanine residues are reversible inhibitors of chymotrypsin in the dark, and irreversible ones in the light. Photoaffinity labelling of chymotrypsin has been reported by the authors in FEBS Letters 46, 347 (1974).  相似文献   

13.
13C-NMR linewidths and spin-lattice relaxation times have been determined for soluble and crosslinked polystyrenes containing quaternary phosphonium and ammonium ions. Solubilities and NMR linewidths show that the solvating abilities toward tri-n-butylphosphonium ions are CDCl3 > CH3OH > D2O > benzene, toluene, toward the trimethylammonium ions, CH3OH > CDCl3 > D2O > toluene, and toward the nonpolar polymer backbone, CDCl3, benzene > toluene > CH3OH > D2O.  相似文献   

14.
近几年来,氢氧化镁作为一种无机阻燃剂由于其具有制备条件相对温和,生产工艺简单且产品与自然环境友好等特点,在研究及生产活动方面备受关注且得到了长足的发展[1~4].目前采用氢氧化钠法进行反应一水热制备高分散阻燃级氢氧化镁的工艺路线已经比较成熟[5~8].然而,不利的是,氢氧化钠偏高的价格导致了产品的制造成本较高.而采用石灰法制备氢氧化镁阻燃剂具有价格低廉的特点,引起了人们的关注.  相似文献   

15.
Zusammenfassung Es wurden VerbindungenLnCl3·H2 DBox * (Ln=La, Ce, Pr, Nd, Sm, Eu, Gd, Tb) und La2(DBox)3·6 CH3OH, Pr2(DBox)3·6 CH3OH, Nd2(DBox)3·3 C2H5OH, Sm2(DBox)3· ·3 C2H5OH, Gd2(DBox)3·3 C2H5OH, Tb2(DBox)3·2 C2H5OH, DyH(DBox)2·2 C2H5OH, HoH(DBox)2·2 C2H5OH, ErH(DBox)2, YH(DBox)2·H2O isoliert. Diese wurden mittels Thermoanalyse und Röntgenstreuung untersucht sowie ihre IR-Absorptionspektren gemessen und diskutiert.
Compounds of rare earth elements with -4-dimethylaminobenzoinoxime
The compoundsLnCl3·H2 DBox * (Ln=La, Ce, Pr, Nd, Sm, Eu, Gd, Tb) and La2(DBox)3·6 CH3OH, Pr2(DBox)3·6 CH3OH, Nd2(DBox)3·3 C2H5OH, Sm2(DBox)3·3 C2H5OH, Gd2(DBox)3· ·3 C2H5OH, Tb2(DBox)3·2 C2H5OH, DyH(DBox)2·2 C2H5OH, HoH(DBox)2·2 C2H5OH, ErH(DBox)2 and YH(DBox)2·H2O were isolated, and studied by thermoanalysis and X-ray diffraction. Their IR absorption spectra were recorded and are discussed.
  相似文献   

16.
The violet superoxo complex, [(H2O)4(OH)RhIII(O2)RhIII(OH)(H2O)4]3+, formed by treatment of (RhII)24+ with O2 in HClO4, is converted to a le? reduction product, the corresponding μ-peroxo complex, by the reductants I?, IrCl63?, and the trinuclear aquamolybdenum(III) cation, (MoIII)3. Each reaction is first-order in both redox partners, and the le? reduction by IrCl63? is followed by a much slower conversion to a peroxide-free complex. Among the rapid reductions of the superoxo derivative examined here and in a previous study, only that by IrCl63? is accelerated by increases in acidity; the rate law for this reaction features both an acid-independent and a [H+]-proportional component, the latter stemming from partial conversion of the oxidant to its conjugate acid (pKA < ?1.0). Rate laws for reductions by other metal-center reagents generally exhibit inverse-[H+] terms, reflecting deprotonation of the reductant. All reductions thus far observed involving this superoxo species appear to be outer-sphere. Treatment of acid-independent rate constants within the framework of the Marcus model, allows estimates of the self-exchange rate, k11, for the (RhIII)2-bound superoxo-peroxo couple. Because values of k11 calculated from the several reductions span a range of 104.5, reductions of the superoxo complex cannot be taken to conform satisfactorily to the Marcus treatment, being in this respect comparable to the systems VO(OH)+/2+, Mn2+/3+, Eu2+/3+, and Ti(OH)2+/3+, each of which exhibits similar divergences. The wide range of calculated self-exchange rates appears to invalidate an earlier suggestion that reduction of the superoxo complex by Fe2+ proceeds primarily through a bridged path. © 1994 John Wiley & Sons, Inc.  相似文献   

17.
《Electroanalysis》2006,18(7):719-729
The interaction between cadmium or zinc and AMPSO was investigated by DCP and GEP, at fixed total ligand to total metal concentration ratios and various pH values, at 25.0 °C and 0.1 M KNO3 ionic strength. For Cd–(AMPSO)x–(OH)y system, CdL and CdL(OH) species, were identified, with stability constants values set to (as log β): 2.1±0.1 and 6.2±0.2, respectively. For Zn–(AMPSO)x–(OH)y system, the proposed final model with stability constants set to (as log β) is: ZnL=2.5±0.1 and ZnL(OH)2=12.9±0.2. For both systems, the fact that AMPSO deprotonation occurs in the metal hydrolysis and M(OH)2 precipitation and the complexes formed are not too strong added a real challenge to data interpretation.  相似文献   

18.
The optical transient and kinetics characterizations of the transients formed in the reaction of OH with benzotrifluoride (BTF) were performed by a laser flash photolysis technique. The results indicated that the formation of π‐type adduct of C6H5(OH)CF3 was the major reaction channel, and the δ‐type adduct of C6H5CF3OH formation was an additional minor process in the oxidation reaction of BTF attacked by OH radicals yielded from the photolysis of H2O2. Addition of OH to the CF3 group led to the fluoride ion elimination to yield α,α‐difluorophenylcarbinol (C6H5CF2OH). Trifluoromethylphenol (HOC6H4CF3) of meta‐, para‐ and ortho‐substituted isomers resulted from the addition of OH to the BTF aromatic ring.  相似文献   

19.
When di-2-pyridyl ketone (dpk) was allowed to react with [Mo(CO)6] in toluene under reflux in air [Mo(O)2(μ-O)(η 3-dpkO,OH)]2 where η 3-dpkO,OH is N,O,N-hydroxybis(2-pyridyl)methanolato was isolated. Infrared and 1H-NMR spectral data measured on solutions of the isolated compound indicate the absence of the carbonyl groups and the presence of coordinated η 3-dpkO,OH, terminal and bridging oxo-groups and elemental analysis confirmed the formulation of the product as [Mo(O)2(μ-O)(η 3-dpkO,OH)]2. Crystals of [Mo(O)2(μ-O)(η 3-dpkO,OH)]2·2dmso obtained from a dimethyl sulfoxide (dmso) solution of [Mo(O)2(μ-O)(η 3-dpkO,OH)]2 are in the monoclinic P21/c space group. X-ray structural analysis confirmed the identity of [Mo(O)2(μ-O)(η 3-dpkO,OH)]2 and shows a dioxomolybdenum oxo-bridged dimer with two terminal oxygen atoms and one tridentate N,O,N-dpkO,OH occupying the coordination sites of the high-valent molybdenum(VI) in a pseudo-octahedral coordination geometry. The molecular packing shows parallel stacks of [Mo(O)2(μ-O) (η 3-dpkO,OH)]2·2dmso and disclose an extensive network of non-covalent interactions within each stack.  相似文献   

20.
Summary Substitution of the (CH3)2CH=CH group by (CH3)2C=N in linalool as well as by the (CH3)2CH-CH2 group in linalool and in sila-linalool does not lead to noticeable changes of their scent qualities. On the contrary, substitution of the OH group at the tertiary C atom by NH2 or CH3 — hydride isosteric to OH according toGrimm — affords fishy or etheric instead of the original flowery smells thus indicating a transition to different basic classes of odor. Similar results were obtained with the linalool-like scents of benzyldimethylcarbinol and phenylethyldimethylcarbinol. Therefore the theory ofAmoore, after which only shape and size of molecules are ruling their odor qualities, must be called in question.
  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