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.
The coupling reaction of polyisoprenyllithium with 1,2-dibromoethane (DBE) gives a multimodal molecular weight distribution (MWD). Species with molecular weights three and four times (P3, P4) higher than the base polymer (P) are present beside the main expected one with double molecular weight (P2). The relative abundancies of P1, P2, P3, and P4 depend on the experimental conditions. Gas-phase chromatographic analysis carried out during the coupling reaction show the presence of ethylene whose formation is related to a lithium–bromine exchange reaction competing with direct alkylation. The lithium–bromine exchange reaction is more effective at T < 80°C and results in the formation of allyl bromide-terminated polyisoprene, while direct alkylation is effective at T > 100°C and yields alkyl bromide chain ends. The allylic and alkylic bromides react differently with the remaining polyisoprenyllithium: the former adds only to polyisoprenyllithium yielding P2, while the later also yields P3 and P4 through radical reactions. © 1997 John Wiley & Sons, Inc.  相似文献   

2.
Co2(CO)8 and Me2P(S)P(S)Me2 react to form the two cluster complexes: Co4(CO)9S(PMe2)2) (1) and Co3(CO)7S(SPMe2) (2). The strucure of1 and of the disubstituted triphenyl phosphine derivative of2. Co3(CO)5(PPh3)2S (SPMe3) (2a) were determined. Compound1 contains a quasi-planar rhomboidal Co4 cluster formed by two Co3 isosceles triangles sharing a Co-Co edge. One triangle is capped by a sulfur atom, the other triangle has two edge-bridging PMe2 moieties. Electron counting gives 64 electrons corresponding to a planar system; the distribution of long Co-Co distances, in particular in the triangle bearing PMe2 bridges, suggests that the excess electrons are located on Co-Co antibonding ortibals. Compound2a contains a Co3S cluster with one side bridged by a SPMe2 unit forming a four-membered Co2SP ring. The substitution of two CO groups with two PPh3 causes a large deformation of the cluster Co-Co bondscis to these two phosphorus atoms. Crystal data for1, space group P1,a = 9.728(2) Å,b = 10.288(2) Å,c = 11.860(3) Å, = 86.41(2)°, = 76.20(2)°, = 80.37(5)°,Z = 2, 5300 reflections,R = 0.0398; for2a, space group P1,a = 9.78(3) Å,b = 13.05(4) Å,c = 18.28(6) Å, = 93.23(3)°, = 99.17(2)°, = 97.26(6)°,Z = 2, 2976 reflections,R = 0.0579.  相似文献   

3.
Reactions of a number of nitriles with camphene in the presence of the heteropoly acids H3PW12O40, H7PMo12O42, and H4SiW12O40 as catalysts were studied. In all cases, N-substituted amides were obtained in sufficiently high yields. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 2, pp. 374–376, February, 2006.  相似文献   

4.
The concept of a reaction network, initially suggested by Ugi and coworkers, in the framework of the graph-theoretical model of organic chemistry is elaborated. The reaction network for a pair of isomeric educt molecular (G E) and product molecular graphs (G P) is determined as an oriented graph. Its edge, beginning at a graph-vertexG i –1 and ending at a graph-vertexG i , corresponds to a feasible transformation (chemical reaction) constrained by a condition of descending chemical distance from the product graphG P, i.e.CD(G i –1,G P) >CD(G i ,G P). In the reaction network, an oriented path which begins at GE and ends atG P corresponds to the decomposition of the overall transformationG E G P into a sequence of elementary transformationsG 0 =G E G 1 G 2 ... G i–1 G i =G P that may be assigned to intermediates of the overall transformation.  相似文献   

5.
With a newly developed analytical technique, i.e. high temperature/pressure IR cell coupled to the reactor, it was possible to study the mechanism of hydroformylation at reaction conditions. It has been conclusively found that the hydrogenolysis of the acyl cobalt complex is performed by HCo(CO)4 and not by molecular H2, as proposed byHeck andBreslow.Therefore the formation of HCo(CO)4 from Co2(CO)8 is an intermediate step in the sequence of hydroformylation reaction steps. The rate of hydroformylation of any of the olefins is smaller than the rate of formation of HCo(CO)4 from Co2(CO)8. The IR spectra reveal that always more than 30% of the cobalt is in the form of HCo(CO)4 under the reaction conditions.It is found that the formation of HCo(CO)4 from Co2(CO)8 is the slowest and most temperature-dependent step of the hydroformylation reaction. Also the reaction between olefin and HCo(CO)4 is slower than the hydrogenolysis of the acyl complex.The experiments were carried out under industrial oxo conditions. The diffusional effects were eliminated.With 6 FiguresPart of the Ph.D. dissertation 1974. N. H. Alemdarolu, J. M. L. Penninger, andE. Oltay, Mechanism of Hydroformylation, Part II. Mh. Chem.107, 1043 (1976).  相似文献   

6.
Reaction data described by the second-order growth function A(t) = At) (1 + αt)?1, where A is the ultimate value of the product concentration A(t), can be linearized by plotting a suitable function F(t) against the time (t). The slope of the straight line obtained is (2α), where α is the product of the rate constant (k2) and the initial concentration of either reactant, with the result that k2 can be determined without knowledge of A?. Optimal determination of the parameter α requires that data taking be limited to the interval 0 ≤ tT, where (αT) is approximately 4.0. Numerical data derived from an experiment on the exchange of lead by zinc ions in the enzyme carbonic anhydrase are analyzed to illustrate the method. The effects of small errors in the initial concentrations and of small deviations from second-order kinetics are briefly discussed.  相似文献   

7.
Rate constants and activation parameters for the isotopic exchange reactions between (PhO)2PSCl and M36Cl (M = Me4N+, Et4N+, n-Bu4N+, Et3HN+, EtH3N+, Li+) in acetonitrile were measured in order to find the effect of the cation nature onthe kinetics of the reaction. The rate constants measured for a range of concentrations of Et3HN36Cl, EtH3N36Cl, and Li36Cl were analyzed using the Acree equation. The equivalent conductance of LiCl in acetonitrile was determined. The nature of the cation has no effect on the mechanism of the reaction. The cation changes only the experimental rate constant proportionally to the dissociation degree of the salt. Smaller values of the rate constant and smaller activation parameters ΔH? and ΔS? for the reaction with Li36Cl indicate the existenceof the intermolecular interaction between lithium ions and O,O-diphenylphosphorochloridothionate.  相似文献   

8.
A closed oscillation system comprised of alanine, KBrO3, H2SO4 and acetone catalyzed by tetraazamacrocyclic nickel(II) complex is introduced, and quantitatively characterized with kinetic parameters, namely the rate constant (k in, k p), the apparent activation energy (E in, E p) and pre-exponential constant (A in, A p) and thermodynamic functions (ΔH in, ΔG in, ΔS in and ΔH p, ΔG p, ΔS p), where indexes “in” and “p” mean “induction period” and “oscillation period,” respectively. The results indicate that tetraazamacrocyclic nickel(II) complex can catalyze alanine oscillating reaction and the reaction corresponds exactly to the feature of irreversible thermodynamics as the entropy of system is negative.  相似文献   

9.
Cationic micelles of alkyltrimethylammonium chloride and bromide (alkyl = n? C12H25, n? C14H29, and n? C16H33) catalyze and anionic micelles of sodium dodecyl sulfate inhibit the reaction of hydroxide ion with 2-phenoxyquinoxaline (1). Inert anions such as chloride, nitrate, mesylate, and n-butanosulfonate inhibit the reaction in CTABr by competing with OH? at the micellar surface. The overall micellar effects on rate in cationic micelles and dilute electrolyte can be treated quantitatively in terms of the pseudo-phase ion-exchange model. The determined second-order rate constants in the micellar pseudo-phase are smaller than the second-order constants in water. © 1994 John Wiley & Sons, Inc.  相似文献   

10.
The reactivity of carbonyl oxides toward benzaldehyde was characterized by thek 33/k 33 ratio, wherek 33 andk 31 are the rate constants of the reactions of RCOO with PhCHO and diphenyldiazomethane Ph2CN2, respectively. Thek 33/k 31 ratios obtained at 60°C in acetonitrile range from 0.61·10−2 (m-BrPh2CN2) to 20·10−2 (Ph2MeCHO). The reactions are probably preceded by the formation of a charge-transfer complex (CTC) with charge transfer from aldehyde to RCOO. The carbonyl oxide reacts with aldehydes by both the nucleophilic pathway (at the C atom of the—CHO group to form 1,2,4-trioxolane) and electrophilic pathway (by the attack at the aromatic ring with the intermediate formation of CTC). In the latter case, either 1,2,4-trioxolane or oxidation products of the phenyl ring are formed. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 4, pp. 650–654, April, 2000.  相似文献   

11.
A vacuum ultraviolet photolysis of C2H5Br at 147 nm was studied over a pressure range of 0.5–50 torr at 298 K. The effects of additives He and NO were also investigated. The principal reaction products were found to be C2H4 and C2H6, with lesser yields of CH4 and C2H2. With increasing pressure the product quantum yields Φi of C2H4, CH4, and CH2H6 remained constant, while that of C2H2 decreased from 0.03 to almost 0. The effect of He as an additive was found to be extremely small on the quantum yields of the major products. Addition of NO completely suppresses the formation of CH4, C2H2, and C2H6, and reduces partially the production of C2H4. The primary processes appear to involve two electronically excited states. One state mainly yields C2H4 by molecular elimination of HBr and is thought to be due to a Rydberg transition. The other state decomposes to C2H5 and Br radicals by C? Br bond fission. These two competitive reaction modes contribute to the photodecomposition in proportions of 50% and 50%. The extinction coefficient for C2H5Br at 147 nm and at 298 K has been determined as ? = (1/PL) In(Io/It) = 712 ± 7 atm?1 · cm?1.  相似文献   

12.
The equilibrium constant for the reaction CH2(COOH)2 + I3? ? CHI(COOH)2 + 2I? + H+, measured spectrophotometrically at 25°C and ionic strength 1.00M (NaClO4), is (2.79 ± 0.48) × 10?4M2. Stopped-flow kinetic measurements at 25°C and ionic strength 1.00M with [H+] = (2.09-95.0) × 10?3M and [I?] = (1.23-26.1) × 10?3M indicate that the rate of the forward reaction is given by (k1[I2] + k3[I3?]) [HOOCCH2COO?] + (k2[I2] + k4[I3?]) [CH(COOH)2] + k5[H+] [I3?] [CH2(COOH)2]. The values of the rate constants k1-k5 are (1.21 ± 0.31) × 102, (2.41 ± 0.15) × 101, (1.16 ± 0.33) × 101, (8.7 ± 4.5) × 10?1M?1·sec?1, and (3.20 ± 0.56) × 101M?2·sec?1, respectively. The rate of enolization of malonic acid, measured by the bromine scavenging technique, is given by ken[CH2(COOH)2], with ken = 2.0 × 10?3 + 1.0 × 10?2 [CH2(COOH)2]. An intramolecular mechanism, featuring a six-member cyclic transition state, is postulated to account for the results on the enolization of malonic acid. The reactions of the enol, enolate ion, and protonated enol with iodine and/or triodide ion are proposed to account for the various rate terms.  相似文献   

13.
1,4-Diphenyl-2,3-dioxabicyclo[2.2.1]hept-5-ene ( 2 ), on treatment with a catalytic amount of trimethylsilyl trifluoromethanesulfonate (Me3SiOTf) in CH2Cl2 at ?78°, reacts with excess (?)-menthone ( 10 ) to give (1S,2S,4′aS,5R,7′aS)-4′a,7′a-dihydro-2-isopropyl-5-methyl-6′,7′-diphenylspiro[cyclohexane-1,3′-[7′H]cyclopenta-[1,2,4]trioxine] ( 11 ) and its (1R,2S,4′aR,5R,7′aR)-diastereoisomer 12 in a 1:1 ratio and in 21% yield. Repeating the reaction with 1.1 equiv. of Me3SiOTf with respect to 2 affords 11 , 12 , and (1S,2S,3′a.R,5R,6′aS)-3′a,6′a-dihydro-2-isopropyl-5-methyl-3′a-phenoxy-5′-phenylspiro[cyclohexane-l,2′-[4′H]cyclopenta[1,3]dioxole] ( 13 ) together with its(1R,2S,3′aS,5R,6′aR)-diastereoisomer 14 in a ratio of 3:3:3:1 and in 56% yield. (+)-Nopinone( 15 ) in excess reacts with 2 in the presence of 1.1 equiv. of Me3SiOTf to give a pair of 1,2,4-trioxanes ( 16 and 17 ) analogous to 11 and 12 , and a pair of 1,3-dioxolanes ( 18 and 19 ) analogous to 13 and 14 , in a ratio of 8:2:3:3 and in 85% yield. (?)-Carvone and racemic 2-(tert-butyl)cyclohexanone under the same conditions behave like 15 and deliver pairs of diastereoisomeric trioxanes and dioxolanes. In general, catalytic amounts of Me3SiOTf give rise to trioxanes, whereas 1.5 equiv. overwhelmingly engender dioxolanes. Adamantan-2-one combines with 2 giving only (4′aRS,7′aRS)-4′a,7′a-dihydro-6′.7′a-diphenylspiro[adamantane-2,3′-[7′H]cyclopenta[1,2,4]trioxine] in 98% yield regardless of the amount of Me3SiOTf used. The reaction of 1,4-dipheny 1-2,3-dioxabicyclo[2.2.2]oct-5-ene ( 32 ) with 10 and 1.1 equiv. of Me3SiOTf produces only the pair of trioxanes 33 and 34 homologous to 11 and 12 . Treatment of the (S,S)-diastereoisomer 33 with Zn and AcOH furnishes (1S,2S)-1,4-diphenylcyclohex-3-ene-1,2-diol. The crystal structures of 11 – 13 and 16 are obtained by X-ray analysis. The reaction courses of 10 and the other chiral cyclohexanones with prochiral endoperoxides 2 and 32 to give trioxanes are rationalized in terms of the respective enantiomeric silylperoxy cations which are completely differentiated by the si and re faces of the ketone function. The origin of the 1,3-dioxolanes is ascribed to 1,2 rearrangement of the corresponding trioxanes, which occurs with retention of configuration of the angular substituent.  相似文献   

14.
The studies on reaction of newly obtained aminophosphonic acid diethyl ester derivatives of fluorene with Cr2O7 2–, CrO4 2–, CrO3Cl and CrO3 have been investigated using electronic, infrared,Raman and NMR spectral methods. It has been found that the resulting compounds are of the type (AH)2Cr2O7, whereA stands for the organic part of the molecule. The organic cation and solvent effects on the electronic states of pseudotetrahedrally arranged Cr(VI) anions are discussed.
Spektroskopische Untersuchungen über Reaktionen von Chrom(VI)-Verbindungen mit Aminophosphonsäure-Estern
Zusammenfassung Es wurden Elektronen-, Ultrarot-,Raman- und NMR-spektroskopische Untersuchungen beschrieben, die an neuen Verbindungen durchgeführt wurden, welche als Reaktionsprodukte von Cr2O7 2–, CrO4 2–, CrO3Cl und CrO3 mit Aminophosphonsäurediethylesterderivaten von Fluorene synthetisiert worden waren. Es wurde festgestellt, daß diese Verbindungen mit einer allgemeinen Formel (AH)2Cr2O7 beschrieben können werden, in derA den organischen Teil der Verbindungen bedeutet. Der Einfluß des organischen Kations wie auch der von Lösungsmitteln auf die Elektronenzustände des pseudotetraedrischen Cr(VI)-Anions wurde gleichfalls untersucht.
  相似文献   

15.
The reaction between formic acid and HOBr in strongly acid aqueous media was studied by absorption spectrophotometry at 298 K. Bromine, the monitored species, displays a transient behavior, gradually rising up to a maximum and then decaying with first-order kinetics. Reaction rates, expressed as R = -dCBr2/dt, depend on the concentrations of HCOOH (0.1–1.4M), HOBr (0.3–1.5 × 10?3 M), and H+ (0.1–1.0M). The mechanism with k1 =1.04 ± 0.20M?1· s?1, k3 = 20.2M?1, quantitatively accounts for all observations within experimental error.  相似文献   

16.
The third order rate coefficients for the addition reaction of Cl with NO2, Cl + NO2 + M → ClNO2 (ClONO) + M; k1, were measured to be k1(He) = (7.5 ± 1.1) × 10?31 cm6 molecule?2 s?1 and k1(N2) = (16.6 ± 3.0) × 10?31 cm6 molecule?2 s?1 at 298 K using the flash photolysis-resonance fluorescence method. The pressure range of the study was 15 to 500 torr He and 19 to 200 torr N2. The temperature dependence of the third order rate coefficients were also measured between 240 and 350 K. The 298 K results are compared with those from previous low pressure studies.  相似文献   

17.
Laser flash photolysis coupled with resonance fluorescence detection of Br atoms was employed to investigate the temperature dependence of the reaction Br + neo‐C5H12 (1) between 688 and 775 K. The following Arrhenius preexponential factor and activation energy were determined (±1 σ): A1 = (6.89 ± 2.27) 1014 cm3 mol−1 s−1 and EA,1 = 57.61 ± 2.05 kJ mol1 The only other kinetic parameters reported for the reaction of Br atoms with neo‐C5H12 were obtained from competitive kinetic experiments relative to Br + C2H6. Comparison with our direct results is hampered by uncertainties in the kinetic data for the reference reaction that may need reinvestigation. The standard enthalpy of formation for the neo‐C5H11 radical was estimated to be 34.7 and 41.6 kJ mol−1, depending on the value of the activation energy assumed for the reverse reaction neo‐C5H11 + HBr (−1). © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 33: 49–55, 2001  相似文献   

18.
The bimolecular reaction of HI in CO2, which was excited vibrationally by irradiation of a continuous-wave CO2 laser light, was investigated in the temperature range of 721–980 K. An enhancement of the reaction rate by a factor of about 2.5 was observed in the 1:1 HI? CO2 mixture in comparison with the rate in pure HI when both sample gases were irradiated by a CO2 laser (50 W) at 1 torr. However, in the HI-SF6 mixtures the decomposition rate of HI was not accelerated by irradiation of the CO2 laser. Thus the enhancement is attributed to vibrational excitation of HI through collisional energy transfers from laser-excited CO2 (00°1). At lower total pressures or at lower partial pressures of HI in HI-CO2 mixtures the enhancement was more significant because of inefficient vibrational deactivation of excited HI. A model calculation gave the result in agreement with the experimental one if the effective activation energy is assumed as Ea? = Ea - αEvib, where Ea is the activation energy for the thermal reaction, Evib is the vibrational energy of two colliding HI molecules, and α is estimated to be about 0.7. This means that a part of the vibrational energy of reacting HI is employed to reduce the activation energy for the translational or rotational degree of freedom.  相似文献   

19.
The reaction of 1,2,5,6-tetrathiacyclooctane with Ru3(CO)12 in methylene chloride solvent at 40°C has yielded two new isomericbis-thane-1,2-dithiolate triruthenium carbonyl cluster complexes:anti-Ru3(CO)7(-SCH2CH2S)2, 1 andsyn-Ru3(CO)7(-SCH2CH2S)2,2, and the previously reported diruthenium compound, Ru2(CO)6(-SCH2CH2S).3 in 24 %, 5 %, and 26 % yields, respectively. Compounds1 and2 were characterized by a single crystal X-ray diflraction analyses. Both compounds consist of a open triangular triruthenium clusters with seven terminal carbonyl ligands and a bridging ethanedithiolate ligand across each of the metal metal bonds in the complex. When heated to 60° C, compound1 was trans[formed into a mixture of2 and3. Crystallographic data for1: Ru3S4O7C11H8, space group, P21/a,a= ll.830(2)A,b= 10.576(1)A,c= 16.012(1)A,= 100.53(2)°,Z=4, 1808 reflections,R= 0.029. For2: Ru3 S4O7C11H8, space group P1,a= 9.945(l)A,b= 11.323(1)A.c= 9,788(1)A,a= 108.73(1)°,= 104,67(1)°,y= 103.59(2)°,Z = 2, 2046 rellections.R = 0.021.  相似文献   

20.
A laser flash photolysis-long path absorption technique has been employed to study the kinetics of the reaction products as a function of temperature (248–346 K), pressure (16–800 torr), and buffer gas identity (N2, CF4). The reaction is found to be in the falloff regime between third and second-order over the entire range of conditions investigated. This is the first study where temperature-dependent measurements of k1(P, T) have been reported at pressures greater than 12 torr; hence, our results help constrain choices of k1(P, T) for use in models of lower stratospheric BrOx chemistry. Approximate falloff parameters in a convenient form for atmospheric modeling are derived. © 1993 John Wiley & Sons, Inc.  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