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.
A rapid and simple method for determining the plant growth regulator hexanoic acid 2-(diethylamino) ethyl ester (DA-6) in pakchoi and soil using gas chromatography-mass spectrometry (GC-MS) has been developed. For this purpose, a single step was used to extract DA-6 with dichloromethane from aqueous-acetone extracts of vegetables and soil. Average recoveries of DA-6 in pakchoi and soil were between 85% and 104% at both spiking levels 0.01 and 0.1 mg kg?1. Relative standard deviations (RSD) were less than 11% for all of the recovery tests. The degradation of DA-6 in pakchoi and soil was studied. The results showed that DA-6 degradation in pakchoi and soil coincided with C = 3.9903 e?0.0516 t , C = 0.3476 e?0.0224 t , respectively; the half-lives were 13.43 h and 30.94 h in pakchoi and soil in Beijing, respectively.  相似文献   

2.
A method for determining validamycin A residue in soil utilizing pre-column derivatization and gas chromatography was developed. Validamycin A was extracted from soil samples by 10% ammonia solution without additional clean-up step and then was derivatized with N,O-bis(trimethylsilyl)acetamide. Derivative of validamycin A was analyzed by gas chromatography on a capillary column. The method showed good linearity in the range of 4.4–142.0 mg/L (r 2 = 0.9997). The average recoveries were 98.82 and 98.61% with relative standard deviations 1.95 and 2.87% at spiking levels 0.2 and 2.0 mg/kg, respectively. The limit of detection of validamycin A in soil was 0.2 mg/kg. The degradation dynamics of validamycin A in soil was studied. The result shows that validamycin A degradation in soil in the open field in Beijing suburb can be described by the equation c = 6.6109e −0.0196t ; the half-life is 35.4 h. The article is published in the original.  相似文献   

3.
Chemistry of Polyfunctional Molecules. 119 [1]. Tetracarbonyl-dicobalt-tetrahedrane Complexes with the Ligands Bis(diphenylphosphanyl)-amine, 2-Butin-1,4-diol, and tert.-Butylphosphaacetylene — Crystal Structure of the Phosphaalkyne Derivative Co2(μ-CO)2(CO)4(μ-Ph2P? NH? PPh2P,P′) · 1/2C6H5CH3 ( 4 · 1/2C6H5CH3) reacts with 2-butine-1,4-diol, HOCH2? C?C? CH2OH ( 5 ), to the dark-red tetrahedrane complex Co2(CO)4(μ-η22-HOCH2? C?C? CH2OH? C2, C3) · (μ-Ph2P? NH? PPh2? P,P′) · THF (6 · THF). With t-butyl-phosphaacetylene, tBu? C?P ( 7 ), 4 · THF forms Co2(CO)4(μ-η22-tBu? C?P)(μ-Ph2P? NH? PPh2? P,P′) ( 8 ), which also belongs to the tetrahydrane type. The compounds were characterized by their mass, IR, 31P{1H} NMR, 13C{1H} NMR, and1H NMR spectra. Crystals suitable for X-ray structure analyses have been obtained for 8 from dioxane. The dark red blocks crystallize in the monoclinic P21/c space group with the lattice constants a = 1404,1(5), b = 1330,0(7), c = 2578,8(10)pm; β = 90,82(3)°.  相似文献   

4.
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   

5.
[NMe4]2[TCNE]2 (TCNE=tetracyanoethenide) formed from the reaction of TCNE and (NMe4)CN in MeCN has νCN IR absorptions at 2195, 2191, 2172, and 2156 cm?1 and a νCC absorption at 1383 cm?1 that are characteristic of reduced TCNE. The TCNEs have an average central C?C distance of 1.423 Å that is also characteristic of reduced TCNE. The reduced TCNE forms a previously unknown non‐eclipsed, centrosymmetric π‐[TCNE]22? dimer with nominal C2 symmetry, 12 sub van der Waals interatomic contacts <3.3 Å, a central intradimer separation of 3.039(3) Å, and comparable intradimer C???N distances of 3.050(3) and 2.984(3) Å. The two pairs of central C???C atoms form a ?C?C???C?C of 112.6° that is substantially greater than the 0° observed for the eclipsed D2h π‐[TCNE]22? dimer possessing a two‐electron, four‐center (2e?/4c) bond with two C???C components from a molecular orbital (MO) analysis. A MO study combining CAS(2,2)/MRMP2/cc‐pVTZ and atoms‐in‐molecules (AIM) calculations indicates that the non‐eclipsed, C2 π‐[TCNE]22? dimer exhibits a new type of a long, intradimer bond involving one strong C???C and two weak C???N components, that is, a 2e?/6c bond. The C2 π‐[TCNE]22? conformer has a singlet, diamagnetic ground state with a thermally populated triplet excited state with J/kB=1000 K (700 cm?1; 86.8 meV; 2.00 kcal mol?1; H=?2 JSa?Sb); at the CAS(2,2)/MBMP2 level the triplet is computed to be 9.0 kcal mol?1 higher in energy than the closed‐shell singlet ground state. The results from CAS(2,2)/NEVPT2/cc‐pVTZ calculations indicate that the C2 and D2h conformers have two different local metastable minima with the C2 conformer being 1.3 kcal mol?1 less stable. The different natures of the C2 and D2h conformers are also noted from the results of valence bond (VB) qualitative diagram that shows a 10e?/6c bond with one C???C and two C???N bonding components for the C2 conformer as compared to the 6e?/4c bond for the D2h conformer with two C???C bonding components.  相似文献   

6.
This study describes the application of the electrochemically generated molybdenum‐based catalyst system MoCl5? e?? Al? CH2Cl2 to ring‐opening metathesis polymerization of bicyclo[2.2.1]hept‐2‐ene (norbornene). The results are compared with those previously obtained by the WCl6? e?? Al? CH2Cl2 system. The polymer product has been characterized by 1H and 13C NMR, IR and gel‐permeation chromatography techniques. This molybdenum‐based catalyst system has led to a mainly trans stereoconfiguration (ca 60%) of the double bonds, in contrast to the polymer obtained with the tungsten‐based analogue, where the cis content is 60%. Analysis of the poly(1,3‐cyclopentylenevinylene) microstructure by 13C NMR spectroscopy revealed that the polymer having σc = 0.41 (fraction of double bonds with cis configuration) contains a slightly blocky distribution (rtrc > 1) of the double‐bond dyads (rtrc = 1.44). In addition, the influence of reaction parameters, e.g. reaction time, electrolysis time and catalyst aging time, on conversion has been analysed in detail. Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

7.
A series of zinc(II) silylenes was prepared by using the silylene {PhC(NtBu)2}(C5Me5)Si. Whereas reaction of the silylene with ZnX2 (X=Cl, I) gave the halide‐bridged dimers [{PhC(NtBu)2}(C5Me5)SiZnX(μ‐X)]2, with ZnR2 (R=Ph, Et, C6F5) as reagent the monomers [{PhC(NtBu)2}(C5Me5)SiZnR2] were obtained. The stability of the complexes and the Zn?Si bond lengths clearly depend on the substitution pattern of the zinc atom. Electron‐withdrawing groups stabilize these adducts, whereas electron‐donating groups destabilize them. This could be rationalized by quantum chemical calculations. Two different bonding modes in these molecules were identified, which are responsible for the differences in reactivity: 1) strong polar Zn?Si single bonds with short Zn?Si distances, Zn?Si force constants close to that of a classical single bond, and strong binding energy (ca. 2.39 Å, 1.33 mdyn Å?1, and 200 kJ mol?1), which suggest an ion pair consisting of a silyl cation with a Zn?Si single bond; 2) relatively weak donor–acceptor Zn?Si bonds with long Zn?Si distances, low Zn?Si force constants, and weak binding energy (ca. 2.49 Å, 0.89 mdyn Å?1, and 115 kJ mol?1), which can be interpreted as a silylene–zinc adduct.  相似文献   

8.
In contrast to previous reports e?t in 3MHx produced at 4 K shows trap relaxation when warmed to 77 K. The relaxation in 3 MHx and 3MP occurs without loss of e?t, and results in an increase in ?max as well as a shift of λmax.  相似文献   

9.
A new method is described for the determination of lead based on the cathodic adsorptive stripping of the lead–nuclear fast red (NFR) at a carbon paste electrode (CPE). The differential pulse voltammograms of the adsorbed complex of lead–NFR are recorded from ?0.10 to ?0.60 V (versus Ag/AgCl electrode). Optimal conditions were found to be an electrode containing 25% paraffin oil and 75% high purity graphite powder, 4.0×10?5 mol L?1 NFR; buffer solution (pH of 3.0), accumulation potential and time, ?0.20 V, 60 and 120 s (for high and low concentration of lead), respectively. The results show that the complex can be adsorbed on the surface of the CPE, yielding one peak at ?0.34 V, corresponding to reduction of NFR in the complex at the electrode. The detection limit was found to be 0.2 ng mL?1 with a 120s accumulation time. The linear ranges are from 0.5 to 50 (tacc=120 s) and 50 to 200 ng mL?1 (tacc=60 s). Application of the procedure to the determination of lead in lake water, bottled mineral water, synthetic samples and sweet fruit‐flavored powder drinks samples gave good results.  相似文献   

10.
The first two persistent silenyl radicals (R2C=Si.?R), with a half‐life (t1/2) of about 30 min, were generated and characterized by electron paramagnetic resonance (EPR) spectroscopy. The large hyperfine coupling constants (hfccs) (a(29Siα)=137.5–148.0 G) indicate that the unpaired electron has substantial s character. DFT calculations, which are in good agreement with the experimentally observed hfccs, predict a strongly bent structure (?C=Si?R=134.7–140.7°). In contrast, the analogous vinyl radical, R2C=C.?R (t1/2≈3 h), exhibits a small hfcc (a(13Cα)=26.6 G) and has a nearly linear geometry (?C=C?R=168.7°).  相似文献   

11.
A thermotropic main-chain polyether based on bis(4-hydroxyphenoxy)-p-xylylene and 1,11-dibromoundecane has been studied by variable-temperature solid-state 13C NMR. Between isotropization and glass transition temperatures, the material can be identified to be semicrystalline, consisting of two conformationally and motionally distinguishable phases. The more mobile component is liquid-like and thus, can be attributed to an amorphous phase. In the more rigid component, the molecules have a conformationally disordered methylene sequence. In the low-temperature ordered phase approximately ? of the carbon-carbon bonds are trans (t). Starting from the bond between the oxygen and the first methylene carbon, the bond conformations are: d? t? d? t? t? t? t? t? t? d? t? d, where d stands for disordered (i.e., it represents the common dynamic interchange between gauche and trans with an overall gauche content of perhaps 40%). The motion of the αα′-diphenoxy-p-xylylene unit consists mainly of 180° ring-flips, which cause no entropy increase relative to ordered phenylene groups in a crystal, but significantly changes the 13C NMR spectra. The central p-xylylene ring starts its flipping motion at a lower temperature than the two phenoxy rings. The high-temperature mesophase contains a methylene sequence of the bond conformations: d? t? d? d? d? t? t? d? d? d? t? d. Thus, the difference between the low-temperature and high-temperature mesophases consists of different degrees of conformational disorder. Thermal analysis seems to indicate that additional mesophases may be possible. © 1994 John Wiley & Sons, Inc.  相似文献   

12.
X‐ray crystal structure analysis of the lithiated allylic α‐sulfonyl carbanions [CH2?CHC(Me)SO2Ph]Li ? diglyme, [cC6H8SO2tBu]Li ? PMDETA and [cC7H10SO2tBu]Li ? PMDETA showed dimeric and monomeric CIPs, having nearly planar anionic C atoms, only O?Li bonds, almost planar allylic units with strong C?C bond length alternation and the s‐trans conformation around C1?C2. They adopt a C1?S conformation, which is similar to the one generally found for alkyl and aryl substituted α‐sulfonyl carbanions. Cryoscopy of [EtCH?CHC(Et)SO2tBu]Li in THF at 164 K revealed an equilibrium between monomers and dimers in a ratio of 83:17, which is similar to the one found by low temperature NMR spectroscopy. According to NMR spectroscopy the lone‐pair orbital at C1 strongly interacts with the C?C double bond. Low temperature 6Li,1H NOE experiments of [EtCH?CHC(Et)SO2tBu]Li in THF point to an equilibrium between monomeric CIPs having only O?Li bonds and CIPs having both O?Li and C1?Li bonds. Ab initio calculation of [MeCH?CHC(Me)SO2Me]Li ? (Me2O)2 gave three isomeric CIPs having the s‐trans conformation and three isomeric CIPs having the s‐cis conformation around the C1?C2 bond. All s‐trans isomers are more stable than the s‐cis isomers. At all levels of theory the s‐trans isomer having O?Li and C1?Li bonds is the most stable one followed by the isomer which has two O?Li bonds. The allylic unit of the C,O,Li isomer shows strong bond length alternation and the C1 atom is in contrast to the O,Li isomer significantly pyramidalized. According to NBO analysis of the s‐trans and s‐cis isomers, the interaction of the lone pair at C1 with the π* orbital of the CC double bond is energetically much more favorable than that with the “empty” orbitals at the Li atom. The C1?S and C1?C2 conformations are determined by the stereoelectronic effects nC–σSR* interaction and allylic conjugation. 1H DNMR spectroscopy of racemic [EtCH?CHC(Et)SO2tBu]Li, [iPrCH?CHC(iPr)SO2tBu]Li and [EtCH?C(Me)C(Et)SO2tBu]Li in [D8]THF gave estimated barriers of enantiomerization of ΔG=13.2 kcal mol?1 (270 K), 14.2 kcal mol?1 (291 K) and 14.2 kcal mol?1 (295 K), respectively. Deprotonation of sulfone (R)‐EtCH?CHCH(Et)SO2tBu (94 % ee) with nBuLi in THF at ?105 °C occurred with a calculated enantioselectivity of 93 % ee and gave carbanion (M)‐[EtCH?CHC(Et)SO2tBu]Li, the deuteration and alkylation of which with CF3CO2D and MeOCH2I, respectively, proceeded with high enantioselectivities. Time‐dependent deuteration of the enantioenriched carbanion (M)‐[EtCH?CHC(Et)SO2tBu]Li in THF gave a racemization barrier of ΔG=12.5 kcal mol?1 (168 K), which translates to a calculated half‐time of racemization of t1/2=12 min at ?105 °C.  相似文献   

13.
汪敦佳  方正东  魏先红 《中国化学》2005,23(12):1600-1606
A new polyoxometalate (CPFX·HCl)3H4SiW12O40·10H2O was prepared from ciprofloxacin hydrochloride and H4SiW12O40·nH2O in aqueous solution, and characterized by elemental analysis, IR spectra and DTA-TG-DTG techniques. The IR spectrum confirmed the presence of Keggin structure and the characteristic functional group for ciprofloxacin in the compound. The TG-DTA-DTG curves showed that its thermal decomposition was a four-step process consisting of simultaneous collapse of Keggin type structure. The residue of decomposition was the mixture of WO3 and SiO2, confirmed by X-ray diffraction and IR spectroscopy. The decomposition mechanism and nonisothermal kinetic parameters of the polyoxometalate were obtained from an analysis to the TG-DTG curves by the single scanning methods (the Achar method and Coats-Redfern method) and the multiple scanning methods (the Kissinger method, Flynn-Wall-Ozawa method and Starink method). The results indicate that the kinetic equationswith parameters describing the thermal decomposition reaction are dα/dt=6.65×10^6[3(1-α)^2/3]e^-10495.5/T with E=87.26 kJ/mol and A=6.65×10^6 s^-1 for the second step,dα/dt=7.01×10^9(1-α)e^-18770.7/T with E=156.06 kJ/mol and A=7.01×10^9 s^-1 for the third step,dα/dt=9.77×10^43[(1-α)^2]e^-88980.0/T with E=739.78 kJ/mol and A=9.77×10^43 s^-1 for the fourth step.  相似文献   

14.
The first two persistent silenyl radicals (R2C=Si.?R), with a half‐life (t1/2) of about 30 min, were generated and characterized by electron paramagnetic resonance (EPR) spectroscopy. The large hyperfine coupling constants (hfccs) (a(29Siα)=137.5–148.0 G) indicate that the unpaired electron has substantial s character. DFT calculations, which are in good agreement with the experimentally observed hfccs, predict a strongly bent structure (?C=Si?R=134.7–140.7°). In contrast, the analogous vinyl radical, R2C=C.?R (t1/2≈3 h), exhibits a small hfcc (a(13Cα)=26.6 G) and has a nearly linear geometry (?C=C?R=168.7°).  相似文献   

15.
A new unsymmetrical, solid, Schiff base (H2LLi) was synthesized using L-lysine, o-vanillin and salicylaldehyde. An Er(III) complex of this ligand [Er(H2L)(NO3)](NO3)?·?2H2O was prepared and characterized by elemental analysis, IR, UV and molar conductance. The thermal decomposition kinetics of the complex for the second stage was studied under non-isothermal conditions by TG and DTG methods. The kinetic equation may be expressed as, dα/dt?=?A?·?e?E/RT ?·?1/2(1???α)[?ln(1???α)]?1. The kinetic parameters (E,?A), activation entropy S and activation free-energy G were also determined.  相似文献   

16.
To accelerate the convergence of the HH expansion, we modified the HH –GLF method, a new simple hyperspherical harmonic method proposed recently by us, into the CFHH –GLF method. Applications of the CFHH –GLF method to the three-body systems He and e? e? e+ exhibit very fast convergence with number of HH basis sets. With only 36 HH and five GLF , we obtain the ground-state energy of ?2.90371 au for He, compared with the exact value of ?2.90372 au, and with only 36 HH and 10 GLF , we obtained the ground-state energy of -0.26188 au for e? e? e+, compared with the exact value of ?0.26200 au. We formulate the CFHH –GLF method in this article. © 1994 John Wiley & Sons, Inc.  相似文献   

17.
Perylene (Py)‐containing polyacetylenes with different skeleton structures ? [HC?C(C6H4)CO2? Py]n? (P 1 ), ? [HC?C(CH2)8CO2? Py]n? (P 2 ), and ? {[(C6H5) C?C(CH2)9NH2]? co? [(C6H5)C?C(CH2)9? Py]}n? (P 3 ) are synthesized in satisfactory yields by Rh‐catalyzed polymerization (for P 1 and P 2 ) and polymer reaction (for P 3 ). All the polymers are soluble and possess high molecular weights (Mw up to 2.8 × 105). Their structures and properties are characterized and evaluated by IR, NMR, UV, TGA, PL, and photovoltaic (PV) analyses. The polymers are thermally stable, losing little of their weights when heated to 330 °C. When their solutions are irradiated, their perylene pendants emit intense red fluorescence at 610 nm. PV cells with a configuration of ITO/PEDOT:PSS/polymer/LiF/Al are fabricated, which show maximum current density of 10.3 μA/cm2. The external quantum efficiency is sensitive to the polymer structure, with P 3 exhibiting the highest value of 0.23%. © 2008 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 2025–2037, 2008  相似文献   

18.
A sterically encumbering multidentate β‐diketiminato ligand, tBuL2 (tBuL2=[ArNC(tBu)CHC(tBu)NCH2CH2N(Me)CH2CH2NMe2]?, Ar=2,6‐iPr2C6H3), is reported in this study along with its coordination chemistry to zirconium(IV). Using the lithio salt of this ligand, Li(tBuL2) ( 4 ), the zirconium(IV) precursor (tBuL2)ZrCl3 ( 6 ) could be readily prepared in 85 % yield and structurally characterized. Reduction of 6 with 2 equiv of KC8 resulted in formation of the terminal and mononuclear zirconium imide‐chloride [C(tBu)CHC(tBu)NCH2CH2N(Me)CH2CH2NMe2]Zr(=NAr)(Cl) ( 7 ) as the result of reductive C=N cleavage of the imino fragment in the multidentate ligand tBuL2 by an elusive ZrII species (tBuL2)ZrCl ( A ). The azabutadienyl ligand in 7 can be further reduced by 2 e? with KC8 to afford the anionic imide [K(THF)2]{[CH(tBu)CHC(tBu)NCH2CH2N(Me)CH2CH2N(Me)CH2]Zr=NAr} ( 8‐2THF ) in 42 % isolated yield. Complex 8‐2THF results from the oxidative addition of an amine C?H bond followed by migration to the vinylic group of the formal [C(tBu)CHC(tBu)NCH2CH2N(Me)CH2CH2NMe2]? ligand in 7 . All halides in 6 can be replaced with azides to afford (tBuL2)Zr(N3)3 ( 9 ) which was structurally characterized, and reduction with two equiv of KC8 also results in C=N bond cleavage of tBuL2 to form [C(tBu)CHC(tBu)NCH2CH2N(Me)CH2CH2NMe2]Zr(=NAr)(N3) ( 10 ), instead of the expected azide disproportionation to N3? and N2. Solid‐state single crystal structural studies confirm the formation of mononuclear and terminal zirconium imido groups in 7 , 8‐Et2O , and 10 with Zr=NAr distances being 1.8776(10), 1.9505(15), and 1.881(3) Å, respectively.  相似文献   

19.
Data on tensile strength and elongation at break for a series of Viton A-HV vulcanizates are discussed. The data were obtained at various extension rates at temperatures from ?5 to 230°C (25 ? TTg ? 260°C) on seven vulcanizates having crosslink densities ve (estimated from C1 in the Mooney-Rivlin equation) from 0.46 × 10?5 to 24.4 × 10?5 mole/cm3. At an extension rate of 1 min?1, an increase in ve affects the tensile strength σb (based on the undeformed cross-sectional area) and the true tensile strength σbσb (based on the cross-sectional area of a deformed specimen) as follows: σb is essentially constant at a low temperature; it passes through a decided maximum at intermediate temperatures; and it increases to a plateau at elevated temperatures. In contrast, λbσb decreases markedly at all temperatures, an exception being the most lightly crosslinked vulcanizate(s). Application of time—temperature superposition to the ultimate-property data gave log aT; its temperature dependence is that typical of nonpolar rubbery polymers. Data on the vulcanizates were compared in corresponding temperature states by plotting log 273σb/T, log 273λbσb/T, and (λb — 1)/(λb — 1)max against logtb/(tb)max, where tb is the temperature-reduced time to break and (tb)max is the value at which the ultimate extension ratio λb attains its maximum, (λb)max. Except for the most lightly crosslink vulcanizate, the comparison shows that 273λbσb/T and (λb — 1)/(λb — 1)max are substantially independent of (or only weakly dependent on) crosslink density, that 273λb/T increases with ve, and that 273λb/T ∝? ve0.6 and λb ∝? ve?0.4 at a large value of tb/(tb)max.  相似文献   

20.
A series of mono‐, bis‐, and tris(phenoxy)–titanium(IV) chlorides of the type [Cp*Ti(2‐R? PhO)nCl3?n] (n=1–3; Cp*=pentamethylcyclopentadienyl) was prepared, in which R=Me, iPr, tBu, and Ph. The formation of each mono‐, bis‐, and tris(2‐alkyl‐/arylphenoxy) series was authenticated by structural studies on representative examples of the phenyl series including [Cp*Ti(2‐Ph? PhO)Cl2] ( 1 PhCl2 ), [Cp*Ti(2‐Ph? PhO)2Cl] ( 2 PhCl ), and [Cp*Ti(2‐Ph? PhO)3] ( 3 Ph ). The metal‐coordination geometry of each compound is best described as pseudotetrahedral with the Cp* ring and the 2‐Ph? PhO and chloride ligands occupying three leg positions in a piano‐stool geometry. The mean Ti? O distances, observed with an increasing number of 2‐Ph? PhO groups, are 1.784(3), 1.802(4), and 1.799(3) Å for 1 PhCl2 , 2 PhCl , and 3 Ph , respectively. All four alkyl/aryl series with Me, iPr, tBu, and Ph substituents were tested for ethylene homopolymerization after activation with Ph3C+[B(C6F5)4]? and modified methyaluminoxane (7% aluminum in isopar E; mMAO‐7) at 140 °C. The phenyl series showed much higher catalytic activity, which ranged from 43.2 and 65.4 kg (mmol of Ti?h)?1, than the Me, iPr, and tBu series (19.2 and 36.6 kg (mmol of Ti?h)?1). Among the phenyl series, the bis(phenoxide) complex of 2 PhCl showed the highest activity of 65.4 kg (mmol of Ti?h)?1. Therefore, the catalyst precursors of the phenyl series were examined by treating them with a variety of alkylating reagents, such as trimethylaluminum (TMA), triisobutylaluminum (TIBA), and methylaluminoxane (MAO). In all cases, 2 PhCl produced the most catalytically active alkylated species, [Cp*Ti(2‐Ph? PhO)MeCl]. This enhancement was further supported by DFT calculations based on the simplified model with TMA.  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