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 609 毫秒
1.
Electrospray Ionization Mass Spectrometry (ESI/MS) has been used to determine the association constants (KAs) and binding stoichiometries for parent para-Sulphonato-calix[n]arenes and their derivatives with bovine serum albumin (BSA). KA values were determined by titration experiments using a constant concentration of protein. KA measurements were carried out in a methanol–formic acid solution. 5,11,17,23–tetra-Sulphonato-calix[4]arene (1a) and 25-mono-(2-aminoethoxy)-5,11,17,23-tetra-Sulphonato-calix[4]arene (1d) interact strongly with BSA showing 3 non-equivalent binding sites with KA1 = 7.69 × 105 M−1, KA2 = 3.85 × 105 M−1, KA3 = 0.33 × 105 M−1 and KA1 = 1.69 × 105 M−1, KA2 = 2.94 × 105 M−1, KA3 = 0.60 × 105 M−1, respectively. The strength of the interactions between the calixarene and BSA is inversely proportional to the size of macrocyclic ring: n = 4 > n=6>>n=8.  相似文献   

2.
The kinetics of the aquation of (H2O)5Cr(O2CCCl3)2+ have been examined at 35–55°C and 1.00M ionic strength with [H+] = 0.01?1.00M. The reaction follows the rate equation -d ln [Crtotal]/dt = (a[H+]?1 + b + c[H+])/(1 + d[H+]), where [Crtotal] is the stoichiometric concentration of the complex. At 45°C a = (1.41 ± 0.03) × 10?7M/s, b = (1.66 ± 0.02) × 10?5 s?1, c = (7.0 ± 0.8) × 10?5M?1·S?1 and d = 2.3 ± 0.3M?1. Two mechanisms consistent with this rate law are discussed, with evidence being presented in favor of an ester hydrolysis mechanism involving steady-state intermediates. Equilibrium and activation parameters were determined.  相似文献   

3.
The kinetics of oxidation of tartaric acid (TAR) by peroxomonosulfate (PMS) in the presence of Cu(II) and Ni(II) ions was studied in the pH range 4.05–5.20 and also in alkaline medium (pH ~12.7). The rate was calculated by measuring the [PMS] at various time intervals. The metal ions concentration range used in the kinetic studies was 2.50 × 10?5 to 1.00 × 10?4 M [Cu(II)], 2.50 × 10?4 to 2.00 × 10?3M [Ni(II)], 0.05 to 0.10 M [TAR], and µ = 0.15 M. The metal(II) tartarates, not TAR/tartarate, are oxidized by PMS. The oxidation of copper(II) tartarate at the acidic pH shows an appreciable induction period, usually 30–60 min, as in classical autocatalysis reaction. The induction period in nickel(II) tartarate is small. Analysis of the [PMS]–time profile shows that the reactions proceed through autocatalysis. In alkaline medium, the Cu(II) tartarate–PMS reaction involves autocatalysis whereas Ni(II) tartarate obeys simple first‐order kinetics with respect to [PMS]. The calculated rate constants for the initial oxidation (k1) and catalyzed oxidation (k2) at [TAR] = 0.05 M, pH 4.05, and 31°C are Cu(II) (1.00 × 10?4 M): k1 = 4.12 × 10?6 s?1, k2 = 7.76 × 10?1 M?1s?1 and Ni(II) (1.00 × 10?3 M): k1 = 5.80 × 10?5 s?1, k2 = 8.11 × 10?2 M?1 s?1. The results suggest that the initial reaction is the oxidative decarboxylation of the tartarate to an aldehyde. The aldehyde intermediate may react with the alpha hydroxyl group of the tartarate to give a hemi acetal, which may be responsible for the autocatalysis. © 2011 Wiley Periodicals, Inc. Int J Chem Kinet 43: 620–630, 2011  相似文献   

4.
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   

5.
The syntheses of a series of l‐methyl‐3‐aryl‐substituted titanocene and zirconocene dichlorides are reported. These complexes are synthesized by the reaction of 2‐ and 3‐methyl‐6, 6‐dimethylfulvenes (1:4) with aryllithium, followed by the reaction with TiCl4·2THF, ZrCl4 and (CpTiCl2)2O respectively, to give complexes 1–5. The complex [η5‐1‐methyl‐3‐(α, α‐dimethylbenzyl) cyclopentadienyl] titanium dichloride has been studied by X‐ray diffraction. The red crystal of this complex is monoclinic, space group P2t/C with unit cell parameters: a =6.973(6) × 10?1 nm, b =36.91(2) × 10?1 nm, c = 10.063(4) × 10?1 nm, α=β= γ = 93.35(5)°, V = 2584(5) × 10?3 nm3 and Z = 4. Refinement for 1004 observed reflections gives the final R of 0.088. There are four independent molecules per unit cell.  相似文献   

6.
The reaction between C2H5 and O2 at 295 K has been studied with a flow reactor sampled by a mass spectrometer. With helium as the carrier gas the rate coefficient was found to increase from (1.2 ± 0.3) × 10?12 to (3.6 ± 0.9) × 10?12 cm3/s as [He] was increased from 2 × 1016 to 3.4 × 1017 cm?3. The importance of has been determined from a knowledge of the initial C2H5 concentration together with a measurement of the C2H4 produced in reaction (5). F, the fraction of the C2H5 radicals removed by path (5), was found to decrease from 0.15 to 0.06 as [He] increased from 2 × 1016 to 3.4 × 1017 cm?3. The rate coefficient for reaction (5) was found to be independent of [He] and to have a value of (2.1 ± 0.5) × 10?13 cm3/s. The variation in F reflects the fact that k1b increases as [He] increases. These observations are taken as evidence for a direct mechanism for C2H4 production and a collision-stabilized route for C2H5O2 formation. Calculations indicate that the high-pressure limit for reaction (1b) is ~4.4 × 10?12 cm3/s and that in the polluted troposphere the branching ratio for reactions (1b) and (5) will be ~l20.  相似文献   

7.
Two samples of cellulose (molecular weight 2.97 × 105 and 1.25 × 105) were transformed into carbanilates (CTC) which were then fractionated by the elution method at a constant composition of the acetone-water elution mixture with the column temperature gradually increasing from ?30°C to 30°C, and by the GPC method in acetone and tetrahydrofuran. Tetrahydrofuran appeared to be a more suitable solvent. The molecular weights of fractions obtained by the elution fractionation were determined by the light-scattering method in tetrahydrofuran. The width of fractions was determined by the GPC method (average M w/M n = 1.37); the [η] values and the Mark-Houwink constants (K = 5.3 × 10-3, a = 0.84) for tetrahydrofuran at 25°C were determined. The calibration curve for the GP method was constructed by means of the fractions thus obtained; it was demonstrated that the universal calibration curve according to Benoit can also be used. It was demonstrated that the molecular weight distribution of cellulose can be conveniently determined by conversion into CTC followed either by the elution fractionation (for preparative purposes) or by fractionation by the GPC method (for analytical purposes).  相似文献   

8.
Various types of soluble crosslinked polymers obtained from the copolymerization of methylmethacrylate (MMA) and p-divinylbenzene (p-DVB) in the presence of a transfer agent (CBr4) have been discussed in relation to the variation of the structure during the reaction time. When [p-DVB]/[MMA] = 1.49 × 10?3 and [CBr4]/[MMA] = 1.28 × 10?4, only linear polymers (primary polymer; M n = 1.0 × 105) with pendant vinyl groups are formed intially. Considerable branched structure is attained in rather large polymers (M n = 2.5 × 105), but the number of pendant double bonds is not enough to reach the gelation. As the concentration of the transfer agent becomes high, the intermolecular crosslinking is depressed, and the formed polymers contain loops and short chains. At [p – DVB]/[MMA] = 7.43 × 10?3 and [CBr4]/[MMA] = 1.28 × 10?3, the shape of polymer with the same M n became compact gradually with increasing reaction time. These results are considered to be useful for the preparation of soluble crosslinked polymer with controlled structure.  相似文献   

9.
Multiarm star‐branched polymers based on poly(styrene‐b‐isobutylene) (PS‐PIB) block copolymer arms were synthesized under controlled/living cationic polymerization conditions using the 2‐chloro‐2‐propylbenzene (CCl)/TiCl4/pyridine (Py) initiating system and divinylbenzene (DVB) as gel‐core‐forming comonomer. To optimize the timing of isobutylene (IB) addition to living PS⊕, the kinetics of styrene (St) polymerization at −80°C were measured in both 60 : 40 (v : v) methyl cyclohexane (MCHx) : MeCl and 60 : 40 hexane : MeCl cosolvents. For either cosolvent system, it was found that the polymerizations followed first‐order kinetics with respect to the monomer and the number of actively growing chains remained invariant. The rate of polymerization was slower in MCHx : MeCl (kapp = 2.5 × 10−3 s−1) compared with hexane : MeCl (kapp = 5.6 × 10−3 s−1) ([CCl]o = [TiCl4]/15 = 3.64 × 10−3M; [Py] = 4 × 10−3M; [St]o = 0.35M). Intermolecular alkylation reactions were observed at [St]o = 0.93M but could be suppressed by avoiding very high St conversion and by setting [St]o ≤ 0.35M. For St polymerization, kapp = 1.1 × 10−3 s−1 ([CCl]o = [TiCl4]/15 = 1.82 × 10−3M; [Py] = 4 × 10−3M; [St]o = 0.35M); this was significantly higher than that observed for IB polymerization (kapp = 3.0 × 10−4 s−1; [CCl]o = [Py] = [TiCl4]/15 = 1.86 × 10−3M; [IB]o = 1.0M). Blocking efficiencies were higher in hexane : MeCl compared with MCHx : MeCl cosolvent system. Star formation was faster with PS‐PIB arms compared with PIB homopolymer arms under similar conditions. Using [DVB] = 5.6 × 10−2M = 10 times chain end concentration, 92% of PS‐PIB arms (Mn,PS = 2600 and Mn,PIB = 13,400 g/mol) were linked within 1 h at −80°C with negligible star–star coupling. It was difficult to achieve complete linking of all the arms prior to the onset of star–star coupling. Apparently, the presence of the St block allows the PS‐PIB block copolymer arms to be incorporated into growing star polymers by an additional mechanism, namely, electrophilic aromatic substitution (EAS), which leads to increased rates of star formation and greater tendency toward star–star coupling. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 1629–1641, 1999  相似文献   

10.
The redox system of potassium persulfate–thiomalic acid (I1–I2) was used to initiate the polymerization of acrylamide (M) in aqueous medium. For 20–30% conversion the rate equation is where Rp is the rate of polymerization. Activation energy is 8.34 kcal deg?1 mole?1 in the investigated range of temperature 25–45°C. Mn is directly proportional to [M] and inversely to [I1]. The range of concentrations for which these observations hold at 35°C and pH 4.2 are [I1] = (1.0–3.0) × 10?3, [I2] = (3.0–7.5) × 10?3, and [M] = 5.0 × 10?2–3.0 × 10?1 mole/liter.  相似文献   

11.
Absolute rate constants for H-atom abstraction by OH radicals from cyclopropane, cyclopentane, and cycloheptane have been determined in the gas phase at 298 K. Hydroxyl radicals were generated by flash photolysis of H2O vapor in the vacuum UV, and monitored by time-resolved resonance absorption at 308.2 nm [OH(A2Σ+X2Π)]. The rate constants in units of cm3 mol−1 s−1 at the 95% confidence limits were as follows: k(c C3H6) = (3.74 ± 0.83) × 1010, k(c C5H10) = (3.12 ± 0.23) × 1012, k(c C7H14) = (7.88 ± 1.38) × 1012. A linear correlation was found to exist between the logarithm of the rate constant per C H bond and the corresponding bond dissociation energy for several classes of organic compounds with equivalent C H bonds. The correlation favors a value of D(c C3H5–H) = (101 ± 2) kcal mol−1.  相似文献   

12.
张洪林  于秀芳  聂毅  刘晓静  张刚 《中国化学》2003,21(11):1466-1469
IntroductionMostcomplicatedreactionshappenedinlivingcrea tures ,amongthemenzymecatalyzedreactionisanimpor tantclass .Itissignificantinboththeoryandpracticetoinvestigateenzymecatalyzedreaction .Therearemanyex perimentalmethodssuchasspectrophotometry ,titrimetry ,isotopemethod ,microcalorimetryandsoon ,inwhichmi crocalorimetryisanewoneduetoitshighsensitivityandaccuracy .Wecanstudythewholeprocessoftheheatef fectusingamicrocalorimeter .Sincetheabsorptionorpro ductionofheatisanintrinsicpropertyofe…  相似文献   

13.
The complexes [Pt(3′′‐clpbpy)Cl2] ( 1 ) [3′′‐clpbpy = 4‐(3′′‐chlorophenyl)‐6‐phenyl‐2, 2′‐bipyridine], [Pt(4′′‐clpbpy)Cl2] ( 2 ) [4′′‐clpbpy = 4‐(4′′‐chlorophenyl)‐6‐phenyl‐2, 2′‐bipyridine], [Pt(3′′‐brpbpy)Cl2] ( 3 ) [3′′‐brpbpy = 4‐(3′′‐bromophenyl)‐6‐phenyl‐2, 2′‐bipyridine], and [Pt(4′′‐brpbpy)Cl2] ( 4 ) [4′′‐brpbpy = 4‐(4′′‐bromophenyl)‐6‐phenyl‐2, 2′‐bipyridine] were synthesized and characterized. The binding of the complexes with herring sperm DNA (HS DNA) was investigated by absorption titration and viscosity measurements. It was found that the complexes have ability of interaction with DNA by covalent mode. The intrinsic binding constant Kb of the complexes with HS DNA is 8.76 × 104 ( 1 ), 9.89 ×104 ( 2 ), 1.52 × 105 ( 3 ), and 2.31 × 105 ( 4 ) M–1. The slight depression in relative specific viscosity was observed, which also attributes to covalent binding of complexes with DNA bases. Gel electrophoresis assay demonstrated the ability of the complexes to unwind negatively supercoiled pUC19 plasmid by 14° ( 1 ), 13° ( 2 ), 13° ( 3 ), and 11° ( 4 ). The in vitro cytotoxic property of the synthesized metal complexes was also carried out against brine shrimp bioassay.  相似文献   

14.
Organic field‐effect transistors incorporating planar π‐conjugated metal‐free macrocycles and their metal derivatives are fabricated by vacuum deposition. The crystal structures of [H2(OX)] (H2OX=etioporphyrin‐I), [Cu(OX)], [Pt(OX)], and [Pt(TBP)] (H2TBP=tetra‐(n‐butyl)porphyrin) as determined by single crystal X‐ray diffraction (XRD), reveal the absence of occluded solvent molecules. The field‐effect transistors (FETs) made from thin films of all these metal‐free macrocycles and their metal derivatives show a p‐type semiconductor behavior with a charge mobility (μ) ranging from 10?6 to 10?1 cm2 V?1 s?1. Annealing the as‐deposited Pt(OX) film leads to the formation of a polycrystalline film that exhibits excellent overall charge transport properties with a charge mobility of up to 3.2×10?1 cm2 V?1 s?1, which is the best value reported for a metalloporphyrin. Compared with their metal derivatives, the field‐effect transistors made from thin films of metal‐free macrocycles (except tetra‐(n‐propyl)porphycene) have significantly lower μ values (3.0×10?6–3.7×10?5 cm2 V?1 s?1).  相似文献   

15.
The kinetics of the atmospherically important gas-phase reactions of acenaphthene and acenaphthylene with OH and NO3 radicals, O3 and N2O5 have been investigated at 296 ± 2 K. In addition, rate constants have been determined for the reactions of OH and NO3 radicals with tetralin and styrene, and for the reactions of NO3 radicals and/or N2O5 with naphthalene, 1- and 2-methylnaphthalene, 2,3-dimethylnaphthalene, toluene, toluene-α,α,α-d3 and toluene-d8. The rate constants obtained (in cm3 molecule?1 s?1 units) at 296 ± 2 K were: for the reactions of O3; acenaphthene, <5 × 10?19 and acenaphthylene, ca. 5.5 × 10?16; for the OH radical reactions (determined using a relative rate method); acenaphthene, (1.03 ± 0.13) × 10?10; acenaphthylene, (1.10 ± 0.11) × 10?10; tetralin, (3.43 ± 0.06) × 10?11 and styrene, (5.87 ± 0.15) × 10?11; for the reactions of NO3 (also determined using a relative rate method); acenaphthene, (4.6 ± 2.6) × 10?13; acenaphthylene, (5.4 ± 0.8) × 10?12; tetralin, (8.6 ± 1.3) × 10?15; styrene, (1.51 ± 0.20) × 10?13; toluene, (7.8 ± 1.5) × 10?17; toluene-α,α,α-d3, (3.8 ± 0.9) × 10?17 and toluene-d8, (3.4 ± 1.9) × 10?17. The aromatic compounds which were observed to react with N2O5 and the rate constants derived were (in cm3 molecule?1 s?1 units): acenaphthene, 5.5 × 10?17; naphthalene, 1.1 × 10?17; 1-methylnaphthalene, 2.3 × 10?17; 2-methylnaphthalene, 3.6 × 10?17 and 2,3-dimethylnaphthalene, 5.3 × 10?17. These data for naphthylene and the alkylnaphthalenes are in good agreement with our previous absolute and relative N2O5 reaction rate constants, and show that the NO3 radical reactions with aromatic compounds proceed by overall H-atom abstraction from substituent-XH bonds (where X = C or O), or by NO3 radical addition to unsaturated substituent groups while the N2O5 reactions only occur for aromatic compounds containing two or more fused six-membered aromatic rings.  相似文献   

16.
The quantum yields of phosphorescence (Φp) of biacetyl have been determined in pure biacetyl, biacetyl-SO2, and biacetyl-c-C6H12 mixtures in experiments using bands of radiation centered at 3450, 3650, 3880, and 4348 Å. It has been shown that the unexpected effect of gas concentration on the quantum yields of the sulfur dioxide triplet-sensitized phosphorescence of biacetyl resulted largely from the significant destruction of biacetyl triplets at the wall of the cell. The kinetics of the variation of Φp with [Ac2], wavelength of the absorbed light, and added gases provide new estimates of the energy relations and the rate constants for the decomposition reaction of vibrationally rich biacetyl molecules in the first excited singlet state (1Ac2?): 1Ac2? → products (1), 1Ac2? + Ac21Ac2 + Ac2 (2); the minimum energy necessary in 1Ac2? for reaction (1) to occur is estimated to be about 72.8 kcal/mole above the ground state of biacetyl: k1/k2 = (4.3 ± 0.1) × 10?3M at 3450 Å, (4.07 ± 0.04) × 10?4M at 3650 Å, and (5.6 ± 0.4) × 10?5M at about 3800 Å. The variation of the rate constant ratio is shown to be consistent with the expectations of the simple theory of excited molecule decomposition. Biacetyl triplet (3Ac2) rate constants were determined by measurements of Φp in O2 and NO-containing mixtures: 3Ac2 + S → (Ac2–S, products) (8); for O2 = S, k8 = (5.76 ± 0.40) × 108 (3650 Å experiments), (5.76 ± 0.27) × 108 (4358 Å); for NO = S, k8 = (3.34 ± 0.20) × 109 (3650 Å), (3.33 ± 0.18) × 109 1./mole-sec (4358 Å). A comparison between these and previous findings of the SO2 triplet (3SO2)-sensitized excitation of biacetyl [5,6] show that the decomposition of the initial 3Ac2 product of the exothermic energy transfer reaction 3SO2 + Ac2 → SO2 + 3Ac2 is unimportant.  相似文献   

17.
Short-range interactions between chain units of random copolymers in solution may be influenced by the composition or precisely by the distribution of sequence lengths of the same monomer units. Steric factors were derived for random copolymers of styrene and acrylonitrile with different compositions from the relation between the limiting viscosity number and the molecular weight. Mark-Houwink relations were obtained in methyl ethyl ketone (MEK) or in N,N′-dimethylformamide (DMF) at 30°C. for random copolymers containing 0.383 (Co-1) and 0.626 (Co-2) mole fraction of acrylonitrile, the expressions are: [η] = 3.6 X 10?4 M w0.62, for Co-1 in MEK; [η] = 5.3 X 10?4 M w0.61, for Co-2 in MEK; [η] = 1.2 × 10?4M w0.77 for Co-2 in DMF. With the Stockmayer-Fixman expression, these correlations become, respectively: [η]/M1/2 = 1.24 × 10?3 + 8.0 × 10?7 M1/2; and [η]/M1/2 = 1.70 × 10?3 + 6.3 × 10?7 M1/2; and [η]/M1/2 = 1.68 × 10?3 + 31.3 × 10?7 M1/2. From the unperturbed mean-square end-to-end distances, 〈L20, determined from the first terms of the latter expressions, together with 〈L20f calculated by assuming the completely free rotation, gives the steric factor σ = (〈L20/〈L20f)1/2 as 2.25 ± 0.05 for Co-1, and 2.31 ± 0.10 for Co-2. These values of σ are close to those for polystyrene (σ = 2.22 ± 0.05) and for polyacrylonitrile (σ = 2.20 ± 0.05). Therefore, it is concluded that the dimensions of random copolymers of styrene and acrylonitrile in solution are not significantly influenced by the composition. In other words, the unperturbed dimensions are not affected by a change in the alternation tendency between styrene units with phenyl side groups having a large molar volume and acrylonitrile units with nitrile groups responsible for the electrostatic interactions. On the other hand, the long-range interactions reflect the effect of sequence length. The Huggins constant and the second virial coefficient obtained from the light-scattering measurements have optimum values at about 0.5 mole fraction of acrylonitrile, where the greatest tendency for alternation seems to exist.  相似文献   

18.
Heterojunctions between polyaniline (PANI) and n-type porous silicon (PS), Al/PS-PANI/Au cell,were fabricated, and the rectifying parameters of this heterojunction diode were measured as a function of thepreparation conditions of PANI and PS, the electronic structure of PANI as well as cell structure. Therectifying parameters of Al/PS-PANI/Au cell were determined to be γ= 1 .8×10~1~ 1 .0×10~5 for the rectifyingratio at 3V, n = 3 ~12 for the ideal factor,j_0 = 8.0×10~(-5)~5.6×10~(-2) mA/cm~2 for the reversed saturated currentdensity, and φ_b = 0.67~ 0.83 V for the barrier height, respectively. The best rectifying heterojunction diodemade between PANI and n-type PS with higher rectifying factor (γ= 1 .0×10~5 at 3V ), output current (>1500mA/cm~2 at 3V) and lower ideal factor (n = 3.3) was obtained by preventing the oxidation of PS beforeevaporating Al electrode.  相似文献   

19.
At pH 4.5 (citrate buffer), D -gluconhydroximo-lactone ( 2 ), the N-methylurethane 3 and the N-phenylurethane 4 inhibit competitively the hydrolysis of p-nitrophenyl β-D -glucopyranoside by emulsin. The IC50 values of 2, 3 , and 4 were 1.6 × 10?4, 1.0 × 10?4, and 5.8 × 10?6 M , respectively. The Ki values of 2 and 4 were 9.8 × 10?5 and 2.3 × 10?6 M , respectively, while D-glucono-1,5-lactone ( 1 ) showed IC50 = 1.1 × 10?4 M and Ki = 3.7 × 10?5 M .  相似文献   

20.
Phenylacetylene was polymerized by WCl6·Ph4Sn (1:1) in 1,4-dioxane to provide in high yield a polymer whose molecular weight reached 1 × 105. The polymerization also proceeded in other oxygen-containing solvents (ethers, esters, and ketones) but the polymer molecular weights were lower than 1 × 104. Certain hydrocarbon solvents such as cyclohexene, tetralin, and indan also afforded high-molecular-weight polyphenylacetylene [M n = (5–8) × 104], as compared with those (M n ≤ 1.5 × 104) obtained in conventional aromatic hydrocarbons like benzene. A high polymer (M n = 1.6 × 105) was also formed from β-naphthylacetylene in 1,4-dioxane. It was inferred that the active hydrogens of these solvents prevent the formed polymer from being decomposed by a radical mechanism and/or modify the nature of active species.  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