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.
Phenyl‐modified natural rubber was prepared in latex stage by bromination of deproteinized natural rubber followed by Suzuki‐Miyaura cross‐coupling reaction. First, the bromination of natural rubber was carried out using N‐bromosuccinimide in latex stage. The bromine atom content increased as amount of N‐bromosuccinimide increased. Second, the allylic bromine atom was replaced with a phenyl group using phenyl boronic acid in the presence of a palladium catalyst, according to the Suzuki‐Miyaura cross‐coupling reaction in latex stage. The resulting products were characterized by nuclear magnetic resonance (NMR) spectroscopy. Signal at 7.13 ppm was assigned to the phenyl group of the product, while signals at 3.98, 4.14, and 4.44 ppm were assigned to the remaining allylic brominated cis‐1,4‐isoprene units. The estimated phenyl group content and the conversion of the Suzuki‐Miyaura cross‐coupling reaction were 1.32 and 23.7 mol%, respectively. Glass transition temperature (Tg) of deproteinized natural rubber increased from ?62°C to ?46.7°C, when the phenyl group was introduced into the rubber.  相似文献   

2.
Ultra‐performance convergence chromatography, which integrates the advantages of supercritical fluid chromatography and ultra high performance liquid chromatography technologies, is an environmentally friendly analytical method that uses dramatically reduced amounts of organic solvents. An ultra‐performance convergence chromatography method was developed and validated for the quantification of decursinol angelate and decursin in Angelica gigas using a CSH Fluoro‐Phenyl column (2.1 mm × 150 mm, 1.7 μm) with a run time of 4 min. The method had an improved resolution and a shorter analysis time in comparison to the conventional high‐performance liquid chromatography method. This method was validated in terms of linearity, precision, and accuracy. The limits of detection were 0.005 and 0.004 μg/mL for decursinol angelate and decursin, respectively, while the limits of quantitation were 0.014 and 0.012 μg/mL, respectively. The two components showed good regression (correlation coefficient (r2) > 0.999), excellent precision (RSD < 2.28%), and acceptable recoveries (99.75–102.62%). The proposed method can be used to efficiently separate, characterize, and quantify decursinol angelate and decursin in Angelica gigas and its related medicinal materials or preparations, with the advantages of a shorter analysis time, greater sensitivity, and better environmental compatibility.  相似文献   

3.
Soyalkaloid A was isolated from Portulaca oleracea L. for the first time in our laboratory and then a rapid and sensitive ultra‐high‐performance liquid chromatography electrospray ionization quadrupole–time of flight mass spectrometry (UHPLC–ESI–Q–TOF/MS) method with hesperidin as internal standard (IS) was developed and validated to investigate the pharmacokinetics of soyalkaloid A in rats after oral and intravenous administrations. The analysis was achieved on an Agilent Zorbax Eclipse Plus C18 Column (2.1 × 50 mm, 1.8 μm) by elution with acetonitrile and water (containing 0.1% formic acid), at a flow rate of 0.3 mL/min. The MS analysis was performed in the positive ion mode with monitored ion m/z 227.0814 [M + H]+ and 611.1971 [M + H]+ for soyalkaloid A and IS, respectively. The linear range was established over the concentration range 7.5–6000 ng/mL (r = 0.9951). The intra‐ and inter‐assay accuracy and precision were between ?4.86‐4.49 and 1.93–9.66, respectively. The lower limits of detection and quantitation observed were 2.1 and 7.4 ng/mL, respectively. The rapid, sensitive and specific UHPLC–ESI–Q–TOF/MS method was successfully applied to a pharmacokinetic study of soyalkaloid A. Moreover, its antioxidant was studied via a 1,1‐diphenyl‐2‐picryl‐hydrazyl radical scavenging assay, the IC50 value being 20.73 ± 0.51 μM.  相似文献   

4.
A multi‐analyte screening method for the quantification of 50 acidic/neutral drugs in human plasma based on on‐line solid‐phase extraction (SPE)–HPLC with photodiode array detection (DAD) was developed, validated and applied for clinical investigation. Acetone and methanol for protein precipitation, three different SPE materials (two electro‐neutral, one strong anion‐exchange, one weak cation‐exchange) for on‐line extraction, five HPLC‐columns [one C18 (GeminiNX), two phenyl‐hexyl (Gemini C6‐Phenyl, Kinetex Phenyl‐Hexyl) and two pentafluorophenyl (LunaPFP(2), KinetexPFP)] for analytical separation were tested. For sample pre‐treatment, acetone in the ratio 1:2 (plasma:acetone) showed a better baseline and fewer matrix peaks in the chromatogram than methanol. Only the strong anion‐exchanger SPE cartridge (StrataX‐A, pH 6) allowed the extraction of salicylic acid. Analytical separation was carried out on a Gemini C6‐Phenyl column (150 × 4.6 mm, 3 µm) using gradient elution with acetonitrile–water 90:10 (v/v) and phosphate buffer (pH 2.3). Linear calibration curves with correlation coefficients r ≥ 0.9950/0.9910 were obtained for 46/four analytes. Additionally, this method allows the quantification of 23 analytes for therapeutic drug monitoring. Limits of quantitation ranged from 0.1 (amobarbital) to 23 mg/L (salicylic acid). Inter‐/intra‐day precisions of quality control samples (low/high) were better than 13% and accuracy (bias) ranged from ?14 to 10%. A computer‐assisted database was created for automated detection of 223 analytes of toxicological interests. Four cases of multi‐drug intoxications are presented. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

5.
This study aimed to investigate the tolerance of a melanized wild‐type strain of Aspergillus niger (CON1) and its light‐colored mutant (MUT1) to UV–C light and the concomitantly generated ozone. Treatments were segregated into four groups based on whether UV irradiation was used and the presence or absence of ozone: (?UV, ?O3), (?UV, +O3), (+UV, ?O3) and (+UV, +O3). The survival of CON1 and MUT1 conidia under +UV decreased as the exposure time increased, with CON1 showing greater resistance to UV irradiation than MUT1. Ozone induced CON1 conidium inactivation only under conditions of UV radiation exposure. While, the inactivation effect of ozone on MUT1 was always detectable regardless of the presence of UV irradiation. Furthermore, the CON1 conidial suspension showed lower UV light transmission than MUT1 when examined at the same concentration. Compared with the pigment in MUT1, the melanin in CON1 exhibited more potent radical‐scavenging activity and stronger UV absorbance. These results suggested that melanin protected A. niger against UV disinfection via UV screening and free radical scavenging. The process by which UV–C disinfection induces a continual decrease in conidial survival suggests that UV irradiation and ozone exert a synergistic fungicidal effect on A. niger prior to reaching a plateau.  相似文献   

6.
Novel, biodegradable poly(?‐caprolactone)‐block‐poly(trans‐4‐hydroxy‐N‐benzyloxycarbonyl‐L ‐proline)‐block‐poly(?‐caprolactone) triblock copolymers were synthesized by ring‐opening polymerization from dihydroxyl‐terminated macroinitiator poly(trans‐4‐hydroxy‐N‐benzyloxycarbonyl‐L ‐proline) (PHpr) and ?‐caprolactone (?‐CL) with stannous octoate as the catalyst. The molecular weights were characterized with gel permeation chromatography and matrix‐assisted laser desorption/ionization time‐of‐flight mass spectrometry. With an increase in the contents of ?‐CL incorporated into the copolymers, a decrease in the glass‐transition temperature (Tg) was observed. The Tg values of copoly(4‐phenyl‐?‐caprolactone) and copoly(4‐methyl‐?‐caprolactone) were higher than Tg of copoly(?‐caprolactone). Their micellar characteristics in an aqueous phase were investigated with fluorescence spectroscopy, dynamic light scattering, and transmission electron microscopy. The block copolymers formed micelles in the aqueous phase with critical micelle concentrations in the range of 1.00–1.36 mg L?1. With higher molecular weights and hydrophobic components in the copolymers, a higher critical micelle concentration was observed. As the feed weight ratio of antitriptyline hydrochloride (AM) to the polymer increased, the drug loading increased. The micelles exhibited a spherical shape, and the average size was less than 250 nm. The in vitro hydrolytic degradation and controlled drug release properties of the triblock copolymers were also investigated. © 2006 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 44: 4268–4280, 2006  相似文献   

7.
A reversed‐phase high performance liquid chromatography method has been developed and validated for determination and quantitation of the natural sesquiterpene (−)‐α‐bisabolol. Furthermore the application of the method was done by characterization of chitosan milispheres and liposomes entrapping Zanthoxylum tingoassuiba essential oil, which contains appreciable amount of (−)‐α‐bisabolol. A reversed‐phase C18 column and gradient elution was used with the mobile phase composed of (A) acetonitrile–water–phosphoric acid (19:80:1) and (B) acetonitrile. The eluent was pumped at a flow rate of 0.8 mL/min with UV detection at 200 nm. In the range 0.02–0.64 mg/mL the assay showed good linearity (R2 = 0.9999) and specificity for successful identification and quantitation of (−)‐α‐bisabolol in the essential oil without interfering peaks. The method also showed good reproducibility, demonstrating inter‐day and intra‐day precision based on relative standard deviation values (up to 3.03%), accuracy (mean recovery of 100.69% ± 1.05%) and low values of detection and quantitation limits (0.0005 and 0.0016 mg/mL, respectively). The method was also robust for showing a recovery of 98.81% under a change of solvent in standard solutions. The suitability of the method was demonstrated by the successful determination of association efficiency of the (−)‐α‐bisabolol in chitosan milispheres and liposomes. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

8.
Radical‐induced oxidation of reversible addition‐fragmentation chain transfer (RAFT) agents is investigated with respect to the effect of molecular structure on oxidation rate. The radicals are generated by homolysis of either azobisisobutyronitrile or alkoxyamine and transformed in situ immediately into peroxy radicals through transfer to molecular oxygen. It is found that the oxidation rate depends on the structure of Z‐ and R‐group of thiocarbonylthio compounds. For dithioesters with identical Z‐phenyl substituent, the oxidation rate decreases in the order of cyanoisopropyl (? C(Me)2CN) > cumyl (? C(Me)2Ph) > phenylethyl (? CH(Me)Ph) > 2‐methoxy‐1‐methyl‐2‐oxoethyl (? CH(Me) ? C(?O)OCH3) > benzyl (? CH2Ph). For dithioesters with identical R‐group, the oxidation rate decreases in the order of Z = phenyl? ~ benzyl? > RS? (trithiocarbonates) > RO? (xanthates). The stability of the RAFT agents toward oxidation correlates well with the chain transfer abilities as those previously reported by Rizzardo and coworkers. The priority of the oxidation reaction over the RAFT process, and the subsequent influence on RAFT polymerization are also investigated. © 2011 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2011  相似文献   

9.
Extraction and pre‐concentration of a bioactive marker compound, phenyl‐1,3,5‐heptatriyne from Bidens pilosa, prior to HPLC has been demonstrated using both organic and ecofriendly solvents. Non‐ionic surfactants, viz. Triton X‐100, Triton X‐114 and Genapol X‐80, were used for extraction. No back‐extraction or liquid chromatographic steps were required to remove the target phytochemical from the surfactant‐rich extractant phase. The optimized cloud point extraction procedure has been shown to be a potentially useful methodology for the preconcentration of the target analyte, with a preconcentration factor of 4–99. Moreover, the method is simple, sensitive, rapid and consumes lesser solvent than traditional methods. An isocratic chromatographic separation and quantitation was accomplished on a C18 column with acetonitrile–acidified aqueous as mobile phase at a flow rate of 1.0 mL/min, UV detection at 254 nm and specificity with photo diode‐array detector (PDA) and MS. Under the optimum experimental conditions recovery was satisfactory (99.18–100.33%) without interference from the surfactant. The method seems to be reliable with intraday precision and interday precision below 2.0%. Good linearity was obtained in the working range from 7.5 to 30 µg/mL with correlation coefficient >0.99. The limits of detection and quantitation were 1.84 and 6.13 µg/mL, respectively. The method was validated following international guidelines and successfully applied for quantitative assays of cytotoxic compound phenyl‐1,3,5‐heptatriyne in Bidens pilosa. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

10.
A new pre‐column derivative high‐performance liquid chromatography (HPLC) method for determination of d ‐glucose with 3‐O‐methyl‐d ‐glucose (3‐OMG) as the internal standard was developed and validated in order to study the gluconeogenesis in HepG2 cells. Samples were derivatized with 1‐phenyl‐3‐methy‐5‐pyrazolone at 70°C for 50 min. Glucose and 3‐OMG were extracted by liquid–liquid extraction and separated on a YMC‐Triart C18 column, with a gradient mobile phase composed of acetonitrile and 20 mm ammonium acetate solution containing 0.09% tri‐ethylamine at a flow rate of 1.0 mL/min. The eluate were detected using a UV detector at 250 nm. The assay was linear over the range 0.39–25 μm (R2 = 0.9997, n = 5) and the lower limit of quantitation was 0.39 μm (0.070 mg/mL). Intra‐ and inter‐day precision and accuracy were <15% and within ±3%, respectively. After validation, the HPLC method was applied to investigate the gluconeogenesis in Dulbecco's modified Eagle medium (DMEM) cultured HepG2 cells. Glucose concentration was determined to be about 1–2.5 μm in this gluconeogenesis assay. In conclusion, this method has been shown to determine small amounts of glucose in DMEM successfully, with lower limit of quantitation and better sensitivity when compared with common commercial glucose assay kits. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

11.
A new route for the synthesis of substituted 8‐methyl‐6‐phenyl‐5,6‐dihydro‐4H‐1,3,2‐benzodioxaphosphocine‐2‐oxide derivatives has been developed by using cinnamic acid and p‐cresol via condensation, reduction, and followed by phosphorylation steps. The title compounds were characterized by IR, 1H, 13C, 31P, and mass spectral studies and elemental analysis. The title compounds have been investigated for their antioxidant activity with respect to their IC50 values using 2,2‐diphenyl‐1‐picrylhydrazyl, NO radical scavenging activities, and reducing power assay. The results obtained from the aforementioned methods revealed that 2‐phenylamino derivatives have shown greater free radical scavenging activity when compared with those of the phenoxy derivatives and is attributed to the presence of secondary amino group, which is able to produce free radicals easily.  相似文献   

12.
The substituent effect of electron‐withdrawing groups on electron affinity and gas‐phase basicity has been investigated for substituted propynl radicals and their corresponding anions. It is shown that when a hydrogen of the α‐CH3 group in the propynyl system is substituted by an electron‐withdrawing substituent, electron affinity increases, whereas gas‐phase basicity decreases. These results can be explained in terms of the natural atomic charge of the terminal acetylene carbon of the systems. The calculated electron affinities are 3.28 eV (?C?C? CH2F), 3.59 eV (?C?C? CH2Cl) and 3.73 eV (?C?C? CH2Br), and the gas‐phase basicities of their anions are 359.5 kcal/mol (?:C?C? CH2F), 354.8 kcal/mol (:C?C? CH2Cl) and 351.3 kcal/mol (?:C?C? CH2Br). It is concluded that the larger the magnitude of electron‐withdrawing, the greater is the electron affinity of radical and the smaller is the gas‐phase basicity of its anion. © 2009 Wiley Periodicals, Inc. J Comput Chem, 2009  相似文献   

13.
New heterocyclic compounds 1‐(3‐methyl‐9H‐dibenzo[b,f][1,2,4]triazolo[4,3‐d][1,4]diazepin‐6‐yl)ethanone 8a , 1‐(3‐methyldibenzo[b,f][1,2,4]triazolo[4,3‐d][1,4]oxazepin‐6‐yl)ethanone 8b , and 1‐(3‐methyldibenzo[b,f][1,2,4]triazolo[4,3‐d][1,4]thiazepin‐6‐yl)ethanone 8c are synthesized from benzodiazepinone, benzoxazepinone, and benzothiazepinone derivatives. These heterocyclic scaffolds have wide medicinal importance. Best results were obtained in antibacterial screening against Escherichia coli, Enterobacter cloacae, and Staphylococcus aureus and antifungal screening against Candida albicans and Fusarium oxysporum. 1,1‐Diphenyl‐2‐picrylhydrazyl radical scavenging activities of compounds 6c , 7c , and 8c were tested in doses 10, 20, 30, 40, and 50 μg/mL and were expressed as IC50 values and percent of inhibition with means ± standard deviation of three different concentrations of synthesized compounds. The assignment of the structures of synthesized compounds was made by thin‐layer chromatography, elemental analysis, IR, 1H‐NMR, 13C‐NMR, and liquid chromatography–mass spectrometry.  相似文献   

14.
Polymerization of diazoketones mediated by organoaluminum compounds was investigated. Trialkylaluminum R3Al (R = iBu, Et, Me) and diisobutylaluminum hydride (DIBAL) polymerized (E)‐1‐diazo‐3‐nonen‐2‐one ( 1 ) to give polymers with Mn = 2000–3500, which contained nearly 33 mol % of azo group (? N?N? ) along with the dominant acylmethylene unit in the main chain. On the other hand, when (E)‐1‐diazo‐4‐phenyl‐3‐buten‐2‐one ( 2 ) was used as a monomer for the organoaluminum‐mediated polymerization, the resulting polymers had ethylidene (? CH[CH3]? ) units in the main chain along with acylmethylene and azo group, as a result of reductive cleavage of the acyl group during the polymerization. © 2007 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 45: 5209–5214, 2007  相似文献   

15.
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   

16.
We investigated the mechanism of the ring‐opening copolymerization of ?‐caprolactam (?‐CL) with glycidyl phenyl ether (GPE) to afford poly(?‐CL‐co‐GPE) as a model reaction of the thermal curing of certain epoxy resins with ?‐CL. The reaction of ?‐CL and GPE proceeded efficiently in the presence of 1,8‐diazabicyclo[5.4.0]undec‐7‐ene (DBU) at 170°C for 2 h. The monomer reactivities r1 of ?‐CL and r2 of GPE calculated according to the Fineman‐Ross method and the Kelen‐Tüdös method were 0.58 and 5.52, respectively. These values indicate that poly(?‐CL‐co‐GPE) has a pseudo‐block gradient copolymer. Based on these results, we examined the thermal curing reactions of certain epoxy resins with ?‐CL. The corresponding novel cured products were obtained quantitatively, and each of them showed a high glass transition temperature and high thermal stability, presumably due at least in part to a pseudo‐block gradient primary structure resembling that of poly(?‐CL‐co‐GPE). © 2016 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2016 , 54, 2220–2228  相似文献   

17.
A high‐performance liquid chromatography method was developed for simultaneous detection and quantitation of total cysteine, glutathione, homocysteine and cysteinylglycine in human plasma. The two key steps in the analysis are reduction of disulfides and treatment with 1‐benzyl‐2‐chloropyridinium bromide, which rapidly and quantitatively reacts with thiol groups to form stable S‐pyridinium derivatives with intense UV absorption. The derivatives are well separated on a Zorbax SB C18 column using reversed‐phase high‐performance liquid chromatography and monitored at 315 nm. The calibration graphs were linear over concentration ranges covering most experimental and clinical cases with a regression coefficients better than 0.999. The detection and quantitation limits for all analytes were 0.2 and 0.5 µmol/L, respectively. The recoveries were 99.25–101.68%. The intra‐ and interassay imprecisions were 0.88–4.24 and 1.68–5.14%, respectively. The method was applied for plasma samples donated by apparently healthy volunteers. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

18.
A simple and sensitive liquid chromatography‐tandem mass spectrometry (LC–MS/MS) method has been developed and validated for the quantitation of exemestane (Exe) and its main metabolite 17‐dihydroexemestane (DhExe) in human plasma. The analytes were extracted by protein precipitation with acetonitrile, containing stable 13C‐labelled Exe (13C3‐Exe) as internal standard, and measured by LC–MS/MS. The best chromatographic separationof the analytes from the interferences was achieved by using a Phenyl column operating under isocratic regime conditions. The total chromatographic runtime was 5.0 min and the elution of Exe and DhExe occurred at 2.5 min and 2.9 min, respectively. Quantitation was performed by employing the positive electrospray ionization (ESI) technique and multiple reaction monitoring mode (MRM). The monitored precursor to product‐ion transitions for Exe, DhExe and 13C3‐Exe internal standard were m/z 297.0 → 120.8, m/z 299.1 → 134.9 and m/z 300.0 → 123.2, respectively. The lower limit of quantitation (LLOQ) was 0.1 ng/ml for DhExe and 0.2 ng/ml for Exe. The method was linear up to 36–51 ng/ml with r2 ≥ 0.998. The intra‐ and inter‐assay precision were ≤7.7% and 5.1% for Exe and ≤8.1 and 4.9% for DhExe while deviations from nominal values were in the 1.5–13.2% and ? 9.0–5.8% ranges for Exe and DhExe, respectively. The analytical method resulted robust and suitable for pharmacokinetic monitoring of Exe and its main metabolite during adjuvant therapy in patients with breast cancer. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

19.
A robust method for the quantitation of epigallocatechin gallate peracetate in plasma for pharmacokinetic studies is lacking. We have developed a validated method to quantify this compound using liquid chromatography with quadrupole time‐of‐flight mass spectrometry with isopropanol and tert‐butyl methyl ether (3:10) extraction and thin‐layer chromatography purification. The epigallocatechin gallate peracetate‐1‐13C8 isotope was used as an internal standard. The linear range (r2 > 0.9950) was from 0.05 to 100.00 μg/mL. The lower limit of quantification of the method was 0.05 μg/mL. Reproducibility, coefficient of variation, was between 0.7 and 12.6% (n = 6), accuracy between 83.7 and 104.6% (n = 5), and recovery ranged from 82.4 to 109.0% (n = 4). Ion suppression was approximately 40%. No mass spectral peaks were found to interfere between the standard and internal standard or the blank plasma extracts. Epigallocatechin gallate peracetate in plasma was stably stored at –80°C over three months even after three freeze–thaw cycles. Extracts were stable in the sampler at 4°C for over 48 h. Plasma levels were maintained at 1.36 μg/mL for 360 min after intraorbital intravenous injection at 50 mg/kg in mice. This method can be used to reliably measure epigallocatechin gallate peracetate in plasma for pharmacokinetic studies.  相似文献   

20.
Upon activation with diethylaluminium chloride (Et2AlCl), a series of phenyl‐substituted α‐diimine nickel precatalysts conducted 4‐methyl1pentene (4MP) and ethylene (E) (co)polymerizations via controlled chain‐walking to generate branched amorphous polymers with high molecular weight and narrow molecular weight distribution (Mw/Mn < 1.6). The obtained poly(4MP)s were amorphous elastomers with glass transition temperature (Tg) of ?10 ~ ?24 °C, which are higher than that of E‐4MP copolymer ( ? 63.0 °C). At room temperature (25 °C), 4MP polymerization proceeds in a living manner. The microstructures of the produced poly(4MP)s indicated the 2,1 and 1,2insertion followed by chain‐walking, the latter being predominant. The NMR analyses of the polymers showed that the obtained poly(4MP) possessed methyl, isobutyl, 2,4dimethylpentyl and 2methylhexyl groups, while the isobutyl and 2,4dimethylalkyl branches derived from 4MP were observed in the E‐4MP copolymer. The branch structures and the insertiontype of monomer were depended on the polymerization temperature, and the content of methyl branch increased with an increase in the polymerization temperature.  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