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 46 毫秒
1.
N-n-Propyl-2-pyridylmethanimine, 1, N-n-octyl-2-pyridylmethanimine, 2, N-n-lauryl-2-pyridylmethanimine, 3, and N-n-octadecyl-2-pyridylmethanimine, 4 have been used in conjunction with copper(II) bromide and azo initiators for the reverse atom transfer radical polymerisation of a range of methacrylates. AIBN to CuIIBr2 ratios of 0.5:1, 0.75:1 and 1:1 give PMMA with Mn 11 500 g mol−1 (PDi = 1.24) (at 22% conversion), 12 500 g mol−1 (PDi = 1.06) (at 83% conversion) and 10 900 g mol−1 (PDi = 1.11) (at 84% conversion), respectively. A CuIIBr2 complex is demonstrated to be needed at the start of the reaction for good control over molecular weight and polydispersity as reactions using Cu(I)Br as catalyst yielded PMMA of Mn 31 000 g mol−1 (PDi = 2.90), reactions with no copper yield PMMA of Mn 33 000 g mol−1 (PDi = 2.95). The RATRP of styrene was carried out using CuIIBr2 as catalyst. AIBN to CuIIBr2 ratio of 0.5:1, 0.75:1 and 1:1 gave PS with Mn = 12 400 g mol−1 (PDi = 1.27) at low conversion, Mn = 15 500 g mol−1 (PDi = 1.11) and 12 400 g mol−1 (PDi = 1.38), respectively at ∼85% conversion. A series of block copolymers of MMA with BMA, BzMA and DMEAMA (15 600 g mol−1 (PDi = 1.18), 13 300 g mol−1 (PDi = 1.14) 15 300 g mol−1 (PDi) = 1.16), using a PMMA macroinitiator were prepared. Emulsion polymerisation of MMA using [initiator]:[Cu(II)Br2] ratio = 0.5:1 with Brij surfactant gave a linear increase of Mn with respect to conversion, final Mn = 112 800 g mol−1 (PDi = 1.42). Further reactions were carried out with [initiator]:[Cu(II)Br2] ratio = 0.75:1 and 1:1. Both giving PMMA with Mn ∼ 32 000 g mol−1 (PDi ∼ 2.4). These reactions exhibit no control, this is because the azo initiator is present in excess and all of the monomer is consumed by a free radical polymerisation as opposed to a controlled reaction. Particle size analysis (DLS) showed the particle size between 160 and170 nm in all cases.  相似文献   

2.
In this work, the syntheses of poly(butyl methacrylate-b-methyl methacrylate-b-butyl methacrylate) triblock copolymer and poly(methyl methacrylate-b-butyl methacrylate-b-methyl methacrylate-b-butyl methacrylate-b-methyl methacrylate) pentablock copolymers using copper mediated living radical polymerisation are reported. Living radical polymerisations were performed using the system CuIBr/N-(n-propyl)-2-pyridylmethanimine as catalyst in conjunction with a difunctional initiator, the 1,4-(2-bromo-2-methylpropionoto)benzene (1). The syntheses of poly(MMA), poly(BMA-b-MMA-b-BMA) and poly(MMA-b-BMA-b-MMA-b-BMA-b-MMA) are described in detail using 1H NMR spectroscopy and size exclusion chromatography. The living behaviour and the blocking efficiency of these polymerisations were investigated in each case. Difunctional initiator, 1, based on hydroquinone was synthesised and fully characterised and subsequently used to give difunctional poly(methyl methacrylate) macroinitiators with molecular weights up to 54,000 g mol−1 and polydispersity between 1.07 and 1.32; molecular weights were close to the theoretical values. The difunctional macroinitiators were used to reinitiate butyl methacrylate to give triblock copolymers of Mn between 17,500 and 45,700 g mol−1. Polydispersities remained narrow below 25,000 g mol−1 but broadened at higher masses. The difunctional triblock macroinitiators were subsequently used to reinitiate methyl methacrylate to give ABABA pentablock copolymers with Mn up to 37,000 g mol−1 with polydispersity=1.13. Under certain conditions radical-radical reaction led to a broadening of polydispersity index.  相似文献   

3.
A series of AB and ABA block copolymers of pDEGMEMA-b-pCHMA and pCHMA-b-pDEGMEMA-b-pCHMA cyclohexyl methacrylate (CHMA) and di(ethylene glycol) methyl ether methacrylate (DEGMEMA) with Mn ranging between 18,000 and 50,000 g mol−1 and PDI = 1.09-1.32 were prepared via copper(I) mediated living radical polymerization with pyridylmethanimine ligands. Aggregation properties were investigated using a combination of 1H NMR, dynamic and static light scattering. For comparative purposes poly(CHMA) and poly(DEGMEMA) homopolymers were prepared. The CAC values estimated for the di- and triblock copolymers soluble in cyclohexane are lower than 0.005 g L−1 whereas the values found for block copolymers in methanol solutions are less than 0.070 g L−1. DLS analysis showed the presence of micellar aggregates with diameters ranging from 25 to 40 nm with particle polydispersity indexes between 0.003 and 0.183. The pCHMA-b-pDEGMEMA-b-pCHMA micelles solubilized the aqueous phase in petrol/gasoline. The block copolymer-based micelles incorporate water within their hydrophilic domains, potentially overcoming a number of practical problems such as the formation of biphasic mixtures in solvent blends due to undesired water accumulation.  相似文献   

4.
Asymmetric poly(styrene-b-methyl methacrylate) (PS-b-PMMA) diblock copolymers of molecular weight Mn = 29,700 g mol−1 (MPS = 9300 g mol−1MPMMA = 20,100 g mol−1, PD = 1.15, χPS = 0.323, χPMMA = 0.677) and Mn = 63,900 g mol−1 (MPS = 50,500 g mol−1, MPMMA = 13,400 g mol−1, PD = 1.18, χPS = 0.790, χPMMA = 0.210) were prepared via reversible addition-fragmentation chain transfer (RAFT) polymerization. Atomic force microscopy (AFM) was used to investigate the surface structure of thin films, prepared by spin-coating the diblock copolymers on a silicon substrate. We show that the nanostructure of the diblock copolymer depends on the molecular weight and volume fraction of the diblock copolymers. We observed a perpendicular lamellar structure for the high molar mass sample and a hexagonal-packed cylindrical patterning for the lower molar mass one. Small-angle X-ray scattering investigation of these samples without annealing did not reveal any ordered structure. Annealing of PS-b-PMMA samples at 160 °C for 24 h led to a change in surface structure.  相似文献   

5.
Copolymerization of an excess of methyl methacrylate (MMA) relative to 2-hydroxyethyl methacrylate (HEMA) was carried out in toluene at 80 °C according to both conventional and controlled Ni-mediated radical polymerizations. Reactivity ratios were derived from the copolymerization kinetics using the Jaacks method for MMA and integrated conversion equation for HEMA (rMMA = 0.62 ± 0.04; rHEMA = 2.03 ± 0.74). Poly(ethylene glycol) α-methyl ether, ω-methacrylate (PEGMA, Mn = 475 g mol−1) was substituted for HEMA in the copolymerization experiments and reactivity ratios were also determined (rMMA = 0.75 ± 0.07; rPEGMA ∼ 1.33). Both the functionalized comonomers were consumed more rapidly than MMA indicating the preferred formation of heterogeneous bottle-brush copolymer structures with bristles constituted by the hydrophilic (macro)monomers. Reactivity ratios for nickel-mediated living radical polymerization were comparable with those obtained by conventional free radical copolymerization. Interactions between functional monomers and the catalyst (NiBr2(PPh3)2) were observed by 1H NMR spectroscopy.  相似文献   

6.
The thermo-sensitive swelling behaviour of hydrogels based on 2-(2-methoxyethoxy)ethyl methacrylate (MEO2MA) and synthesized by free radical polymerization has been investigated. The homopolymer hydrogel presents a low critical solution temperature (LCST) close to room temperature, which can be modulated by copolymerization with longer oligo(ethylene glycol) side chain methacrylates (OEGxMA). Then, three series of copolymeric hydrogels synthesized with MEO2MA and several low ratios of OEGxMA with Mn = 475 g mol−1 (OEG8MA), Mn = 1100 g mol−1 (OEG23MA) and Mn = 2080 g mol−1 (OEG45MA) were studied. In addition to conventional tetra(ethylene glycol) dimethacrylate (TEGDMA) crosslinker, the use of biodegradable oligo(caprolactone) dimethacrylate (OCLDMA) was also tested. The hydrophilic/hydrophobic balance, function of the short and the long OEG side chains, establishes a swelling behaviour depending on monomer composition, side chain length and temperature. The swelling at equilibrium increases with increasing the amount of OEGxMA in the copolymer and, at the same time, the collapsing moves progressively to higher temperature. The temperature dependent volumetric response of some of these hydrogels can be compare with the most extended thermo-sensitive hydrogel, which is based on poly(N-isopropylacrylamide) (P(N-iPAAm)). Therefore, they are potential candidates to replace it in applications where biocompatibility is required.  相似文献   

7.
Well-defined poly(MMA-b-DMS-b-MMA) triblock copolymers were prepared by copper(I) mediated living radical polymerization. This was achieved by polymerization of methylmethacrylate (MMA) with different concentrations of 2-bromoisobutyrate terminated polydimethylsiloxane (PDMS). The polymerization occurred in controlled manner with the molecular weight found by 1H NMR close to that predicted and a narrow molecular weight distribution (Mw/Mn∼1.2). Copolymers were obtained with Mn=2100, 4900, 10 100 and 29 500 g mol−1 respectively with poly(MMA) (PMMA) terminal blocks and a central PDMS block of 5500 g mol−1 in each case.DSC analysis showed most of the poly(MMA-b-DMS-b-MMA) triblock copolymers exhibits two Tg’s, one at low temperature corresponding to the Tg of PDMS microphase and a second at high temperature corresponding to the Tg of the PMMA microphase. TEM images show microphase segregation morphology in bulk for the triblock copolymers, with a higher degree of segregation for copolymers containing higher PDMS content. XPS measurements were performed to determine the chemical composition at the surface. For all the copolymers PDMS enrichment is observed at the surface. Copolymers containing higher percentage of PDMS exhibit higher phase separation and better enrichment of PDMS at the surface. The surface tension determined by contact angle measurements of the copolymer film containing 59 mol% of PDMS was 19.15 mN m−1.  相似文献   

8.
This study describes the miscibility phase behavior in two series of biodegradable triblock copolymers, poly(l-lactide)-block-poly(ethylene glycol)-block-poly(l-lactide) (PLLA-PEG-PLLA), prepared from two di-hydroxy-terminated PEG prepolymers (Mn = 4000 or 600 g mol−1) with different lengths of poly(l-lactide) segments (polymerization degree, DP = 1.2-145.6). The prepared block copolymers presented wide range of molecular weights (800-25,000 g mol−1) and compositions (16-80 wt.% of PEG). The copolymer multiphases coexistance and interaction were evaluated by DSC and TGA. The copolymers presented a dual stage thermal degradation and decreased thermal stability compared to PEG homopolymers. In addition, DSC analyses allowed the observation of multiphase separation; the melting temperature, Tm, of PLLA and PEG phases depended on the relative segment lengths and the only observed glass transition temperature (Tg) in copolymers indicated miscibility in the amorphous phase.  相似文献   

9.
In this study, the oxidative polycondensation reaction conditions of 2-[(4-fluorophenyl) imino methylene] phenol (FPIMP) with air oxygen and NaOCl were studied in an aqueous alkaline medium between 60 and 90 °C. Synthesized oligo-2-[(4-fluorophenyl) imino methylene] phenol was characterized by 1H-NMR, FT-IR, UV-Vis, size exclusion chromatography (SEC) and elemental analysis techniques. The yield of oligo-2-[(4-fluorophenyl) imino methylene] phenol (OFPIMP) was found to be 62.00% (for air O2 oxidant) and 97.70% (for NaOCl oxidant) at the optimum reaction conditions. According to the SEC analysis, the number-average molecular weight (Mn), weight-average molecular weight (Mw) and polydispersity index (PDI) values of OFPIMP were found to be 1370 g mol−1, 1979 g mol−1 and 1.45, using NaOCl, 2105 g mol−1, 2557 g mol−1, and 1.22, using air O2, respectively. During the oxidative polycondensation reaction, (2.88%) a part of -CHN group oxidized to carboxylic acid (-COOH). TG and TG-DTA analyses were shown to be more stable of oligo-2-[(4-fluorophenyl) imino methylene] phenol and its oligomer metal complexes than monomer against thermo-oxidative decomposition. The weight loss of OFPIMP was found to be 97.00% at 900 °C. The weight losses of OFPIMP-Co, OFPIMP-Ni OFPIMP-Cu oligomer-metal complex compounds were found to be 88.66%, 94.36% and 83.21%, respectively, at 1000 °C.  相似文献   

10.
Two nickel(II) complexes (A and B) bearing β-iminoamine ligands, [2-(ArNCH)-C6H4-NMe2] (La, Ar = 2,6-i-Pr2C6H3; Lb, Ar = 2,6-Me2C6H3), were synthesized and characterized by elemental analyses and 1H NMR. X-ray crystal structure of complex B reveals that the six-membered chelate ring adopts a envelope conformation, with nickel(II) atom deviating from the plane of backbone aromatic ring by 1.164 Å. In the presence of methylaluminoxane (MAO), both complexes showed moderate activities of 105 g molNi−1 h−1 for norbornene polymerization. β-iminoamine Ni(II)/MAO catalysts gave unimodal polymers (Mw, 3.16-8.02 × 10g/mol) with a relatively narrow MWD (Mw/Mn, 1.59-2.14), indicative of single-site catalyst behavior. The obtained polymers are vinyl-type polynorbornenes (PNBs), which are soluble in common solvents such as toluene, cyclohexane and dichlorobenzene.  相似文献   

11.
This contributions shows with a series of ab initio MP2 and DFT (BP86 and B3-LYP) computations with large basis sets up to cc-pVQZ quality that the literature value of the standard enthalpy of depolymerization of Sb4F20(g) to give SbF5(g) (+18.5 kJ mol−1) [J. Fawcett, J.H. Holloway, R.D. Peacock, D.R. Russell, J. Fluorine Chem. 20 (1982) 9] is by about 50 kJ mol−1 in error and that the correct value of (Sb4F20(g)) is +68 ± 10 kJ mol−1. We assign , , and values for SbnF5n with n = 2-4 and compare the results to available experimental gas phase data. Especially the MP2/TZVPP values obtained in an indirect procedure that rely on isodesmic reactions or the highly accurate compound methods G2 and CBS-Q are in excellent agreement with the experimental data, and reproduce also the fine experimental details at temperatures of 423 and 498 K. With these data and the additional calculation of [SbnF5n+1] (n = 1-4), we then assessed the fluoride ion affinities (FIAs) of SbnF5n(g), nSbF5(g), nSbF5(l) and the standard enthalpies of formation of SbnF5n(g) and [SbnF5n+1](g): FIA(SbnF5n(g)) = 514 (n = 1), 559 (n = 2), 572 (n = 3) and 580 (n = 4) kJ mol−1; FIA(nSbF5(g)) = 667 (n = 2), 767 (n = 3) and 855 (n = 4) kJ mol−1; FIA(nSbF5(l)) = 434 (n = 1), 506 (n = 2), 528 (n = 3) and 534 (n = 4) kJ mol−1. Error bars are approximately ±10 kJ mol−1. Also the related Gibbs energies were derived. ΔfH°([SbnF5n+1](g)) = −2064 ± 18 (n = 1), −3516 ± 25 (n = 2), −4919 ± 31 (n = 3) and −6305 ± 36 (n = 4) kJ mol−1.  相似文献   

12.
Poly(acrylamide) (PAM) with controlled molecular weight and tacticity was prepared by UV-irradiation-initiated controlled/living radical polymerization in the presence of dibenzyl trithiocarbonate (DBTTC) and Y(OTf)3. The rapid and facile photo-initiated controlled/living polymerization at ambient temperature led to controlled molecular weight and narrow polydispersity (Mw/Mn = 1.12-1.24) of PAM. The coordination of Y(OTf)3 with the last two amide groups in the growing chain radical effectively enhanced isotacticity of PAM. The isotactic sequence of dyads (m), triads (mm) and pentads (mmmm) in PAM were 70.32%, 50.95%, and 29.97%, respectively, which were determined by the resonance of methine (CH) groups in PAM under 13C NMR experiment. Factors affecting stereocontrol during the polymerization were studied, including the type of Lewis acids, concentration of Y(OTf)3, and monomer conversion. It is intriguing that the meso tacticity increased gradually with chain propagation and quite higher isotacticity (m = 93.01%, mm = 86.57%) was obtained in the later polymerization stage (conversion 65-85%).  相似文献   

13.
The transfer constant to benzene in the radical polymerisation of di-n-butyl itaconate, DBI, was measured at 60 °C and found to have the value of 2.2 × 10−3. This value is surprisingly high considering that benzene has no readily extractable hydrogen atoms and is usually considered to be a relatively inert solvent in radical polymerisations. The high values of transfer constants in the radical polymerisation of itaconates must be taken into consideration when planning the tailored polymerisation of these monomers.  相似文献   

14.
2-Dimethylaminoethyl methacrylate (DMAEMA) and 2-diethylaminoethyl methacrylate (DEAEMA) block copolymers have been synthesized by using poly(ethylene glycol), poly(tetrahydrofuran) (PTHF) and poly(ethylene butylenes) macroinitiators with copper mediated living radical polymerization. The use of difunctional macroinitiator gave ABA block copolymers with narrow polydispersities (PDI) and controlled number average molecular weights (Mn’s). By using DMAEMA, polymerizations proceed with excellent first order kinetics indicative of well-controlled living polymerization. Online 1H NMR monitoring has been used to investigate the polymerization of DEAEMA. The first order kinetic plots for the polymerization of DEAMA showed two different rate regimes ascribed to an induction period which is not observed for DMAEMA. ABA triblock copolymers with DMAEMA as the A blocks and PTHF or PBD as B blocks leads to amphiphilic block copolymers with Mn’s between 22 and 24 K (PDI 1.24-1.32) which form aggregates/micelles in solution. The critical aggregation concentrations, as determined by pyrene fluorimetry, are 0.07 and 0.03 g dm−1 for PTHF- and PBD-containing triblocks respectively.  相似文献   

15.
Synthesis of aromatic poly(ether ketone) (3) with a narrow molecular weight distribution (Mw/Mn) was investigated via polycondensation. Mns of 3 could be controlled varying the feed ratio of monomer (1) and initiator (2) maintaining relatively narrow Mw/Mns (<1.3). The kinetics of polycondensation obeyed a first-order relationship between polycondensation time and -(1/[2]0) ln([1]/[1]0), and the rate of polycondensation was estimated as 2.57 mol−1 L h−1. MALDI-TOF mass analysis of 3 indicated that polycondensation should proceed via chain growth manner to give 3 having an initiator unit in each chain end.  相似文献   

16.
A facile synthesis of poly(lauryl acrylate) has been achieved by atom transfer radical polymerization using benzyl-2-bromoisobutyrate, copper (I) bromide, and N-(n-octyl)-2-pyridylmethanimine (OPMI). The latter was of great interest as its synthesis was very easy to carry out and as it allowed the reaction mixture to be homogeneous, which was essential for the control of the reaction. The polymerization was controlled under these conditions and was optimized with the addition of copper (II) bromide as deactivator. We proved that the synthesis of poly(lauryl acrylates) with well defined molecular weights and narrow polydispersities was possible using a ligand which does not require difficult synthesis and purification. We also showed the ability of pyridylmethanimine ligands to control ATRP of an acrylate derivative. Best results were obtained at 130 °C in xylene for [Initiator]0/[Cu(I)Br]0/[Cu(II)Br2]0/[OPMI]/[lauryl acrylate] equal to 1/1/0.05/2.2/181, respectively (Mn = 19,942, DPI = 1.28).  相似文献   

17.
The dynamic properties of polyethylene glycol (PEG) molecules on the solid–liquid interface oscillating at MHz were investigated using the quartz crystal microbalance (QCM). The number-average molecular weights (Mn) of the PEG molecules were systematically varied over 4 orders of magnitude. This study makes it clear that the series-resonant frequency shift, ΔF, of the QCM against the square root of the density–viscosity product of the PEG solution is linear and has the intercept. Moreover, systematical analysis reveals that the ΔF slope rapidly decreases with Mn and that the ΔF intercept becomes constant above 4.0 × 103 g mol−1. As a result, those reveal that the resonant length of the PEG molecule moving with the oscillating plate of 9 MHz is 54.2 Å. We also find that the behaviors of ΔF due to Mn are mainly caused by the length of the PEG molecule.  相似文献   

18.
Chlorophyll derivative possessing a trifluoroacetyl group at the 3-position was synthesized as a new chemosensor for alcohols and amines. Intense Qy peak of the trifluoroacetyl-chlorin (701 nm in CHCl3) showed blue shifts to 667 nm in MeOH and 665 nm in n-BuNH2 due to the formation of the corresponding hemiacetal and hemiaminal with visible color changes. Thermodynamic parameters for the complexation between trifluoroacetyl-chlorin and n-BuNH2 in CDCl3 were determined to be ΔH = −48 kJ mol−1 and ΔS = −147 J K−1 mol−1. Ratiometric fluorescence sensing of n-BuNH2 in THF was also demonstrated.  相似文献   

19.
In this study, the reaction conditions of poly-4-[(2-methylphenyl)iminomethyl]phenol (P-2-MPIMP) were studied by using oxidants such as air O2, H2O2 and NaOCl in an aqueous alkaline medium between 50 and 90 °C. The structures of the synthesized monomer and polymer were confirmed by FT-IR, UV-vis, NMR and elemental analysis. The characterization was made by TG-DTA, size exclusion chromatography (SEC) and solubility tests. At the optimum reaction conditions, the yield of poly-4-[(2-methylphenyl)iminomethyl]phenol (P-2-MPIMP) was found to be 20% (for air O2 oxidant), 33% (for H2O2 oxidant), and 74% (for NaOCl oxidant). According to the SEC analysis, the number-average molecular weight (Mn), weight-average molecular weight (Mw) and polydispersity index (PDI) values of P-2-MPIMP were found to be 3300, 4100 g mol−1 and 1.242, using H2O2, and 4550, 5150 g mol−1and 1.132, using air O2 and 5300, 5850 g mol−1 and 1.104, using NaOCl, respectively. According to TG analysis, the weight losses of 4-[(2-methylphenyl)iminomethyl]phenol (2-MPIMP) and P-2-MPIMP were found to be between 75.29% and 48.17% at 1000 °C, respectively. P-2-MPIMP was shown to have a higher stability against thermal decomposition. Also, electrical conductivity of the P-2-MPIMP was measured, showing that the polymer is a typical semiconductor. Electrochemically, the highest occupied molecular orbital (HOMO), the lowest unoccupied molecular orbital (LUMO) and electrochemical energy gaps ( of 2-MPIMP and P-2-MPIMP were found to be −6.01, −6.03; −2.63, −2.82; 3.38 and 3.21 eV, respectively. According to UV-vis measurements, the optical band gap (Eg) of 2-MPIMP and P-2-MPIMP was found to be 3.40 and 2.97 eV, respectively.  相似文献   

20.
Enthalpies for the two proton ionizations of glycine, N,N-bis(2-hyroxyethyl)glycine (“bicine”) and N-tris(hydroxymethyl)methylglycine (“tricine”) were obtained in water-methanol mixtures with methanol mole fraction (Xm) from 0 to 0.360. With increasing methanol the ionization enthalpy for the first proton (ΔH1) of glycine increased from 4.4 to 9.4 kJ mol−1 with a minimum of 4.1 kJ mol−1 at Xm = 0.059. The ionization enthalpy of the second proton (ΔH2) for glycine decreased from 46.3 to 38.1 kJ mol−1. ΔH1 of bicine increased from 3.5 to 7.6 kJ mol−1 at Xm = 0.273 before dropping to 4.1 kJ mol−1 at Xm = 0.360. ΔH2 of bicine increased from 24.9 to 29.4 kJ mol−1. For tricine, ΔH1 increased from 6.7 to 9.8 kJ mol−1 at Xm = 0.194 then dropped to 7.4 kJ mol−1 at Xm = 0.360. ΔH2 for tricine first dropped from 30.8 to 28.5 kJ mol−1 at Xm = 0.059 before increasing to 33.3 kJ mol−1 at Xm = 0.273. The solvent composition was selected so as to include the region of maximum structure enhancement of water by methanol. The results were interpreted in terms of solvent-solvent and solvent-solute interactions.  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