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.
Novel green and sustainable elastomers having both good biodegradability and chemical recyclability properties were designed and synthesized using potentially biobased materials and lipase as an environmentally benign catalyst. High molecular weight poly[(12-hydroxydodecanoate)-co-(12-hydroxystearate)] [poly(12HD-co-12HS)] samples with varying monomer ratios were prepared by the polycondensation of 12-hydroxydodecanoic acid and methyl 12-hydroxystearate using immobilized lipase from Candida antarctica (IM-CA) in toluene in the presence of molecular sieves 4A at 90 degrees C. Although poly(12HD) is a highly crystalline polyester having a melting temperature (T(m)) of 87.6 degrees C and crystalline temperature (T(c)) of 64 degrees C, by the copolymerization of 12HD with 12HS, both the T(m) and T(c) of the copolymer decreased with increasing 12HS contents, and poly(12HD-co-12HS) containing more than 60 mol-% 12HS was a viscous liquid at room temperature. At the same time, the Young's modulus and hardness also decreased with increasing 12HS content, and poly(12HD-co-36 mol-% 12HS) exhibited an elastic behavior, having a hardness of 70 A using a durometer A. In addition, it showed an excellent biodegradability by activated sludge and chemical recyclability by lipase.  相似文献   

2.
Medium to high molecular weight random copolymers of 1,5-dioxepan-2-one (DXO) and L-lactide (L-LA) or ε-caprolactone (ε-CL) of different compositions have been investigated. Polymerization was conducted in bulk at 110°C using stannous 2-ethylhexanoate as catalyst. Poly(DXO-co-L-LA) is a hydrolysable material with a glass transition temperature (Tg) ranging from −36 up to 58°C depending on the molar composition. The material exhibited crystallinity as long as the amount of DXO did not exceed 50 weight%. Reactivity ratios were determined to rL-LA=10 and rDXO=0.1, giving a more blocky structure than expected in a random copolymer. The copolymer between ε-CL and DXO was shown to be a truly random copolymer by 13C NMR, as expected from the reactivity ratios, rDXO=1.6 and rε-CL=0.6. Tg of the material was ranging from −64 up to −39°C. The ability of the poly(DXO-co-ε-CL) to crystallize was retained up to a DXO content of 40 weight%. The melting temperature and crystallinity of both copolymers decrease with increasing amount of DXO. Incorporation of semicrystalline comonomers, L-LA or ε-CL, into the amorphous poly(DXO) creates materials with adjustable properties depending on the molar composition.  相似文献   

3.
Highly isotactic polylactide or poly(lactic acid) is synthesized in a ring-opening polymerization (ROP) of racemic lactide with achiral salen- and homosalen-aluminum complexes (salenH(2)=N,N'-bis(salicylidene)ethylene-1,2-diamine; homosalenH(2)=N,N'-bis(salicylidene)trimethylene-1,3-diamine). A systematic exploration of ligands demonstrates the importance of the steric influence of the Schiff base moiety on the degree of isotacticity and the backbone for high activity. The complexes prepared in situ are pure enough to apply to the polymerizations without purification. The crystal structures of the key complexes are elucidated by X-ray diffraction, which confirms that they are chiral. However, analysis of the (1)H and (13)C NMR spectra unambiguously demonstrates that their conformations are so flexible that the chiral environment of the complexes cannot be maintained in solution at 25 degrees C and that the complexes are achiral under the polymerization conditions. The flexibility of the backbone in the propagation steps is also documented. Hence, the isotacticity of the polymer occurs due to a chain-end control mechanism. The highest reactivity in the present system is obtained with the homosalen ligand with 2,2-dimethyl substituents in the backbone (ArCH==NCH(2)CMe(2)CH(2)N==CHAr), whereas tBuMe(2)Si substituents at the 3-positions of the salicylidene moieties lead to the highest selectivity (P(meso)=0.9(8); T(m)=210 degrees C). The ratio of the rate constants in the ROPs of racemic lactide and L-lactide is found to correlate with the stereoselectivity in the present system. The complex can be utilized in bulk polymerization, which is the most attractive in industry, although with some loss of stereoselectivity at high temperature, and the afforded polymer shows a higher melting temperature (P(meso)=0.9(2), T(m) up to 189 degrees C) than that of homochiral poly(L-lactide) (T(m)=162-180 degrees C). The "livingness" of the bulk polymerization at 130 degrees C is maintained even at a high conversion (97-98 %) and for an extended polymerization time (1-2 h).  相似文献   

4.
We studied the thermal diffusion behavior of poly(N-isopropylacrylamide) (PNiPAM) in ethanol in a temperature range from T = 14.0 degrees C to T = 40.0 degrees C by means of thermal diffusion forced Rayleigh scattering. The obtained Soret coefficient S(T) of PNiPAM was positive for lower temperatures (T < 34 degrees C), while S(T) showed a negative value for higher temperatures (T > 34 degrees C). This means PNiPAM molecules move to the cold side for temperatures T < 34 degrees C, whereas they move to the warm side for T > 34 degrees C. This is the first nonaqueous polymeric system for which a sign change with temperature has been observed. We performed static and dynamic light scattering experiments in the same temperature range. The second virial coefficient determined from dilute solutions by static light scattering (SLS) was positive in the comparable temperature range. The results of SLS for the semidilute solution showed a strong repulsion among PNiPAM chains which was enhanced by increasing temperature. These results imply that the observed thermally induced sign change of S(T) does not depend on the intermolecular interactions among PNiPAM chains.  相似文献   

5.
We present the first pressure-versus-temperature phase diagram for the helix-to-coil transition of double stranded nucleic acids. The thermodynamic stability of a nucleic acid duplex is a complex function of temperature and pressure and strongly depends on the denaturation temperature, T(M), of the duplex at atmospheric pressure. Depending upon T(M), pressure, and temperature, the phase diagram shows that pressure may stabilize, destabilize, or have no effect on the conformational state of DNA. To verify the phase diagram, we have conducted high-pressure UV melting experiments on poly(dIdC)poly(dIdC), a DNA duplex, poly(rA)poly(rU), an RNA duplex, and poly(dA)poly(rU), a DNA/RNA hybrid duplex. The T(M) values of these duplexes have been modulated by altering the solution ionic strength. Significantly, at low salt, these three duplexes have helix-to-coil transition temperatures of 50 degrees C or less. In agreement with the derived phase diagram, we found that the polymeric duplexes were destabilized by pressure if the T(M) is < approximately 50 degrees C. However, these duplexes were stabilized by pressure if the T(M) is > approximately 50 degrees C. The DNA/RNA hybrid duplex, poly(dA)poly(rU), with a T(M) of 31 degrees C in 20 mM NaCl undergoes a pressure-induced helix-to-coil transition at room temperature. This is the first report of pressure-induced denaturation of a nucleic acid duplex and provides new insights into the molecular forces stabilizing these structures.  相似文献   

6.
In this work, small-angle neutron scattering (SANS) is used to probe the structural transformations that accompany temperature-induced gelation of emulsions stabilized by a temperature-responsive polymer. The latter is poly(NIPAM-co-PEGMa) (N-isopropylacrylamide and poly(ethyleneglycol) methacrylate) and contains 86 mol% NIPAM. Turbidity measurements revealed that poly(NIPAM-co-PEGMa) has a lower critical solution temperature (T(LCST)) of 36.5 degrees C in D(2)O. Aqueous polymer solutions were used to prepare perfluorodecalin-in-water emulsions (average droplet size of 6.9 mum). These emulsions formed gels at 50 degrees C. SANS measurements were performed on the poly(NIPAM-co-PEGMa) solutions and emulsions as a function of temperature. The emulsion was also prepared using a D2O/H2O mixture containing 72 vol% D2O in order to make scattering from the droplets negligible (on-contrast). The SANS data were analyzed using a combination of Porod and Ornstein-Zernike form factors. The results showed that the correlation length (xi) of the polymer scaled as xi approximately phi(p)(-0.68) at 32 degrees C, where phi(p) is the polymer volume fraction. The xi value increased for all systems as the temperature increased, which was attributed to a spinodal transition. At temperatures greater than T(LCST), the polymer solution changed to a polymer dispersion of poly(NIPAM-co-PEGMa) aggregates. The aggregates have features that are similar to microgel particles. The average size of these particles was estimated as 160-170 nm. The particles are "sticky" and are gel-forming. The on-contrast experiments performed using the emulsion indicated that the interfacial polymer chains condensed to give a relatively thick polymer layer at the perfluorodecalin-water interface at 50 degrees C. The gelled emulsions appear to consist of perfluorodecalin droplets with an encapsulating layer of collapsed polymer to which sticky microgel particles are adsorbed. The latter act as a "glue" between coated droplets in the emulsion gel.  相似文献   

7.
The DNA mediated assembly of complementary DNA-functionalized gold nanoparticles (DNA-AuNP) was investigated by means of UV/Vis-spectroscopy and Dynamic Light Scattering (DLS). The melting temperature of the aggregates was determined to be T(m) = 31 degrees C. Characterization of the assembly at 20 degrees C, 25 degrees C and 30 degrees C showed a decrease of the initial assembly growth rate with increasing temperature. The correlation of the wavelengths at the absorbance maxima lambda(max) and the hydrodynamic radii R(h) of the AuNP assemblies proved the dependence of the optical properties on the assembly size while at higher assembly temperature (30 degrees C) a larger redshift of lambda(max) with increasing R(h) was observed than at lower temperatures. This tendency might give information about the dependence of the internal structure of the DNA-AuNP assemblies on assembly temperature. It is assumed that at higher temperatures more compact assemblies are built than at lower temperatures of 20 degrees C and 25 degrees C. To the best of our knowledge, this is the first systematic time-resolved in situ investigation of DNA-mediated AuNP assembly by UV/Vis-spectroscopy and DLS.  相似文献   

8.
Hydrodynamic and electrophoretic properties of core-shell poly(styrene/alpha- tert-butoxy-omega-vinylbenzyl-polyglycidol) (P(S/PGL)) microspheres suspended in water are described. The microspheres were obtained by surfactant-free emulsion copolymerization of styrene and alpha- tert-butoxy-omega-vinylbenzyl-polyglycidol macromonomer ( M n = 2800, M w/ M n = 1.05). The process yielded microspheres with number average diameter D n = 270 nm and with low diameter dispersity index D w/ D n = 1.01. Shells of P(S/PGL) microspheres were enriched in polyglycidol. Molar fraction of polyglycidol monomeric units in the shells (determined by X-ray photoelectron spectroscopy) was equal to 0.34, which is much higher than the average molar fraction of polyglycidol monomeric units in whole particles of 0.048. Influences of NaCl concentration and temperature on P(S/PGL) microsphere diameters and on their electrophoretic mobility were investigated. It was found that hydrodynamic diameter of P(S/PGL) microspheres, determined by photon correlation spectroscopy, decreased significantly when temperature did exceed a certain value (transition temperature, T t). It has been found that the decrease is more pronounced for higher concentrations of NaCl in the medium. For microspheres suspended in 10 (-1) M NaCl, the hydrodynamic diameter decreased by 8% whereas for the same particles in pure water the diameter decreased by 5.2%. The process of shrinkage was fully reversible. Values of T t for P(S/PGL) microspheres were lower for higher concentrations of NaCl. Adjustment of salt concentration allowed controlling T t in a range from 44.4 to 49.9 degrees C. 13C NMR relaxation time measurements (T 1) for carbon atoms in polyglycidol macromonomer revealed that T 1 did increase with increasing temperature (in temperature range from 25 to 75 degrees C) indicating higher motion of chains at higher temperature. Addition of NaCl did not induce a substantial change of T 1 in the mentioned temperature range. The swelling-deswelling properties of P(S/PGL) microspheres' interfacial layer affected adsorption of P(S/PGL) particles on modified with (3-aminopropyl)triethoxysilane mica. It was shown that the deposition of P(S/PGL) microspheres at 25 degrees C on mica led to formation of two-dimensional crystal-shape assemblies, whereas at 60 degrees C (far above T t = 49.8 degrees C in H2O) the microspheres were randomly adsorbed without formation of colloidal crystal assemblies.  相似文献   

9.
The temperature dependence of molecular mobility and conformational changes of poly(chlorotrifluoro- ethylene) (PCTFE) have been investigated by solid-state (19)F magic angle spinning (MAS) NMR spectroscopy. The pulse techniques of dipolar-filter and T(1rho)-filter allow selective observation of the amorphous and crystalline domains, respectively. The temperature dependence of T(1rho) (F) revealed that the segmental motion in the amorphous domain becomes vigorous above ca 80 degrees C, which is well above the glass transition (T(g)) temperature (52 degrees C) and more close to the beta-relaxation temperature (95 degrees C). On the other hand, vigorous molecular motions in the crystalline domain occur above 120 degrees C, which is much below the melting temperature (212 degrees C). This indicates that the polymer chains in the PCTFE crystallites are more mobile than those of typical semicrystalline fluoropolymers like poly(vinylidene fluoride) (PVDF), which can be associated with structural imperfections in the crystallites. In addition, the density functional theory (DFT) calculations of (19)F magnetic shielding suggest that the high-frequency shifts observed for the crystalline signals above 80 degrees C can be ascribed to the conformational change around meso diads toward more twisted and/or helical conformations in the main chain.  相似文献   

10.
A poly(lactic acid) (PLA)-like terpolyester consisting of 96 mol% lactate (LA), 1 mol% 3-hydroxybutyrate and 3 mol% 3-hydroxyvalerate was produced in recombinant Escherichia coli LS5218 expressing LA-polymerizing enzyme (LPE). The strain was grown on glucose with a feeding of valerate as the monomer precursor. The glass transition and melting temperatures of the terpolyester were close to those of chemically synthesized poly(L-LA)s (PLLAs) having similar molecular weights. Additionally, a blend of the terpolyester, which was composed entirely of (R)-LA (D-LA) due to the strict enantiospecificity of LPE, with PLLA formed a stereocomplex with higher melting temperature (201.9 °C). These results indicate that the biological PLA-like polyester produced via this one-step microbial process has comparable thermal properties to chemically synthesized PLAs.  相似文献   

11.
The segmental dynamics of poly(methyl acrylate-d3) (PMA-d3) adsorbed in the pores of anopore membranes has been investigated using deuterium NMR over the temperature range 25-80 degrees C. The onset of the NMR glass-transition temperature (Tg) for the adsorbed samples was approximately 15 degrees C higher than that for the bulk sample. The adsorbed polymer contained segments with restricted mobility (glassy), even at the highest temperatures studied, at which the bulk polymer showed only mobile segments. The spectra from samples with different adsorbed amounts of PMA-d3, between 1.1 and 4.2 mg/m2, were similar in their temperature-dependent mobilities. Neither was there much difference in the spectra of PMA-d3 on anopore samples with pore sizes of 0.2 and 0.02 microm. However, for a solvent-washed sample with an adsorbed amount of 0.7 mg/m2, additional restriction in PMA-d3 mobility was observed.  相似文献   

12.
Phase-contrast microscopy and particle tracking algorithms are used to study the near-surface diffusion of poly(N-isopropylacrylamide) (PNIPAAm) brush functionalized micron-sized silica microspheres after sedimentation from aqueous suspension onto planar substrates coated with a similar polymer brush above and below the lower critical solution temperature (LCST) of PNIPAAm, 32 degrees C. A small negative charge on the wall and the particles (zeta potential = -6 mV) prevents adhesion above and below the LCST. The near-surface translational diffusion coefficient (D(surface)) is compared to the bulk-phase translational diffusion coefficient (D(bulk)), which was measured by dynamic light scattering. We find that D(surface)/D(bulk) is approximately equal to 0.6 at temperatures T < 32 degrees C but rises abruptly to approximately 0.8-0.9 at T > 32 degrees C. Near-surface diffusion is expected to be slower than bulk diffusion owing to hydrodynamic coupling to the wall, implying reduced hydrodynamic coupling at the higher temperatures, perhaps mediated by enhanced electrostatic repulsion above the LCST transition.  相似文献   

13.
Rutile TiO(2) particles were partly dissolved into aqueous solutions of H(2)SO(4), and the Ti(4+) ions were reprecipitated by adding NH(3) aq. Rutile-anatase coupled TiO(2) particles were prepared by heating the solid recovered after centrifugation of the suspension. The content of anatase (c(A), wt%) could be controlled arbitrarily by changing the dissolved amount of rutile. The photocatalytic activity for the gas-phase oxidation of acetaldehyde was evaluated. The first-order rate constant, k, strongly depended on both c(A) and heating temperature (T(c)), increasing with an increase in T(c) at T(c)相似文献   

14.
A series of novel biodegradable multi-block copolymers PLGA-(L-Asp-alt-diol)(x)-PLGA with pendant amino groups was synthesized by ring-opening polymerization of D,L-lactide/glycolide(D,L-LA/GA) (75/25) using poly(N-Cbz-L-Asp-alt-diol)s as macroinitiator and stannous octoate as catalyst, in which the N-Cbz-L-Asp represents N-carbobenzyloxy-L-aspartic acid and diols are ethylene glycol, triethylene glycol, PEG200, and PEG600, respectively. Their structures and properties were characterized by FTIR, (1)H NMR, DSC, GPC, and elemental analysis (EA). The contents of the L-Asp unit in the copolymers were increased from 12.9 to 79.3 mmol.g(-1) with decreasing the chain length of the diol, while the glass transition temperatures of the copolymers were decreased from 27.1 to 11.7 degrees C with increasing the chain length of the diol. Thus, the results in this study provide a way to prepare biomaterials with different L-Asp unit densities or different number of bioactive sites as well as different properties through adjusting the chain length of the diol. Synthesis of PLGA-(N-Cbz-L-Asp-alt-diol)(x)-PLGA copolymers.  相似文献   

15.
16.
Liu Y  Cui L  Guan F  Gao Y  Hedin NE  Zhu L  Fong H 《Macromolecules》2007,40(17):6283-6290
Uniform nylon 6 nanofibers with diameters around 200 nm were prepared by electrospinning. Polymorphic phase transitions and crystal orientation of nylon 6 in unconfined (i.e., as-electrospun) and a high T(g) (340 degrees C) polyimide confined nanofibers were studied. Similar to melt-spun nylon 6 fibers, electrospun nylon 6 nanofibers also exhibited predominant, meta-stable gamma crystalline form, and the gamma-crystal (chain) axes preferentially oriented parallel to the fiber axis. Upon annealing above 150 degrees C, gamma-form crystals gradually melted and recrystallized into the thermodynamically stable alpha-form crystals, which ultimately melted at 220 degrees C. Release of surface tension accompanied this melt-recrystallization process, as revealed by differential scanning calorimetry. For confined nanofibers, both the melt-recrystallization and surface tension release processes were substantially depressed; gamma-form crystals did not melt and recrystallize into alpha-form crystals until 210 degrees C, only 10 degrees C below the T(m) at 220 degrees C. After complete melting of nano-confined crystals at 240 degrees C and recrystallization at 100 degrees C, only alpha-form crystals oriented perpendicular to the nanofiber axis were obtained. In the polyimide-confined nanofibers, the Brill transition (from the monoclinic alpha-form to a high temperature monoclinic form) was observed at 180-190 degrees C, which was at least 20 degrees C higher than that in unconfined nylon 6 at approximately 160 degrees C. This, again, was attributed to the confinement effect.  相似文献   

17.
Monodisperse poly(N-isopropylacrylamide-co-acrylic acid) hydrogel microspheres were prepared by a membrane emulsification method using membranes of pore diameters of 0.33, 0.73, 1.15, and 1.70 μm. The hydrogels were synthesized by polymerization of 3.6 M N-isopropylacrylamide (N-IPAAm or NIPAM) and 0.4 M acrylic acid (AAc). Their surface properties were studied by measuring the electrophoretic mobility of the microspheres in electrolyte solutions at pH 7.4 at 25, 30, 33, 35, 40, and 45 degrees C. Poly(N-IPAAm-co-AAc) microspheres have shown negative mobility. More negative values of electrophoretic mobility were obtained with the smaller microspheres than the larger ones at each temperature. The surface charge density of the microspheres increased and their surfaces became harder above 35 degrees C, since the microspheres contained thermosensitive poly(N-IPAAm) moiety and LCST increased by the addition of AAc, while that of poly(N-IPAAm) was 33 degrees C. It has recently been found that the smaller microspheres exhibit the stronger dependence of both surface charge density and softness on the temperature. Copyright 2000 Academic Press.  相似文献   

18.
The helix-coil transition of poly[d(I-C)] and poly[d(A-T)] was studied as a function of hydrostatic pressure, temperature, and sodium ion concentration. These studies were undertaken in light of a recently published phase diagram for double stranded nucleic acids [Dubins et al. J. Am. Chem. Soc. 2001, 123, 9254-9259]. The sign and magnitude of the volume change for the heat-induced helix-coil transition, DeltaV(T), of poly[d(I-C)] and poly[d(A-T)] were dependent on the helix-coil transition temperature, T(M), at atmospheric pressure. The sign of DeltaV(T) changed from negative to positive as T(M) was increased by increasing the sodium ion concentration. For poly[d(I-C)], DeltaV(T) = 0 cm(3) mol(-1), when the sodium ion concentration is such that the spectroscopically monitored T(M) = 55 degrees C at atmospheric pressure. For poly[d(A-T)], the value of DeltaV(T) = 0 under conditions such that T(M) = 47 degrees C at atmospheric pressure. Negative values of DeltaV(T) imply that the helical form is destabilized at high pressure. Under experimental conditions where the DeltaV(T) for the transition is negative, the transition could be caused by increasing the pressure under isothermal conditions. At temperatures below T(M) measured at atmospheric pressure the midpoint of the pressure-induced helix-coil transition, P(M), decreases with increasing temperature. The volume change of the pressure-induced transitions helix-coil transition, DeltaV(P), was calculated assuming a two-state model. The magnitude of DeltaV(P) (per cooperative length) was much larger than the volume change (per base pair) measured for the heat-induced transition, DeltaV(T), calculated using the Clapeyron equation. The ratio of these two volume changes was used to calculate the cooperative length for the pressure-induced transition. This parameter depends strongly on temperature, becoming greater closer to T(M) measured at atmospheric pressure. At temperatures approaching T(M) the magnitude of the cooperative length of the pressure-induced transition is approximately twice that observed for the heat-induced transition (N(T)). On the basis of the temperature dependence of the DeltaV(T) for the two polymers the coefficient of thermal expansion of the two polymers was found to be 0.17 and 0.16 cm(3) K(-1) mol(-1) for poly[d(I-C)] and poly[d(A-T)], respectively.  相似文献   

19.
The simultaneous improvement of the melting temperature (T(m) = 224 °C) and the decomposition temperature (T(10) = 359 °C) of poly(lactide)s was achieved by the stereocomplex formation of poly(l-lactide) and poly(d-lactide) with bio-based aromatic groups at both initiating and terminating chain ends.  相似文献   

20.
A 2'-O-methyluridylic acid derivative 3 having a cyclic structure linked between the 5-position of the uracil residue and the 5'-phosphate group was synthesized. The NMR analysis suggests that this cyclouridylic acid derivative has exclusively the C3'-endo conformation that is in favor of duplex formation with RNA. Two oligonucleotides ?pc3Um(pT)(9) and pc3Um(pU)(9) incorporating this cyclouridylic acid unit at the 5'-terminal site were synthesized by using the fully protected cyclouridylic acid 3'-phosphoramidite derivative 11 in the solid-phase synthesis. To examine the actual effect of this cyclic structure on the thermal stability of duplexes between the modified oligonucleotides and their complementary oligonucleotides, two oligonucleotides ?pUm(pT)(9) and pUm(pU)(9) having an acyclic structure were also synthesized. As the complementary oligonucleotides, dA(pdA)(9) and A(pA)(9) were used for T(m) experiments with these 5'-terminal modified oligonucleotides. The T(m) values of all the possible duplexes were measured. These results clearly show that the duplex of pc3Um(pT)(9)-A(pA)(9) has a higher T(m) value by 5.5 degrees C than that of A(pA)(9)-T(pT)(9). This is rather significant compared with all other cases. Moreover, the T(m) value of pc3Um(pT)(9)-A(pA)(9) is 4.5 degrees C higher than that of pUm(pT)(9)-A(pA)(9). This result suggests that the cyclic structure can considerably contribute to stabilization of the duplex only in the case of the modified oligomer (DNA) and decaadenylate (RNA).  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