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 621 毫秒
1.
Films of poly(l-lactic acid) (PLLA) with different number-average molecular weights (Mn) and d-lactide unit contents (Xd) were made amorphous and the effects of molecular weight and small amounts of d-lactide units on the hydrolytic degradation behavior in phosphate-buffered solution at 37 °C of PLLA were investigated. The degraded films were investigated using gravimetry, gel permeation chromatography, polarimetry, differential scanning calorimetry, X-ray diffractometry, and tensile testing. To exclude the effects of crystallinity on the hydrolytic degradation, the films were made amorphous by melt-quenching. The incorporation of small amounts of d-lactide units drastically enhanced the hydrolytic degradation of PLLA. In the period of 0-32 weeks, the hydrolytic degradation rate constant (k) of PLLA films increased with increasing Xd, while the k values did not depend on Mn. This means that the effects of Xd on the hydrolytic degradation rate of the films are higher than those of Mn. In contrast, in the period of 32-60 weeks neither Xd nor Mn was a crucial parameter to determine k values, probably because in addition to these parameters the differences in the amount of catalytic oligomers accumulated in films and crystallinity affect the hydrolytic degradation behavior of the films. The initially amorphous PLLA films remained amorphous even after the hydrolytic degradation for 60 weeks.  相似文献   

2.
Speed of sound and density values for ternary systems (amino acid + salt + water): l-isoleucine/l-proline/l-glutamine in aqueous solutions of 1.5 M KCl, 1 M KNO3, and 0.5 M K2SO4 have been measured for several concentrations of amino acids at different temperatures (303.15, 308.15, 313.15, 318.15, and 323.15 K). Using speed of sound and density data, the thermodynamic parameters such as isentropic compressibility (κs), change in isentropic compressibility (Δκs) and relative change (Δκs/κ0) in isentropic compressibility have been computed. The isentropic compressibility values decrease with increase in the amino acid concentration as well as with temperature. The decrease in κs values with increase in concentration of l-isoleucine/l-proline/l-glutamine in 1.5 M KCl, 1 M KNO3, and 0.5 M K2SO4 has been ascribed to an increase in the number of incompressible zwitterions in solutions, and the formation of ‘zwitterions-ions’ and ‘zwitterions-water dipole’ entities in solutions. The decrease in κs values with increase in temperature has been attributed to the corresponding decrease of κrelax (a relaxational part of compressibility), which is dominant over the corresponding increase of κ (an instantaneous part of compressibility). The trends of variation of Δκs and Δκs/κ0 with variations in solute concentration and temperature have also been discussed in terms of solute-solute and solute-solvent intermolecular/interionic interactions operative in the systems.  相似文献   

3.
Poly(d,l-lactide) (PDLLA) degraded at processing temperature under air and nitrogen. A random chain scission model was established and used to determine the activation energy Ea, and FT-IR, 1H and 13C NMR were used to elucidate the degradation behavior under different atmospheres. Results showed that there were two to three stages. The 1st stage was dominated by the oligomers containing carboxylic acid groups and hydroxyl groups, during which oxygen and nitrogen had little effect on the degradation, thus they share similar Ea. When the oligomers were consumed over or evaporated, the 2nd stage began, and oxygen had a promoting effect on the thermo-oxidation process, resulting in the great decrease in Ea. The third stage of PDLLA was observed when it degraded under nitrogen over 200 °C, which was caused by the appearance of carboxylic acid substance.  相似文献   

4.
To control the depolymerization process of poly(l-lactic acid) into l,l-lactide for feedstock recycling, the racemization of l,l-lactide as a post-depolymerization reaction was investigated. In the absence of a catalyst, the conversion to meso-lactide increased with increase in the heating temperature and time at a higher rate than the conversion into oligomers. The resulting high composition of meso-lactide suggests that the direct racemization of l,l-lactide had occurred in addition to the known racemization mechanism that occurs on the oligomer chains. In the presence of MgO, the oligomerization rapidly proceeded to reach an equilibrium state between monomers and oligomers. The equilibrium among l,l-, meso-, and d,d-lactides was found to be a convergent composition ratio l,l-:meso-:d,d-lactides = 1:1.22:0.99 (wt/wt/wt) after 120 min at 300 °C. This composition ratio also indicates that in addition to the known racemization reaction on the oligomer chains, direct racemization among the lactides is also a frequent occurrence.  相似文献   

5.
Ultrasonic velocity and density values have been measured for ternary systems (amino acid/di-peptide + salt + water): l-leucine/l-asparagine/glycylglycine each in 1.5 M aqueous solutions of NaCl or NaNO3 or KNO3 used as solvents for several concentrations of amino acids/di-peptide at different temperatures in the range of 298.15-323.15 K. The ultrasonic velocity values have been found to increase with increase in amino acids/di-peptide concentration and temperature in all the systems. The increase in ultrasonic velocity with increase in concentration has been discussed in terms of electrostatic interactions occurring between terminal groups of zwitterions (NH4+ and COO) and Na+, K+, Cl, NO3 ions. The interactions of water dipoles with cations/anions and with zwitterions have also been taken into consideration. It has been observed that the ion-zwitterion and ion-dipole attractive forces are stronger than those of ion-hydrophobic repulsive forces. These interactions comprehensively introduce the cohesion into solutions under investigation. The cohesive forces are further enhanced on successive increases in solute concentration. Using ultrasonic velocity and density data, the parameters such as isentropic compressibility (κs), change (Δκs) and relative change (Δκs/κ0) in isentropic compressibility, specific acoustic impedance (Z) and relative association (RA) have been computed. The isentropic compressibility values decrease with increase in the concentration of solutes as well as with temperature. The decrease in κs values with increase in concentration of l-leucine, l-asparagine and glycylglycine in 1.5 M aqueous solutions of NaCl, NaNO3 and KNO3 have been explained in terms of an increase in the number of incompressible molecules/zwitterions in solutions and the formation of compact zwitterions-water dipole and zwitterions-ions structures in solutions. The decrease in κs values with increase in temperature has been attributed to the corresponding decrease of κrelax. (relaxational part of compressibility), which is dominant over the corresponding increase of κ (instantaneous part of compressibility). The trends of variations of Δκs, Δκs/κ0, Z and RA with change of concentration and temperature have also been interpreted in terms of various intermolecular/interionic interactions existing in the systems.  相似文献   

6.
7.
The predominant mechanism of the hydrolytic degradation of oligo(d,l-lactide)-grafted dextrans in phosphate buffer was followed by quantifying both released dextran and lactic acid from the copolymers. The studied amphiphilic copolymers, with well-defined structure, exhibited various oligo(d,l-lactide) weight fractions (FOLA) while having a quite high extent of free hydroxyl groups (>90%). Depending on their FOLA, oligo(d,l-lactide)-grafted dextrans were soluble either in water or in organic solvents (THF, toluene, …) and different prevailing mechanisms of hydrolytic degradation were observed. The copolymer soluble in THF, with longer oligo(d,l-lactide) grafts and higher FOLA, was found to degrade via a particular mechanism by which the greatest part of dextran was released into buffer medium during the first two weeks of degradation. During the initial stage of degradation, the hydrophilicity of dextran backbone was considered to be the main driving force for the hydrolytic cleavage of the ester linkage between backbone and grafts. Released oligo(d,l-lactide) grafts were found to be degraded via chain-end degradation or random degradation depending on their solubility in buffer medium. In case of water-soluble copolymers with shorter oligo(d,l-lactide) grafts and lower FOLA, the chain-end degradation was exclusively observed.  相似文献   

8.
“Linear” 1-arm and 2-arm poly(dl-lactide) [i.e., poly(dl-lactic acid), or PDLLA] polymers with relatively low number-average molecular weights (Mn in the range 0.2-6 × 104 g mol−1) were synthesized using ring-opening polymerization of dl-lactide initiated with tin(II) 2-ethylhexanoate (i.e., stannous octoate) and coinitiators of dl-lactic acid and ethylene glycol (these PDLLA polymers are hereafter abbreviated as 1-DL and 2-DL, respectively). Their glass-transition properties were monitored by differential scanning calorimetry, and their hydrolytic degradation was investigated using gravimetry and gel permeation chromatography. The results of the present study indicate that the coinitiator-induced molecular structural difference of the terminal groups, the chain directional change, the incorporated coinitiator moiety as an impurity in the middle of the molecule, and the molecular weight each affect both the hydrolytic degradation behavior and rate, and the glass-transition properties of the “linear” 1-DLs and 2-DLs. The glass-transition temperature (Tg) values were higher for the 2-DLs than for the 1-DLs, indicating low chain mobility and a strong inter-chain interaction of 2-arm PDLLA. However, the coinitiator-induced molecular structural difference did not produce a difference in the excess free volume of the end groups between the 1-DLs and 2-DLs, despite the difference produced in the terminal groups. On the other hand, although the hydrolytic degradation of the 1-DLs and 2-DLs proceeds via bulk erosion, significant surface erosion also occurs in the 2-DLs. This should have caused a larger weight loss and lower decrease rate of Mn of the 2-DLs compared to those of the 1-DLs. Moreover, the results of the present study indicate that in 2-arm PDLLA selective chain cleavage at the terminal ester groups or second ester groups from the chain terminals, which are induced by two terminal hydroxyl groups, is the significant hydrolytic degradation route. However, the random cleavage of ester groups, irrespective of their position, is the main hydrolytic degradation route.  相似文献   

9.
An amperometric biosensor system which uses screen-printed electrodes to simultaneously detect d-glucose and l-lactate has been developed and applied for simple and rapid determination of d-glucose and l-lactate levels in lactic fermenting beverages. The system was constructed from three-dimensionally layered electrodes. Taking into consideration the effects of easily oxidized substances contained in the samples, ferricyanide ions, which are electrochemically oxidized at a lower voltage, were chosen as a mediator. A linear relationship between steady-state current and concentration was found over a range of 1-100 mM (d-glucose) and 1-50 mM (l-lactate); the variation coefficients were 1.43% (n = 10) and 3.50% (n = 10) for the d-glucose and l-lactate sensors, respectively. When applied to lactic fermenting beverages, there was good agreement between the results obtained by the proposed sensing system and those obtained by the HPLC method. Using the proposed method, assays were completed within 5 min.  相似文献   

10.
The hydrolytic reactions between various Pt(II) complexes of the type [Pt(L)Cl2] and [Pt(L)(CBDCA-O,O′] (L is ethylenediamine, en; (±)-trans-1,2-diaminocyclohexane, dach; (±)-1,2-propylenediamine, 1,2-pn and CBDCA is the 1,1-cyclobutanedicarboxylic anion) and the N-acetylated l-methionylglycine dipeptide (MeCOMet-Gly) were studied by 1H NMR spectroscopy. All reactions were realized at 37 °C with equimolar amounts of the Pt(II) complex and the dipeptide at pH 7.40 in 50 mM phosphate buffer in D2O. Under these experimental conditions, a very slow cleavage of the Met-Gly amide bond was observed and this hydrolytic reaction proceeds through the intermediate [Pt(L)(H2O)(MeCOMet-Gly-S)]+ complex. In general, it can be concluded that faster hydrolytic cleavage of the MeCOMet-Gly dipeptide was observed in the reaction with the chloride complex than with corresponding CBDCA Pt(II) complexes. The steric effects of the Pt(II) complex on the hydrolytic cleavage of the amide bond in the MeCOMet-Gly dipeptide were also investigated by 1H NMR spectroscopy. It was found that the rate of hydrolysis decreases as the steric bulk of the CBDCA and chlorido Pt(II) complexes increase (en > 1,2-pn > dach). These results contribute to a better understanding of the toxic side effects of Pt(II) antitumor drugs and should be taken into consideration when designing new potential Pt(II) antitumor drugs with preferably low toxic side effects.  相似文献   

11.
This study explores in vitro aging effects on the surface properties of resorbable PLA95 (poly-5d/95l-lactide) bone plates. The in vitro degradation of injection molded PLA95 bone plates was undertaken by soaking them in a PBS solution. Specimens were harvested at 0, 4, 6, 8, 12, 20, and 26 weeks. After each in vitro aging period, the surface morphology, viscosity, chemical structure, wettability, and thermal properties of the PLA95 bone plates were examined by scanning electron microscopy (SEM), capillary viscometers, attenuated total reflection fourier transform infrared spectroscopy (ATR-FTIR), contact angle, and modulated differential scanning calorimetry (MDSC), respectively. The surface morphology of aged PLA95 bone plates exhibited bulk erosion. As hydrolysis progressed, the inherent viscosity (I.V.) of the PLA95 plates gradually decreased from 0.83 ± 0.01 dL/g at week 0-0.46 ± 0.03 dL/g at week 26. However, the absorbance peak intensity ratio between δas CH3 (A1452 cm−1) and νCO (A1750 cm−1) and the contact angle reveal different tendencies than that of molecular weight, which decreases. The contact angle of the PLA95 plates decreased until week 4, increased until week 8, and subsequently decreased again. Peak separation analysis reveals that the equilibrium part of the modulated DSC overlapped curves exhibit triple endothermic peaks. Over time, in vitro degradation changes the position and area of the individual peaks. After different time periods of degradation, the variation of wettability shows a tendency similar to the change of PLA95 plates crystallinity; the intensity ratio of A1452 cm−1 and A1750 cm−1as CH3CO) absorbance peaks varied like the ratio of β/α-crystal heat of fusion. Results also show a similarity in the degradation time dependence in MDSC, contact angle, and ATR-FTIR measurements. During the in vitro aging process, the breakdown and subsequent recrystallization of PLA95 molecular chains might be attributed to a progressive change in wettability and the molecular conformation between δas CH3 and νCO.  相似文献   

12.
Biodegradable poly(l-lactic acid) (PLLA) nanofiber was prepared by a carbon dioxide (CO2) laser supersonic drawing which was carried out by irradiating the laser on an as-spun fiber in a supersonic jet. The supersonic jet was generated by blowing off air into a vacuum chamber from a fiber supplying orifice. The flow velocity from the orifice can be estimated by applying Graham’s theorem from the pressure difference between the atmospheric pressure and the pressure of the vacuum chamber. The fastest flow velocity estimated was 396 m s−1 when the chamber pressure was 6 kPa. The PLLA nanofiber having an average diameter of 0.132 μm was obtained when the supersonic drawing was carried out by irradiating the laser at 177 W cm−2 on the as-spun fiber supplied at 0.1 m min−1 in the vacuum chamber at 6 kPa. The obtained nanofiber had a draw ratio of about 323,000 and a degree of crystallinity of 45%, and its diameter uniformity was high. The CO2 laser supersonic drawing was a new route for preparation of various nanofibers without using any solvent.  相似文献   

13.
Poly(l-lactic acid) (PLLA) nonwoven fabric was obtained by using a carbon dioxide laser-thinning method. The obtained PLLA nonwoven fabric was made of endless microfibers with a uniform diameter without droplets. The fiber diameter can be varied by controlling an airflow rate supplied to the air jet, a supplying speed of an original fiber into a laser-irradiating point, and laser intensity. When the microfiber prepared by irradiating the laser operated at a laser intensity of 66 W cm−2 to the original fiber supplied at Ss = 0.1 m min−1 was dragged at an airflow rate of 30 L min−1, the thinnest microfiber with an average diameter of 3.4 μm was obtained. The obtained microfiber had a degree of crystallinity of 45%, and the degree of crystal orientation of 84%. The existence of highly oriented crystallites suggests that a flow-induced crystallization occurred during the laser-thinning.  相似文献   

14.
Surface properties and enzymatic degradation of poly(l-lactide) (PLLA) end-capped with hydrophobic dodecyl and dodecanoyl groups were investigated by means of advancing contact angle (θa) measurement, quartz crystal microbalance (QCM) and atomic force microscopy (AFM). The θa values of end-capped PLLA films were larger than those of non-end-capped PLLA films, suggesting that the hydrophobic dodecyl and dodecanoyl groups were segregated on the film surface. The weight changes of end-capped PLLA thin films during enzymatic degradation in the presence of proteinase K were monitored by using a QCM technique. The relatively fast weight loss of PLLA film occurred during first few hours of degradation, followed by a decrease in the erosion rate. The erosion rate of PLLA films at the initial stage of degradation was dependent on the chain-end structure of PLLA molecules, and the value decreased with an increase in the amount of hydrophobic functional groups. The surface morphologies of PLLA thin films before and after degradation were characterized by AFM. After the enzymatic degradation, the surface of non-end-capped PLLA films was blemished homogeneously. In contrast, the end-capped PLLA thin films were degraded heterogeneously by the enzyme, and many hollows were formed on the film surface. From these results, it has been concluded that the introduction of hydrophobic functional groups at the chain-ends of PLLA molecules depressed the erosion rate at the initial stage of enzymatic degradation.  相似文献   

15.
Aluminum(III) can be absorbed when it is appropriately complexed. There are several plasma components which can bind weakly Al(III). Many proteins bind Al(III) in solution quite strongly. Carbohydrates bearing an abundance of electronegative functional groups can interact with metal cations. In solution, d-ribose exists as a mixture at equilibrium of many isomers and only a few of them bear a ‘complexing’ sequence of the hydroxyl groups. The presence of d-ribose in an Al(III) solution experiences a decrease of its Brönsted-acid sites. The lowering of the Brönsted acidity of an Al(III)-d-ribose mixture suggests the existence of attractive interactions (‘association’) between Al(III) ion and the complexing sequence of the hydroxyls of d-ribose. There is enhancement in the stability of the interaction complexes between Al(III) and d-ribose through strong intramolecular hydrogen bonding, which offers the possibility to investigate the kinetics of the subsequent proton release reactions. On the basis of the kinetic results, it may be concluded that proton release reactions, which are associated with the complexation reactions, are associatively activated. The complexes (Al(H2O)6−n(d-ribosenH)(3−n)+) resulting from the various ‘complexing’ forms of d-ribose are formed at mainly acidic pH. As the pH increases, the values of the activation enthalpy, ΔH, are changing, because of the formation of mixed hydroxo-complexes (Al(H2O)6−nm(OH)m(d-ribosenH)(3−nm)+); finally, OH displaces d-ribose from the coordination sphere of Al(III) in a rather slow process, i.e. with high values of ΔH; the activation enthalpy values, ΔH, decrease with the progression of the displacement, becoming finally very small due to the formation of a precipitate. Chelate coordination of d-ribose with some divalent and trivalent metal ions has been also reported.  相似文献   

16.
Lijuan Hua  Xueji Zhang 《Talanta》2009,77(5):1654-4893
Water-soluble CdTe quantum dots (QDs) with five sizes (2.25, 2.50, 2.77, 3.12, and 3.26 nm) were synthesized with the hydrothermal method. The electrochemiluminescence (ECL) of CdTe QDs was investigated in detail in air-saturated solution without adding foreign oxidant. It was found that the ECL of CdTe QDs displayed a size-dependent property. With the increasing in the particle size of the CdTe QDs, the ECL intensity was gradually increased, in addition, both ECL peak potentials and ECL onset potentials of CdTe QDs were shifted positively. Influences of some factors on the ECL intensity were investigated. Under the optimal conditions, the ECL intensity had a linear relationship with the concentration of l-cysteine (l-Cys) in the range from 1.3 × 10−6 to 3.5 × 10−5 mol L−1 (R2 0.996) with a detection limit of 8.7 × 10−7 mol L−1 (S/N = 3). The proposed method was applied to the determination of l-Cys in real samples with satisfactory results. Compared with previous reports, it has better selectivity for the determination of l-Cys.  相似文献   

17.
Three new alkali-metal compounds stabilized by aryloxo groups were synthesized and fully characterized. The reactions of carbon-bridged bis(phenol)s MBMPH2 (MBMPH2 = 2,2′-methylene-bis(6-tert-butyl-4-methylphenol)) with sodium and potassium metals in tetrahydrofuran (THF) gave the desired alkali-metal complexes [MBMPNa2(THF)3]2 (1) and [(MBMPK2)2(THF)5]2 (2), respectively, in high isolated yields. A similar reaction of aminophenol [HNOH] ([HNOH] = N-p-methyl-phenyl(2-hydroxy-3,5-di-tert-butyl)benzylamine) with sodium gave the monosodium salt [HNONa(THF)]2 (3). Compounds 1-3 were well characterized, including X-ray structure determination. Compound 1 has a dimeric structure, and each sodium atom is four-coordinated with four oxygen atoms to form a distorted tetrahedron. Compound 2 has a centrosymmetric tetrameric structure. The coordination environments around the four potassium atoms are different. K1 is four-coordinated, K2 is three-coordinated, K3 is five-coordinated, whereas the coordination sphere of K4 is completed by one aryloxo oxygen atom and two oxygen atoms from two THF molecules and six carbon atoms from one arene ring of the bis(phenolate) ligand. Compound 3 has a dimeric structure, and each of the sodium atoms is four-coordinated to form a distorted tetrahedron. It was found that compounds 1-3 can efficiently initiate the ring opening polymerization of l-lactide in the absence of alcohol, yielding polymers with high molecular weights for a wide range of monomer-to-initiator ratios.  相似文献   

18.
Poly(l-lactide) (PLLA) was melt-blended with poly(p-vinyl phenol) (PVPh) using a two-roll mill, and the miscibility between PLLA and PVPh and degradation of the blend films were investigated. It was found that PLLA/PVPh blend has miscibility in the amorphous state because only single Tg was observed in the DSC and DMA measurements. The Tg of the PLLA/PVPh blend could be controlled in the temperature range from 55 °C to 117 °C by changing the PVPh weight fraction. In alkaline solution, degradation rate of PLLA/PVPh blends was faster than that of neat PLLA because PVPh could dissolve in alkaline solution. The surface morphology of degraded PLLA and PLLA/PVPh blend were observed by SEM. The surface morphology of degraded PLLA/PVPh blend was finer than that of PLLA. Young's modulus of PLLA/PVPh blend increased with increasing PVPh content. Yield stress of PLLA/PVPh blends whose PVPh content was less than 30 wt% kept the level of about 55 MPa and that of PLLA/PVPh blend whose PVPh content was 40 wt% is much lower than that of neat PLLA.  相似文献   

19.
Wu-Song Li  Bing-Bing Wang  Yen Wei 《Tetrahedron》2007,63(36):8794-8800
Novel dendrons based on glycine and l-glutamic acid from the first generation (G1) to the third generation (G3) were synthesized and studied for their gelation properties by using transmission electron microscopy (TEM), atomic force microscopy (AFM), fluorescence, IR, circular dichroism (CD), and 1H NMR spectroscopy. It was found that the gelation capability of these dendrons increased from the first generation (G1) to the third generation (G3), and that G3 exhibited the highest efficiency in forming gels. Both the focal and peripheral groups of dendrons had great effects on the formation of organogels. Hydrogen bonding and π-π stacking interactions were proved to be the main driving forces to form the fibrous networks at low concentrations (0.5 wt %). Small-angle X-ray scattering (SAXS) and wide-angle X-ray diffraction (WAXD) measurements indicate that the xerogels of the second generation (G2) from ethyl acetate and ethanol, and G3 xerogel from CH2Cl2 all display lamellar structures with the interlamellar spacing of ca. 36.0 Å for G2 and 40.5 Å for G3, respectively.  相似文献   

20.
A rapid method for the identification and quantification of l-ascorbic acid in wines by direct injection liquid chromatography equipped with a UV detection was developed. The levels of ascorbic acid were determined using a polymeric PLRP-S 100 A (5 μm) column (150 mm × 4.6 mm) with a mobile water/trifluoroacetic acid (99/1, v/v) phase. The method is rapid (less than 5 min) and sensitive (LOQ of 5 mg L−1). The calibration curve of ascorbic acid was linear (r = 0.999) over a concentration range between 1 and 200 mg L−1. Repeatability was less than 2.5% and the recovery over 95%.  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