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 468 毫秒
1.
Asakura K  Hayashi M  Osanai S 《Chirality》2003,15(3):238-241
The influence of latent heat dissipated by the crystallization of 1,1'-binaphthyl in its supercooled molten state on the chiral symmetry breaking transition was investigated. Temperature change in the crystallization system was monitored by infrared thermocamera. Temperature rise due to the dissipation of latent heat in the growing front of polycrystalline aggregate was about 2 degrees C in an unstirred crystallization system. The melting point of racemic mixture and racemic compound of 1,1'-binaphthyl is 145 degrees C and 158 degrees C, respectively. The latent heat generated by the crystallization could thus change the crystallization behavior when the initial temperature of the melt was slightly lower than 145 degrees C. The temperature change in both unstirred and stirred crystallization systems was monitored. In the stirred crystallization system, even in the case when the initial temperature of the melt was about 2 degrees C lower than 145 degrees C, the temperature rose by about 4 degrees C immediately after the onset of crystallization. This indicates that the role of stirring as the critical parameter for the chiral symmetry breaking transition is not only to clone the chiral crystals but also to enhance the dissipation of latent heat due to secondary nucleation.  相似文献   

2.
Chiral symmetry breaking can be realized in stirred crystallization of Na-ClO3. We present experimental and theoretical studies of the random distribution of crystal enantiomeric excess (cee) for various stirring and solvent evaporation rates. For a fixed solvent evaporation rate, as the stirring RPM is increased, the probability distribution of cee initially broadens and subsequently develops a sharp peak close to cee = 1. On further increase of stirring rate, the probability distribution once again broadens. This broad probability distribution becomes narrow, with a sharp peak near cee = 1, if the solvent evaporation rate is decreased. Thus we show some ways in which the probability distribution of cee can be controlled in stirred crystallization. In particular, our study shows that the stirring rate and the solvent evaporation rate can be adjusted to maximize crystal enantiomeric excess. © 1995 Wiley-Liss, Inc.  相似文献   

3.
Although the theory of spontaneous chiral symmetry-breaking in open systems was proposed some time ago, experimental realization of this phenomenon has not been achieved. In this article, we note that the crystal growth front of 1,1'-binaphthyl shows many of the characteristics of an open system in which chiral symmetry-breaking has occurred. We studied the temperature profiles of the crystallizing surface and obtained X-ray diffraction data of the crystals grown from the melt under controlled conditions. The data show that, under appropriate conditions, the observed bimodal probability distribution of enantiomeric excess (ee) with maxima approximately 60% is due exclusively to chiral crystals and not due to racemic crystals of 1,1'-binaphthyl that can also form at large supercooling. The mass independence of the ee shows that the growing front maintains a constant ee, which is a clear signature of open systems in steady state. Chirality 16:131-136, 2004.  相似文献   

4.
Dissolution of sucrose crystals in the anhydrous sorbitol melt   总被引:1,自引:0,他引:1  
The dissolution of a sugar (sucrose as a model) with higher melting point was studied in a molten food polyol (sorbitol as a model) with lower melting point, both in anhydrous state. A DSC and optical examination revealed the dissolution of anhydrous sucrose crystals (mp 192 degrees C) in anhydrous sorbitol (mp 99 degrees C) liquid melt. The sucrose-sorbitol crystal mixtures at the proportions of 10, 30, 60, 100 and 150 g of sucrose per 100 g of sorbitol were heat scanned in a DSC to above melting endotherm of sorbitol but well below the onset temperature of melting of sucrose at three different temperatures 110, 130 and 150 degrees C. The heat scanning modes used were with or without isothermal holding. The dissolution of sucrose in the sorbitol liquid melt was manifested by an increase in the glass transition temperature of the melt and corresponding decrease in endothermic melting enthalpy of sucrose. At given experimental conditions, as high as 25 and 85% of sucrose dissolved in the sorbitol melt during 1 h of isothermal holding at 110 and 150 degrees C, respectively. Optical microscopic observation also clearly showed the reduction in the size of sucrose crystals in sorbitol melt during the isothermal holding at those temperatures.  相似文献   

5.
An early endothermic peak at approximately 150 degrees C was observed for crystalline sucrose by differential scanning calorimetry. The enthalpy at this temperature was found to vary with recrystallised sucrose from different sources. The addition of mineral salts to recrystallisation solutions decreased the enthalpy of the peak at around 150 degrees C, whereas the absence of salts increased it. The presence of organic solvents and polysaccharides in solution had a minor effect compared to the inorganic impurities. The peak was also depleted by increasing the amount of stirring and temperature at which recrystallisation was performed.  相似文献   

6.
Stirred crystallization of racemic (D/L)‐glutamic acid (Glu) in the presence of small amounts of L‐ or D‐lysine (Lys) was studied for the effect of transient chiral resolution by monitoring the time evolution of optical rotation and the concentration of the solution. The presence of a small amount of L‐ or D‐Lys retards the crystallization rate of the corresponding enantiomer of Glu in a chirally selective manner, giving rise to transient optical resolution of racemic Glu during crystallization. The optical rotation of the Glu solution was found to increase from zero to a value corresponding to an enantiomeric excess (ee) of 22–35% and subsequently decreases to zero over a period of many hours. During this process, the ee of the crystallized Gu is nearly 100% during the first 35 min and then it decreases slowly to zero. Our results indicate that the time at which the ee of the solution reaches its maximum and the maximum value of the ee show a nonlinear dependence on the initial mole fraction of the chiral impurity. The effect of the impurity is highly chirally selective, indicating “molecular recognition.” Chirality 11:343–348, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

7.
The influence of the synthesis conditions (method and sequence of adding reagents, reaction temperature, stirring rate) on the luminescence performance of the rare earth ternary complex compound europium-salicylic acid-(1,10-phenanthroline) [Eu(Sal)(3)Phen] was studied. We show that the method and sequence of adding reagents greatly affects the luminescence properties of the products. Within the range of synthesis temperatures (30-70 degrees C), the complex has a better luminescence performance at the lower temperature. However, performance decreased when the temperature decreased to 20 degrees C. Increase of stirring rate led to better performance in the range 200-700 r.p.m. Infrared (IR) spectroscopy, scanning electron microscopy (SEM) and wide-angle X-ray diffraction (WAXD) were utilized to characterize the structure and morphology of Eu(Sal)(3)Phen. The studies demonstrate that the crystallization of the product increased with increasing luminescence of the product. Therefore, it is proposed that increasing the crystallization of the product will enhance its luminescence performance.  相似文献   

8.
The crystallization kinetics of poly(l-lactide), PLLA, is slow enough to allow a quasi-amorphous polymer to be obtained at low temperature simply by quenching from the melt. The PLLA crystallization process was followed by differential scanning calorimetry and optical microscopy after nucleation isothermal treatments at temperatures just below (53 degrees C) and just above (73 degrees C) the glass transition temperature. The crystallization exotherm shown in the heating thermograms shifts toward lower temperatures as the annealing time at 73 degrees C increases. The same effect is shown to a lesser extent when the sample nucleates at 53 degrees C, showing the ability to nucleate in the glassy state, already shown in other polymers. The shape of the DSC thermograms is modeled by using Avrami's theory and allows an estimation of the number of crystallization germs formed. The results of optical microscopy are converted to thermograms by evaluating the average gray level of the image recorded in transmission mode as a function of temperature and calculating its temperature derivative. The shape of such optical thermograms is quite similar to that of the DSC traces but shows some peculiarities after long nucleation treatments. Atomic force microscopy was used to analyze the crystal morphology and is an additional proof of the effect of nucleation in the glassy state. The crystalline morphology observed in samples crystallized after nucleation in the glassy state is qualitatively different from that of samples nucleated above the glass transition temperature, and the number of crystals seems to be much greater than what could be expected from the crystallization kinetics.  相似文献   

9.
He C  Sun J  Zhao T  Hong Z  Zhuang X  Chen X  Jing X 《Biomacromolecules》2006,7(1):252-258
The crystallization behaviors of the poly(ethylene glycol)-poly(epsilon-caprolactone) diblock copolymer with the PEG weight fraction of 0.50 (PEG(50)-PCL(50)) was studied by DSC, WAXD, SAXS, and FTIR. A superposed melting point at 58.5 degrees C and a superposed crystallization temperature at 35.4 degrees C were obtained from the DSC profiles running at 10 degrees C/min, whereas the temperature-dependent FTIR measurements during cooling from the melt at 0.2 degrees C/min showed that the PCL crystals formed starting at 48 degrees C while the PEG crystals started at 45 degrees C. The PEG and PCL blocks of the copolymer crystallized separately and formed alternating lamella regions according to the WAXD and SAXS results. The crystal growth of the diblock copolymer was observed by polarized optical microscope (POM). An interesting morphology of the concentric spherulites developed through a unique crystallization behavior. The concentric spherulites were analyzed by in situ microbeam FTIR, and it was determined that the morphologies of the inner and outer portions were mainly determined by the PCL and PEG spherulites, respectively. However, the compositions of the inner and outer portions were equal in the analysis by microbeam FTIR.  相似文献   

10.
This experimental work was carried out to validate the use of a -152 degrees C ultra-low temperature freezer to freeze and store canine semen. The semen of three dogs was pooled and processed to obtain a final dilution with a concentration of 100 x 10(6) spermatozoa/mL, glycerol at 5% and Equex at 0.5%. Then, four freezing protocols were tested to evaluate the cryosurvival of sperm at 1, 7, 30, 60 and 120 days after freezing: (I) semen was frozen and stored in liquid nitrogen; (II) semen was frozen in liquid nitrogen and stored in the ultra-low freezer at -152 degrees C; (III) semen was frozen in the vapour of liquid nitrogen and stored in the ultra-low freezer at -152 degrees C; (IV) semen was frozen and stored in the ultra-low freezer at -152 degrees C. Data were statistically analyzed by repeated measures analysis of variance to determine the effect of the freezing protocol and time on the sperm characteristics assessed. The percentages of sperm motility and of dead/live spermatozoa were similar throughout the experimental period, with no significant differences (P < 0.05) to be observed between four different freezing techniques tested. At 120 days after freezing, the percentage of abnormal cells and the percentage of sperm cells with abnormal acrosome were not significantly different between the freezing techniques. Although the number of dogs used was slightly low, in vitro results of this preliminary study showed that the use of ultra-freezers at -152 degrees C to freeze and store canine semen could be a viable alternative to liquid nitrogen.  相似文献   

11.
Optically active 1-phenyl 1-propanol is used as a chiral building block and synthetic intermediate in the pharmaceutical industries. In this study, the enantioselective production of 1-phenyl 1-propanol was investigated systematically using response surface methodology (RSM). Before RSM was applied, the effects of the enzyme source, the type of acyl donor, and the type of solvent on the kinetic resolution of 1-phenyl 1-propanol were studied. The best results were obtained with Candida antartica lipase (commercially available as Novozym 435), vinyl laurate as the acyl donor, and isooctane as the solvent. In the RSM, substrate concentration, molar ratio of acyl donor to the substrate, amount of enzyme, temperature, and stirring rate were chosen as independent variables. The predicted optimum conditions for a higher enantiomeric excess (ee) were as follows: substrate concentration, 233 mM; molar ratio of acyl donor to substrate, 1.5; enzyme amount, 116 mg; temperature, 47 °C; and stirring rate, 161 rpm. A verification experiment conducted at these optimized conditions for maximum ee yielded 91% for 3 hr, which is higher than the predicted value of 83%. The effect of microwave on the ee was also investigated and ee reached 87% at only 5 min.  相似文献   

12.
The yeast strain CGMCC 0573 was identified as Citeromyces matriensis and shown to be capable of enantioselectively hydrolyzing ethyl ester of ( R)-Ketoprofen (2-(3-benzoylphenyl)propionic acid). The strain was isolated for the first time from soil samples through a new and efficient screening procedure in which the probability of obtaining active strains was greatly increased by using ethanol and Tween-80 alternatively as additives during the enrichment culture. Studies of the culture conditions and catalytic performance of Citeromyces matriensis CGMCC 0573 showed that the enzyme occurs constitutively in the cells and its production is enhanced by feeding with Tween-80 during the early period of cultivation. Yeast extract was found to be beneficial both for growth and for esterase production. The optimal temperature and pH for the bioconversion were 40 degrees C and pH 8.0, respectively. Biotransformation using resting cells cultured in a flask with baffles and magnetic stirring and in the presence of 50 mM substrate resulted in the production of ( R)-ketoprofen at 93% ee (enantiomeric excess) and at 42.6% conversion.  相似文献   

13.
Taira S  Crépy KV  Imamoto T 《Chirality》2002,14(5):386-392
The copper-catalyzed enantioselective conjugate addition of diethylzinc to 2-cyclohexen-1-one was investigated using (R,R)-bis-(t-butylmethylphosphino)methane (1c) as a chiral ligand. The reaction was carried out at 0 degrees C in THF-toluene as the solvent system and in the presence of 1.2 mol% of CuOTf afforded (S)-3-ethylcyclohexan-1-one with 85% ee.  相似文献   

14.
Aminolytic kinetic resolution (AKR) of trans-stilbene oxide and trans-beta-methyl styrene oxide proceeded smoothly under microwave irradiation using chiral dimeric and polymeric Cr(III) salen complexes as efficient catalysts, giving regio-, diastereo-, and enantioselective anti-beta-amino alcohols in high yields (49%) and chiral purity (ee up to 94%) in case of 4-methylaniline within 2 min. The kinetic resolution system is approximately five times faster than traditional oil bath heating at 70 degrees C and 420 times faster than the reaction conducted at room temperature with concomitant recovery of respective chirally enriched epoxides (ee, 92%) in excellent yields (up to 48%). The catalyst 1 worked well in terms of enantioselectivity than the catalyst 2, but both the catalysts were easily recovered and reused five times with the retention of its efficiency.  相似文献   

15.
Zhao Y  Yang XB  Wang QF  Nan PJ  Jin Y  Zhang SY 《Chirality》2007,19(5):380-385
A new capillary zone electrophoresis (CZE) method was developed to separate three chiral 2,3-dihydroxy-3-phenylpropionate enantiomers using neutral hydroxypropyl-beta-CD (HP-beta-CD) as chiral selector and borate as background electrolyte. The results showed that HP-beta-CD exhibited good enantioselectivity and high resolution was achieved under the optimum condition of pH 10.3, 200 mM borate buffer containing 6% methanol and 50 mM HP-beta-CD at 15 kV and 20 degrees C within 16 min. The precision of the method was <0.9% for migration time and 4.5% for corrected peak area. In addition, the developed method was successfully applied to the determination of enantiomeric excess (ee) of synthetic 2,3-dihydroxy-3-phenylpropionate samples. With this method, low as 0.2% impurity of the undesirable enantiomer in the presence of high amount of target enantiomer was determined. The results demonstrated that the proposed CZE method is a simple and useful technique and is applicable to ee assay of 2,3-dihydroxy-3-phenylpropionate enantiomers.  相似文献   

16.
The aim of this work was to survey how processing parameters affect the crystal growth of beta-sitosterol in suspension. The process variables studied were the cooling temperature, stirring time and stirring rate during recrystallization. In addition, we investigated the effect a commonly used surfactant, polysorbate 80, has on crystal size distribution and the polymorphic form. This study describes the optimization of the crystallization process, with the object of preparing crystals as small as possible. Particle size distribution and habit were analyzed using optical microscopy, and the crystal structure was analyzed using X-ray diffractometry. The cooling temperature had a remarkable influence on the crystal size. Crystals with a median crystal length of approximately 23 microm were achieved with a low cooling temperature (<10 degrees C); however, a fairly large number of crystals over 50 microm appeared. Higher cooling temperatures (>30 degrees C) caused notable crystal growth both in length and width. Rapid (250 rpm), continuous stirring until the suspensions had cooled to room temperature created small, less than 50 micro m long (median <20 microm), needle-shaped crystals. The addition of surfactant slightly reduced the size of the initially large crystals. Both hemihydrate and monohydrate crystal forms occurred throughout, regardless of the processing parameters. By using an optimized process, it was possible to obtain a microcrystalline suspension, with a smooth texture.  相似文献   

17.
The crystallization behavior and crystalline morphologies of poly[(S)-lactide] (P[(S)-LA]) in thin films crystallized isothermally at over 160 degrees C were characterized by transmission electron microscopy and atomic force microscopy (AFM). The dendritic crystal and hexagonal crystal were formed in thin film with thicknesses below 30 nm or over 50 nm, respectively. The crystal structures of dendritic and hexagonal crystals were identical, suggesting that the crystalline morphology of P[(S)-LA] is strongly dependent upon the film thickness. In situ observation of the crystal growth in the P[(S)-LA] thin film at 165 degrees C from the melt was carried out by using temperature-controlled AFM equipped with a heating stage. The initial stage of crystallization and development of lamellae were successfully observed during isothermal crystallization at 165 degrees C. The first forming crystal showed the edge-on orientation, and grew to S-shaped edge-on lamellae. Dendritic flat-on crystals were developed from the S-shaped edge-on lamellae. The growth rates of flat-on and edge-on lamellae were almost identical.  相似文献   

18.
Shi M  Liu YH  Chen LH 《Chirality》2007,19(2):124-128
In the Morita-Baylis-Hillman (MBH) reactions of arylaldehydes with methyl vinyl ketone, it was observed that in the presence of a catalytic amount of a chiral phosphine Lewis base (CPLB) bearing multiple phenol groups, such as CPLB1 (10 mol %), the corresponding MBH adducts could be obtained in moderate to good yields with low to moderate ee's (4-45% ee) at ambient temperature (10 degrees C) in THF.  相似文献   

19.
Poly(L-lactide) [i.e., poly(L-lactic acid) (PLLA)] was hydrolyzed in the melt in high-temperature and high-pressure water at the temperature range of 180-350 degrees C for a period of 30 min, and formation, racemization, and decomposition of lactic acids and molecular weight change of PLLA were investigated. The highest maximum yield of l-lactic acid, ca. 90%, was attained at 250 degrees C in the hydrolysis periods of 10-20 min. Too-high hydrolysis temperatures such as 350 degrees C induce the dramatic racemization and decomposition of formed lactic acids, resulting in decreased maximum yield of L-lactic acid. The hydrolysis of PLLA proceeds homogeneously and randomly via a bulk erosion mechanism. The molecular weight of PLLA decreased exponentially without formation of low-molecular-weight specific peaks originating from crystalline residues. The activation energy for the hydrolysis (deltaE(h)) of PLLA in the melt (180-250 degrees C) was 12.2 kcal x mol(-1), which is lower than 20.0 kcal x mol(-1) for PLLA and 19.9 kcal x mol(-1) for poly(dl-lactide) [i.e., poly(DL-lactic acid)] as a solid in the temperature range below the glass-transition temperature (21-45 degrees C). This study reveals that hydrolysis of PLLA in the melt is an effective and simple method to obtain l-lactic acid and to prepare PLLA having different molecular weights without containing the specific low-molecular-weight chains, because of the removal of the effect caused by crystalline residues.  相似文献   

20.
Study on the metastable zone width of ketoprofen   总被引:1,自引:0,他引:1  
Lu YH  Ching CB 《Chirality》2006,18(4):239-244
With increasing awareness for the need of pure enantiomer drugs, strong emphasis has been focused on the research of chiral drug separation. Compared with other separation methods, crystallization is a simple and economical method, and the metastable zone width (MSZW) is a very important factor for the entire crystallization process. In this paper, the effects of the metastable zones of (R,S)- and (S)-ketoprofen and a 0.94 mole fraction of (S)-ketoprofen in order to enhance the MSZW were studied. Four main factors were studied, namely, temperature, cooling rate, stirring rate, and volume ratio of mixed solvent (water/ethanol). Through the L9 fractional experiment design, it was observed that all samples' MSZWs would increase with an increase in cooling rate and decrease with an increase in the ethanol volume ratio and temperature. The ethanol ratio may have the strongest effect on the process and can greatly enhance the metastable zone, and the other three factors influence the MSZW only slightly. In conclusion, the these four factors for enhancing MSZW have been optimized: water-to-ethanol volume ratio, 1:0.6; temperature, 20 degrees C; stirring rate, 700 rpm; and cooling rate, 12.0 degrees C/h. All of these results will be helpful for the following chiral separation of ketoprofen by crystallization.  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