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.
As an example of acyclic P‐chiral phosphine oxides, the resolution of ethyl‐(2‐methylphenyl)‐phenylphosphine oxide was elaborated with TADDOL derivatives, or with calcium salts of the tartaric acid derivatives. Besides the study on the resolving agents, several purification methods were developed in order to prepare enantiopure ethyl‐(2‐methylphenyl)‐phenylphosphine oxide. It was found that the title phosphine oxide is a racemic crystal‐forming compound, and the recrystallization of the enantiomeric mixtures could be used for the preparation of pure enantiomers. According to our best method, the (R)‐ethyl‐(2‐methylphenyl)‐phenylphosphine oxide could be obtained with an enantiomeric excess of 99% and in a yield of 47%. Complete racemization of the enantiomerically enriched phosphine oxide could be accomplished via the formation of a chlorophosphonium salt. Characterization of the crystal structures of the enantiopure phosphine oxide was complemented with that of the diastereomeric intermediate. X‐ray analysis revealed the main nonbonding interactions responsible for enantiomeric recognition.  相似文献   

2.
An improved method, which is highly reproducible, was developed for the enantioseparation of racemic O‐ethyl phenylphosphonothioic acid ( 1a ) with brucine by introducing seeding to a supersaturated solution of the diastereomeric salt mixture. The present method gave both diastereomeric salts in high yields with a diastereomeric ratio of >99.5:0.5 upon choosing the crystallization solvent (MeOH for the ( (R)-1a salt and MeOH/H2O for the ( (S)-1a salt). The enantiopure acid (R)-1a , (S)-1a showed a good chirality recognition ability for not only strong bases, such as amines and amino alcohols, but also weakly basic alcohols and was applicable as a solvating agent to the 1H NMR determination of the enantiomeric excess of chiral amines, amino alcohols, and alcohols, including aliphatic substrates. Chirality 26:614–619, 2014. © 2014 Wiley Periodicals, Inc.  相似文献   

3.
C2‐symmetric N,N‐bis(phosphinomethyl)amines were prepared by the thermal reaction of aromatic aldehydes with ammonia and hypophosphorus acid as previously described. Both enantiomers of C2‐symmetric N,N‐bis(phosphinomethyl)amine were obtained in a high enantiomeric purity through the diastereomeric salt formation with (–)‐quinine, and subsequent fractional crystallization. X‐ray crystallographic analysis of one of the diastereomeric salts clearly revealed that (–)‐quinine could be an efficient resolving agent for obtaining the single enantiomer (R,R)‐N,N‐bis(phosphinomethyl)amine. Chirality 27:71–74, 2015. © 2014 Wiley Periodicals, Inc.  相似文献   

4.
Protein‐coated microcrystals can be generated by co‐precipitation of protein and a water‐soluble crystalline carrier by addition to excess water miscible organic solvent. We have investigated this novel process for its utility in the concentration and partial purification of a recombinant protein exported into the culture broth during expression by Pichia pastoris. Co‐precipitation with a L ‐glutamine carrier selectively isolated the protein content of the culture broth, with a minimal number of steps, and simultaneously removed contaminants including a novel yeast metabolite. This pigment co‐elutes during aqueous chromatography but its elucidation as a benzoylated glycosamine suggested a simple route of removal by partition during the co‐precipitation process. Scale‐up of the process was readily achieved through in‐line mixing and subsequent reconstitution of the dried protein‐coated microcrystals yielded natively folded, bioactive protein. Additional washing of the crystals with saturated L ‐glutamine facilitated further purification of the recombinant protein immobilized on the L ‐glutamine carrier. Thus, we present a novel method for the harvesting of recombinant protein from culture broth as a dry powder, which may be of general applicability to bioprocessing. Biotechnol. Bioeng. 2010;106: 764–773. © 2010 Wiley Periodicals, Inc.  相似文献   

5.
The specialized salt glands on the epidermis of halophytic plants secrete excess salts from tissues by a mechanism that is poorly understood. We examined the salt glands as putative salt and water bi‐regulatory units that can respond swiftly to altering environmental cues. The tropical mangrove tree species (Avicennia officinalis) is able to grow under fluctuating salinities (0.7–50.0 dS m?1) at intertidal zones, and its salt glands offer an excellent platform to investigate their dynamic responses under rapidly changing salinities. Utilizing a novel epidermal peel system, secretion profiles of hundreds of individual salt glands examined revealed that these glands could secrete when exposed to varying salinities. Notably, rhythmic fluctuations observed in secretion rates were reversibly inhibited by water channel (aquaporin) blocker, and two aquaporin genes (PIP and TIP) preferentially expressed in the salt gland cells were rapidly induced in response to increasing salt concentration. We propose that aquaporins are involved and contribute to the re‐absorption of water during salt removal in Avicennia officinalis salt glands. This constitutes an adaptive feature that contributes to salt balance of trees growing in saline environments where freshwater availability is limited.  相似文献   

6.
A new method for oxidative folding of synthetic polypeptides assembled by stepwise solid phase synthesis is introduced. Folding is obtained in excellent yields by reacting Stert‐butylthiolated polypeptides with a 100‐fold molar excess of cysteine at 37 °C in a slightly alkaline buffer containing chaotropic salts, and in the presence of air‐oxygen. This novel protocol has been applied to the folding of Stert‐butylthiolated human thymus and activation‐regulated chemokine (hu‐TARC) derivatives as well as to larger segments of Plasmodium falciparum and Plasmodium berghei circumsporozoite proteins. Folded P. falciparum polypeptides have been used as substrates of endoproteinase Glu‐C (Glu‐C) and endoproteinase Asp‐N (Asp‐N) in an attempt to identify their disulfide connectivities. Particular practical advantages of the present method are (i) easy purification and storage of the S‐protected peptide derivatives, (ii) elimination of the risk of cysteine alkylation during the acidolytic cleavage deprotection and resin cleavage steps, (iii) possibility to precisely evaluate the extent of folding and disulfide bond formation by mass spectrometry, and (iv) facile recovery of the final folded product. Copyright © 2008 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

7.
The bulky‐headed oxidant hexadecyltrimethylammonium periodate affords the diastereomeric pairs, (Ss)‐(+)/(Rs)‐(+) and (Ss)‐(?)/(Rs)‐(?)‐neomenthyl phenyl sulfoxides in stereochemically pure states with improved diastereomeric excess (48% diastereomeric excess [de]) as compared to its nonbulky counterpart, sodium metaperiodate (28% de) from respective (+)/(?)‐neomenthyl phenyl sulfides. Steric effects involving the head group volume of hexadecyltrimethylammonium periodate is found to play a role in improving the diastereomeric ratio of the products. The two diastereomers can be readily separated by column chromatography. Absolute configuration at the sulfur center in (+)‐neomenthyl phenyl sulfoxide was determined by single‐crystal X‐ray crystallography and found to be Ss. Relative configurations of other sulfoxides were assigned based on the configuration of (+)‐neomenthyl phenyl sulfoxide. Chirality 27:370–374, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

8.
A simple, efficient, and economical method based on the combination of the exceptional behavior of o,o′‐dibenzoyl‐ or o,o′‐di‐p‐toluyl‐(2R,3R)‐tartaric acid in chiral recognition processes, and the coordination ability of calcium or magnesium ion was developed for the resolution of phospholene oxides 1 . The calcium or magnesium salt of (?)‐o,o′‐dibenzoyl‐(2R,3R)‐tartaric acid 2 , 4 ‐ 6 or calcium hydrogen (?)‐o,o′‐di‐p‐toluyl‐(2R,3R)‐tartrate 3 may form crystalline diastereomeric coordination complexes with the appropriate antipode of substituted 3‐phospholene oxides 1 that makes possible efficient resolutions. Optically active phospholene oxides 1 were prepared directly by simply crystallization and digestion of the corresponding diastereomeric complexes so formed. Thermal behavior of the crystalline diastereomeric complexes was studied by simultaneous TG/DTA. The novel method may be of more general value in respect of the resolution of tertiary phosphine oxides. Chirality, 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

9.
The excellent optoelectronic properties demonstrated by hybrid organic/inorganic metal halide perovskites are all predicated on precisely controlling the exact nucleation and crystallization dynamics that occur during film formation. In general, high‐performance thin films are obtained by a method commonly called solvent engineering (or antisolvent quench) processing. The solvent engineering method removes excess solvent, but importantly leaves behind solvent that forms chemical adducts with the lead‐halide precursor salts. These adduct‐based precursor phases control nucleation and the growth of the polycrystalline domains. There has not yet been a comprehensive study comparing the various antisolvents used in different perovskite compositions containing cesium. In addition, there have been no reports of solvent engineering for high efficiency in all‐inorganic perovskites such as CsPbI3. In this work, inorganic perovskite composition CsPbI3 is specifically targeted and unique adducts formed between CsI and precursor solvents and antisolvents are found that have not been observed for other A‐site cation salts. These CsI adducts control nucleation more so than the PbI2–dimethyl sulfoxide (DMSO) adduct and demonstrate how the A‐site plays a significant role in crystallization. The use of methyl acetate (MeOAc) in this solvent engineering approach dictates crystallization through the formation of a CsI–MeOAc adduct and results in solar cells with a power conversion efficiency of 14.4%.  相似文献   

10.
Numbers of resolving factors were investigated to improve resolution of venlafaxine 1 . An effective resolving agent, O,O′‐di‐p‐toluoyl‐(R, R)‐tartaric acid 2 , was screened using similar method of ‘Dutch resolution’ from tartaric acid derivatives. The resolution efficiency was up to 88.4%, when the ratio of rac‐ 1 and 2 was 1:0.8 in THF with little water (10:1 v/v). Enantiomerically pure venlafaxine was prepared with 99.1% ee in 82.2% yield. The chiral resolution mechanism was first explained through X‐ray crystallographic study. One diastereomeric salt with well solubility forms a columnar supramolecular structure as the acidic salt (R)‐ 1 · 2 , while the other diastereomeric salt with less solubility forms a multilayered sandwich supramolecular structure by enantio‐differentiation self‐assembly as the neutral salt 2(S)‐ 1 · 2 . The water molecules play a key role in the optical resolution, as indicated by the special structures of the diastereomeric salts.  相似文献   

11.
In this study, we describe a new approach for the characterization of process‐related impurities along with an in silico tool to generate orthogonal, integrated downstream purification processes for biological products. A one‐time characterization of process‐related impurities from product expression in Pichia pastoris was first carried out using linear salt and pH gradients on a library of multimodal, salt‐tolerant, and hydrophobic charge induction chromatographic resins. The Reversed‐phase ultra‐performance liquid chromatography (UPLC) analysis of the fractions from these gradients was then used to generate large data sets of impurity profiles. A retention database of the biological product was also generated using the same linear salt and pH gradients on these resins, without fraction collection. The resulting two data sets were then analyzed using an in silico tool, which incorporated integrated manufacturing constraints to generate and rank potential three‐step purification sequences based on their predicted purification performance as well as whole‐process “orthogonality” for impurity removal. Highly ranked sequences were further examined to identify templates for process development. The efficacy of this approach was successfully demonstrated for the rapid development of robust integrated processes for human growth hormone and granulocyte‐colony stimulating factor.  相似文献   

12.
Because of safety concerns and product consistency issues with the use of animal‐derived collagen, several recombinant protein expression hosts have been considered for recombinant collagen corn seed. Full length, triple‐helical, recombinant collagen (rCIα1) is expressed as a fusion with a foldon domain, which must later be removed. Here we have examined integration of purification and foldon removal by comparing advantages of removal before or after purification, using salt precipitation as the main purification step. Because expression levels in available maize lines are low, Pichia‐produced recombinant collagens, both with and without foldon, were added to corn seed germ at the extraction step. Salt precipitation of an acidic corn seed extract yielded 100% of the collagen without foldon at >70% purity without the pepsin pretreatment. With pepsin pretreatment, yield was 94.0% with purity of 76.5%. Analysis of the protein molecular weight distribution of the pre‐ and post‐treatment extracts showed that the corn proteins are largely resistant to pepsin proteolysis, explaining why little benefit was obtained by pepsin treatment. In the absence of pepsin treatment, the recovery of rCIα1 with foldon was still above 90% but the purity was only 44%. This still represented at about 13‐fold purification with a 2.7‐fold volume reduction which would reduce the pepsin requirement for post‐recovery foldon cleavage. © 2015 American Institute of Chemical Engineers Biotechnol. Prog., 32:98–107, 2016  相似文献   

13.
R/S mixture of monoterpene alcohol cis‐verbenol can be separated in preparative scale by its conversion into phthalic mono‐ester and subsequent crystallization of its diastereomeric salts with (R)‐α‐methylbenzylamine and (S)‐α‐methylbenzylamine. Finally, basic methanolysis of the resolved phthalic mono‐esters results (S)‐cis‐verbenol and (R)‐cis‐verbenol in high enantiomeric and diastereomeric purity.  相似文献   

14.
The micronization of ampicillin via supercritical gas antisolvent (GAS) process was studied. The particle size distribution was significantly controlled with effective GAS variables such as initial solute concentration, temperature, pressure, and antisolvent addition rate. The effect of each variable in three levels was investigated. The precipitated particles were analyzed with scanning electron microscopy (SEM) and Zetasizer Nano ZS. The results indicated that decreasing the temperature and initial solute concentration while increasing the antisolvent rate and pressure led to a decrease in ampicillin particle size. The mean particle size of ampicillin was obtained in the range of 220–430 nm by varying the GAS effective variables. The purity of GAS-synthesized ampicillin nanoparticles was analyzed in contrast to unprocessed ampicillin by FTIR and HPLC. The results indicated that the structure of the ampicillin nanoparticles remained unchanged during the GAS process.KEY WORDS: ampicillin, nanoparticles, precipitation, supercritical gas antisolvent  相似文献   

15.
The salt and stereoselective cocrystal phenomena associated with 2‐phenylglycinol and 2‐phenylglycine have been studied using X‐ray powder diffraction and differential scanning calorimetry. The chiral identities of the free acids and their sodium salts, or the free bases and their chloride salts, were found to play a determining role as to whether a salt–cocrystal product could or could not be formed. In particular, when cocrystallization of an enantiomerically pure basic or zwitterionic substance with its enantiomerically pure acid addition salt was attempted, a salt–cocrystal was only obtained when the absolute configuration of the two reactants is opposite. On the other hand, it has been found that no stereoselectivity in salt–cocrystal formation existed in the cocrystallization of an enantiomerically pure acidic or zwitterionic substance with its enantiomerically pure base addition salt. Chirality 25:8–15, 2013. © 2012 Wiley Periodicals, Inc.  相似文献   

16.
With the expansion of saline land worldwide, it is essential to establish a model halophyte to study the salt‐tolerance mechanism. The salt glands in the epidermis of Limonium bicolor (a recretohalophyte) play a pivotal role in salt tolerance by secreting excess salts from tissues. Despite the importance of salt secretion, nothing is known about the molecular mechanisms of salt gland development. In this study, we applied RNA sequencing to profile early leaf development using five distinct developmental stages, which were quantified by successive collections of the first true leaves of L. bicolor with precise spatial and temporal resolution. Specific gene expression patterns were identified for each developmental stage. In particular, we found that genes controlling salt gland differentiation in L. bicolor may evolve in a trichome formation, which was also confirmed by mutants with increased salt gland densities. Genes involved in the special ultrastructure of salt glands were also elucidated. Twenty‐six genes were proposed to participate in salt gland differentiation. Our dataset sheds light on the molecular processes underpinning salt gland development and thus represents a first step towards the bioengineering of active salt‐secretion capacity in crops.  相似文献   

17.
With an increased number of applications in the field of the avidin‐biotin technology, the resulting demand for highly‐purified protein avidin has drawn our attention to the purification process of avidin that naturally occurs in chicken egg white. The high‐throughput process development (HTPD) methodology was exploited, in order to evaluate purification process alternatives to commonly used ion‐exchange chromatography. In a high‐throughput format, process parameters for aqueous two‐phase extraction, selective precipitation with salts and polyethylene glycol, and hydrophobic interaction and mixed‐mode column chromatography experiments were performed. The HTPD strategy was complemented by a high‐throughput tandem high‐performance liquid chromatography assay for protein quantification. Suitable conditions for the separation of avidin from the major impurities ovalbumin, ovomucoid, ovotransferrin, and lysozyme were identified in the screening experiments. By combination of polyethylene glycol precipitation with subsequent resolubilization and separation in a polyethylene glycol/sulfate/sodium chloride two‐phase system an avidin purity of 77% was obtained with a yield >90% while at the same time achieving a significant reduction of the process volume. The two‐phase extraction and precipitation results were largely confirmed in larger scale with scale‐up factors of 230 and 133, respectively. Seamless processing of the avidin enriched bottom phase was found feasible by using mixed‐mode chromatography. By gradient elution a final avidin purity of at least 97% and yield >90% was obtained in the elution pool. The presented identification of a new and beneficial alternative for the purification of the high value protein thus represents a successful implementation of HTPD for an industrially relevant purification task. © 2015 American Institute of Chemical Engineers Biotechnol. Prog., 31:957–973, 2015  相似文献   

18.
The optical resolution of (R,S)‐propranolol by the diastereomeric crystallization method was successfully performed using dehydroabietic acid (DHAA) as the resolving agent in methanol. The three important parameters: DHAA amount, solvent (methanol) amount, and crystallization temperature of diastereomeric salts were optimized employing the response surface methodology (RSM). When maintaining a lower limit of 95% for the purity of (S)‐propranolol, the optimal resolution conditions were a DHAA/(R,S)‐propranolol molar ratio of 1.1, solvent/(R,S)‐propranolol ratio of 16.2 mL.g‐1, and crystallization temperature of –5 °C. The desired (S)‐propranolol was prepared with 94.8% optical purity and 72.2% yield under the optimal conditions. Chirality 27:131–136, 2015. © 2014 Wiley Periodicals, Inc.  相似文献   

19.
X‐ray crystallography is the most powerful method for determining three‐dimensional structures of proteins to (near‐)atomic resolution, but protein crystallization is a poorly explained and often intractable phenomenon. Differential Scanning Calorimetry was used to measure the thermodynamic parameters (ΔG, ΔH, ΔS) of temperature‐driven unfolding of two globular proteins, lysozyme, and ribonuclease A, in various salt solutions. The mixtures were categorized into those that were conducive to crystallization of the protein and those that were not. It was found that even fairly low salt concentrations had very large effects on thermodynamic parameters. High concentrations of salts conducive to crystallization stabilized the native folded forms of proteins, whereas high concentrations of salts that did not crystallize them tended to destabilize them. Considering the ΔH and TΔS contributions to the ΔG of unfolding separately, high concentrations of crystallizing salts were found to enthalpically stabilize and entropically destabilize the protein, and vice‐versa for the noncrystallizing salts. These observations suggest an explanation, in terms of protein stability and entropy of hydration, of why some salts are good crystallization agents for a given protein and others are not. This in turn provides theoretical insight into the process of protein crystallization, suggesting ways of predicting and controlling it. © 2016 Wiley Periodicals, Inc. Biopolymers 105: 642–652, 2016.  相似文献   

20.
Hydrophobic interaction chromatography (HIC) is one of the most frequently used purification methods in downstream processing of biopharmaceuticals. During HIC, salts are the governing additives contributing to binding strength, binding capacity, and protein solubility in the liquid phase. A relatively recent approach to increase the dynamic binding capacity (DBC) of HIC adsorbers is the use of salt mixtures. By mixing chaotropic with kosmotropic salts, the DBC can strongly be influenced. For salt mixtures with a higher proportion of chaotropic than kosmotropic salt, higher DBCs were achieved compared with single salt approaches. By measuring the surface tensions of the protein salt solutions, the cavity theory—proposed by Melander and Horváth—that higher surface tensions lead to higher DBCs, was found to be invalid for salt mixtures. Aggregation temperatures of lysozyme in the salt mixtures, as a degree of hydrophobic forces, were correlated to the DBCs. Measuring the aggregation temperatures has proven to be a fast analytical methodology to estimate the hydrophobic interactions and thus can be used as a measure for an increase or decrease in the DBCs. © 2016 American Institute of Chemical Engineers Biotechnol. Prog., 32:346–354, 2016  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