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 437 毫秒
1.
通过1-(3-二甲氨基丙基)-3-乙基碳二亚胺(EDC)与N-羟基琥珀酰亚胺(NHS)活化反应将多肽偶联到金纳米粒子表面,采用荧光光谱研究其形成的酰胺键的反应偶联效率.考察了实验条件,包括缓冲液的种类(HEPES、Tris-HCl、硼酸、PBS缓冲液)、pH值(pH 6.5~9.0)、缓冲液浓度(10, 25, 40和50 mmol/L)、NHS和EDC的浓度(NHS 0.2~1.0 mol/L,EDC 0.01~0.5 mol/L)及二者比例(0,0.5,1.0,2.0和2.5)、偶联反应时间(4, 8, 12, 24和36 h)等对偶联效率的影响,筛选出最佳实验条件.实验结果表明, 25 mmol/L 4-羟乙基哌嗪乙磺酸(HEPES)缓冲液(pH 7.0), NHS/EDC浓度为0.4 mol/L/0.2 mol/L和24 h的反应时间为EDC-NHS活化反应将多肽偶联到金纳米粒子上的最佳实验条件.本研究结果可为相关研究提供技术参考.  相似文献   

2.
用荧光标记技术对含羧基类中药成分进行标记,探索中药成分荧光标记方法,提高其检测灵敏度、为中药药代动力学研究奠定基础。以荧光试剂8-氨基芘-1,3,6-三磺酸三钠盐(APTS)和绿原酸组成研究体系,对标记条件进行优化。反应体系中加入1-乙基-3-(3-二甲氨丙基)碳二亚胺盐酸盐(EDC.HCl)与N-羟基丁二酰亚胺(NHS)可使衍生反应在室温下进行。APTS标记绿原酸的较优反应条件为:绿原酸与APTS的浓度比为1∶5,EDC的浓度为5 mmol/L,NHS的浓度为0.33 mmol/L。以0.1mol/L、pH5.50的NaH2PO4-Na2HPO4为缓冲溶液,绿原酸首先与EDC、NHS避光反应4 h,再与APTS反应4h,即可达到较好的标记效果。此标记方法操作简便,灵敏度高,可以用于含羧基类中药成分的荧光标记研究。  相似文献   

3.
在磷酸盐缓冲溶液中用1-(3-二甲氨基丙基)-3-乙基碳二亚胺(EDC)与N-羟基琥珀酰亚胺(NHS)活化羧甲基纤维素钠(CMC)侧链上的羧基; 在室温下再将活化的CMC与5'端经氨基修饰的单链脱氧核糖核酸(DNA)齐聚物(ODNs)反应, 获得CMC上接枝ODNs的共聚物(CMC-g-ODNs), 以Lambda DNA为模板, 通过聚合酶链式反应(PCR), 将接枝的ODNs扩增为长度为1300个碱基对的双链DNA, 从而制得CMC侧链上接枝DNA的共聚物CMC-g-DNA. 采用傅里叶红外光谱仪测定CMC与NHS形成的中间体; 用水平式琼脂糖凝胶电泳和垂直板变性聚丙烯酰胺凝胶电泳对CMC-g-DNA接枝共聚物进行表征. 结果表明, 合成了CMC-g-DNA接枝共聚物, 且在酸性条件下CMC的活化效果更好; 同时, 接枝在CMC上的DNA在琼脂糖凝胶电泳中迁移速率加快, 而在聚丙烯酰胺凝胶电泳中迁移速率减慢.  相似文献   

4.
A delayed crosslinking system employed with hydrolyzed polyacrylamide (HPAM) and chromium acetate was developed for high profile control in low-temperature reservoirs. The crosslinking system formed strong gel at polymer concentration range of 3000–5000 mg/L, and the gelation time was delayed to 8–30 days, which offered the crosslinking system enough time to flow into deep water-producing zones and plug large pore paths, forcing follow-up fluids to enter low-permeability layer and reduce disproportionate permeability. The effect of polymer hydrolysis degree, polymer concentration, pH, and crosslinker concentration on delaying gelation time was evaluated using bottle testing. Meanwhile, sand-packed tube displacement experiments revealed the plugging performance of delayed crosslinking system. According to Fourier transform infrared spectroscopy (FTIR) analysis, as well as crosslinking mechanism between polymer and chromium acetate, the delayed mechanism of crosslinked system demonstrated that due to stronger affinity of acetate complexes as a ligand, its substitution in situ by carboxylate group of polyacrylamide slowed down the initial rate-determining step of crosslinking reaction.  相似文献   

5.
通过酰氯化法与碳二亚胺缩合法(EDC/NHS)制备氨基化聚乙二醇(PEG1500N)修饰的多壁碳纳米管(MWNTs)并采用FTIR、Raman、TEM、原子力显微镜(AFM)、TGA-DTA-DSC、UV-Vis进行表征与分析。实验结果发现:两种方法PEG1500N都能很好地修饰MWNTs,但EDC/NHS缩合法采用更短的反应时间(反应1 d),达到了更好的接枝效果。EDC/NHS缩合法提高了碳管上羧基的利用率,接枝率大大提高。TGA-DTA分析表明缩合法接枝率为30%,而酰氯化法(反应4 d)为15%。UV-Vis分析表明EDC/NHS缩合法得到的产物溶解性也更好,溶解度由1.19 mg·mL-1(酰氯化法得到的产物的溶解度)提高到2 mg·mL-1以上。  相似文献   

6.
The kinetics of hydrolysis in the lowest excited single state of 6-methoxyquinoline in aqueous solutions can be evaluated from the pH-dependence of its fluorescence spectrum. In ethanol/water solutions, the rate constant for hydrolysis decreases exponentially with increasing mole fraction of ethanol and hydrolysis is completely inhibited at a mole fraction of ethanol greater than about 0.2. The slope of a plot of the logarithm of the hydrolysis rate constant against the mole fraction of ethanol was found to correlate well with the ratio of the diffusion volume of excited 6-methoxyquinoline to the molecular volume of ethanol. This also appears to explain why the linear plot of the logarithm of the hydrolysis rate constant against the logarithm of the water activity has a slope of 31, suggesting that 30 molecules of water, beside the one that is split in hydrolysis, participate in the reaction.  相似文献   

7.
分别通过1-(3-二甲氨基丙基)-3-乙基碳二亚胺盐酸盐/N-羟基丁二酰亚胺(EDC/NHS)偶联反应和席夫碱-还原反应,并以高碘酸钠(NaIO4)为氧化交联剂,得到2种邻苯二酚接枝改性的壳聚糖基水凝胶组织胶黏剂(CHI-C和CCS)。利用红外光谱(FT-IR)、核磁共振氢谱(~1 H-NMR)、抗菌实验和黏结强度测试对壳聚糖衍生物的结构和性能进行了表征。探讨了胶黏剂配比对其成胶性能和黏结强度的影响。结果表明:此种胶黏剂的成胶时间短,邻苯二酚基团和NaIO4含量的增加均有利于凝胶形成,CHI-C的最大黏结强度为40.42kPa,CCS的最大黏结强度为15.68kPa。两种壳聚糖衍生物均具有优良的抑菌性能。  相似文献   

8.
In the present research, molecular modeling methods were used to study novel porous soy protein conjugates with gelatin or alginate, which were recently developed as potential scaffolds for tissue engineering applications. Gelatin (protein) and alginate (polysaccharides) were chemically crosslinked to soy protein isolates (SPI) in order to obtain a porous 3D network. Computational tools were applied to estimate the crosslinking degree and compare the degradation rate of soy–gelatin or soy–alginate conjugates. Soy protein 3D structure was obtained from the Protein Data Bank (PDB). Alginate and gelatin structures were built and subjected to dynamic simulation using the molecular modeling package Material Studio 7.0. The crosslinking degree was estimated by the miscibility of the two reactants and the interaction with the crosslinking agents 1‐ethyl‐3‐(3‐dimethylaminopropyl) carbodiimide (EDC) or glyoxal. The calculations revealed that soy protein mixes well with gelatin but not with alginate. Radial distribution function (RDF) calculations showed that the interaction distance between alginate and EDC is significantly shorter than between gelatin and EDC, probably because of ionic attraction between the ammonium groups of EDC and the carboxylate groups in alginate, which facilitates the crosslinking reaction. The degradation rate of soy protein conjugates was related to their interaction with water. It was found that the solubility of soy–gelatin in water is higher than soy–alginate and that water molecules form more hydrogen bonds with soy–gelatin than with soy–alginate. These findings might be the reason for the observed difference in degradation rate of the two conjugates; the soy–gelatin degrades faster than soy–alginate. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

9.
The characteristics of poly(ethylene glycol) (PEG)–acrylate hydrogel networks were investigated as a function of the ethanol–water solvent composition during free‐radical crosslinking copolymerization. Macromonomer (88% ω‐methoxy‐PEG–acrylate and 10% ω‐phenoxy‐PEG–acrylate) and crosslinker (2% PEG–diacrylate) concentrations were kept constant. As the copolymerization progressed, the polymer solution in 100% ethanol became increasingly turbid, indicating the development of a heterogeneous network structure. In 100% water, however, the initially turbid polymer solution became increasingly transparent as the crosslinking copolymerization progressed. All the gels were optically clear upon equilibration in water. Kinetic studies, with attenuated total reflectance‐infrared, showed a long induction period, along with a lowered reaction rate, in 100% ethanol, and a decrease in conversion with an increase in ethanol content. These results agree with the UV analysis of the sol fractions, which indicated an increase in the amounts of unreacted PEG–acrylates with an increase in the ethanol content. The gels which were formed with a high ethanol concentration exhibited lower Young's modulus and higher swelling ability, suggesting that the network structure was significantly affected by the solvent composition during free‐radical crosslinking copolymerization. From the stress–strain and swelling experiments, the Flory–Huggins interaction parameter was evaluated. The creep characteristics of the hydrogels were modeled with two Kelvin elements. © 2002 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 40: 2677–2684, 2002  相似文献   

10.
Hydrophilic carboxyl-containing fluoromembranes were obtained by preirradiation grafting of acrylic acid onto ethylene-tetrafluoroethylene film. The dependence of the grafting reaction on temperature, monomer concentration, nature and concentration of inhibitor, crosslinking agent, solvent, and on the preirradiation dose was investigated. The grafting rates increase with temperature, whereas the saturation degree of grafting (SDG) decreases. Addition of inhibitor minimizes homopolymerization and at the same time hinders the grafting reaction. The SDG increases markedly with monomer concentration until it reaches a maximum and thereafter decreases. The grafting rates increase with preirradiation dose. Addition of crosslinking agent initially decreases the SDG, and thereafter increases. The highest grafting rates are obtained using water as solvent followed by methanol and ethanol. The results are discussed on the basis of various parameters: interaction between monomer diffusibility and the viscosity of the monomer bath, the mutual reactivity of monomer, and the crosslinking agent. An agreement is observed between the values of the electrical resistance and the saturation degree of grafting. © 1996 John Wiley & Sons, Inc.  相似文献   

11.
Novel biomimetic composite was prepared by the reaction of collagen and poly(γ‐benzyl L ‐glutamate)‐co‐poly(glutamic acid) (PBLG‐co‐PGA), which were crosslinked by non‐toxic crosslinking reagents 1‐ethyl‐(dimethylaminopropyl) carbodiimide (EDC) and N‐hydroxysuccinimide (NHS). The composite was characterized by FTIR and DSC. FTIR results confirmed that the collagen in the composite was successfully crosslinked with PBLG‐co‐PGA. DSC results showed that the composites possessed higher shrinkage temperature and higher thermal stability than the collagen. The water absorption test showed that the water absorbency of the composites increased with the increase in PBLG‐co‐PGA content in the composite. The studies of collagenase degradation and the tensile strength showed that the biostability and the tensile strength of the composites were significantly improved in comparison with that of the collagen. According to the investigations of cell adherent ratio and cell proliferation in vitro, the composite possessed good biocompatibility. Copyright © 2007 John Wiley & Sons, Ltd.  相似文献   

12.
The low-temperature chemical crosslinking of a copolymer of vinylidene fluoride and tetrafluoroethylene was studied through various physicochemical methods. The reaction was conducted in solution in the presence of diethylenetriamine as a crosslinking agent. The penetration of the mixture molecules and the crosslinking agent only in the amorphous phase of the copolymer was provided via selection of the ratio between a good solvent (dimethylformamide) and a poor solvent (ethanol). Owing to this, the crosslinking reaction occurred in the amorphous phase and hardly involved the crystals. This outcome was confirmed by wide-angle X-ray diffraction data. The structural and chemical changes in the amorphous phase during crosslinking were recorded with the use of IR spectroscopy and differential scanning calorimetry. It is shown that crosslinking is accompanied by the formation of C=C bonds in the copolymer chains. The study of high-voltage polarization and conductivity during exposure to bipolar rectangular pulses suggested that crosslinking leads to an increase in the carrier concentration. It was found that the surface potentials in the films increase with an increase in the number of high-voltage pulses applied to the sample. This circumstance is attributed to the fact that the double bonds formed in the copolymer chains can effectively trap negatively charged carriers.  相似文献   

13.
以苯乙烯(St)为主单体,丙烯酸(AA)为功能单体,乙二醇二甲基双丙烯酸酯(EGDMA)为交联单体,聚乙烯吡咯烷酮(PVP)为稳定剂,偶氮二异丁腈(AIBN)为引发剂,乙醇/水混合溶剂为分散介质,用分散聚合法一步合成了功能性单分散大粒径(10~20μm)交联聚苯乙烯微球.研究了PSt-AA-EGDMA三元分散共聚合体系的动力学,由转化率-时间关联得到动力学方程;Rp=k[I]0.13([St]1.87+[AA]0.13+[EGDMA]0.2)·(1+[PVP]0.2)exp(-E/RT).详细讨论了AA,EGDMA,PVP的浓度和溶剂极性对羧基分布的影响,阐述了AA和EGDMA对粒子形态、粒径及粒径分布的影响.  相似文献   

14.
Electrogenerated PPy doped with pSA was used as a substrate for immobilization of GOD. This was achieved via covalent bonding of carboxyl groups of the main chain of alginate with amino groups of the enzyme. The pH-induced aggregation behavior of SA in aqueous solution was employed to provide optimum conditions for electrochemical preparation of PPy by galvanostatic methods. GOD was attached to the electrode surface by reaction between the carboxyl groups in the main chain of pSA with amino groups of GOD after treatment with EDC and NHS. The linkage of GOD enzyme to the conductive surface was characterized by ATR spectroscopy and SEM CV was used to demonstrate the bioactivity of the enzyme electrode toward glucose.  相似文献   

15.
In this article, the fixed gel/nongel point and critical gel point experiments are designed for formulating the regulation of esterification reaction between epoxy resin and carboxyl acrylic polymer in the presence of tertiary amine in organic solvent. Ternary diagrams are employed to illustrate experimental results. It is disclosed that some factors, such as mole ratio of the ternary reactants, molecular weight of epoxy resin, even the hydrophilic solvent content of system, play important roles in the gelling effect of esterification reaction. In order to explain these experimental results, a suggested reaction process is put forward and, subsequently, a critical gel equation is derived from Carothers gelation theory. The equation provides an explicit connection between the critical gel composition and the relevant parameters of the reaction system such as average carboxyl functionality of acrylic resin, average molecular weight of epoxy resin, the base intensity of tertiary amine, and the solid content of system. The regressive critical gel curve based on the equation is well consistent with the experiment data. It expresses that the suggested reaction process is reasonable. Finally, based on the critical gel equation, the important factors that influence the esterification result are discussed qualitatively. All the achievements are beneficial to understanding the reaction process and avoiding useless gelation in preparing the water-reducible compositions.  相似文献   

16.
The kinetics of the hydrolysis of fenuron by hydrochloric acid in aqueous methanol solution was studied spectrophotometrically. The influence of cationic micelles of cetyltrimethylammonium bromide and anionic micelles of sodium lauryl sulfate on the rate of hydrolysis of fenuron have also been studied. The anionic micelles increased the rate of reaction, while the cationic micelles decreased the rate of hydrolysis. The reaction followed first-order kinetics in [fenuron]. The rate of reaction was increased with increase in [HCl] in lower range, but become constant at higher concentration in aqueous and micellar pseudophases. The reaction starts with the protonation of amino group of fenuron followed by rate-determining attack of water. The results in micellar media are accounted for on the basis of distribution of substrate into micellar and aqueous pseudophases.  相似文献   

17.
The kinetics of base-catalyzed hydrolysis of 4-nitrophenyl acetate was studied in water and water-ethanol mixtures under ultrasound (21.1 kHz) using a probe equipped with a quartz horn and, by comparison, without sonication. The ethanol concentration was varied from 0 to 50 wt % and pH from 7.5 to 9. The increase of reaction rate was inversely proportional to the strength of substrate interactions with the solvent structure. It was found that the effect of ultrasound could be related to the perturbation of solute-solvent interactions. This conclusion seems to apply to both acid-catalyzed and base-catalyzed hydrolysis reactions. Parallel experiments with a titanium horn revealed the catalytic action of undetermined chemical species formed from dispersed titanium metal.  相似文献   

18.
In this work, an electrochemical method based on the diazonium-coupling reaction mechanism for the immobilization of okadaic acid (OA) on screen printed carbon electrode was developed. At first, 4-carboxyphenyl film was grafted by electrochemical reduction of 4-carboxyphenyl diazonium salt, followed by terminal carboxylic group activation by N-hydroxysuccinimide (NHS), N-(3-dimethylaminopropyle)-N′-ethyle-carbodiimide hydrochloride (EDC). Hexamethyldiamine was then covalently bound by one of its terminal amine group to the activated carboxylic group. The carboxyl group of okadaic acid was activated by EDC/NHS and then conjugated to the second terminal amine group on other side of the hexamethyldiamine through amide bond formation. After immobilization of OA, an indirect competitive immunoassay format was employed to detect OA. The immunosensor obtained using this novel approach allowed detection limit of 1.44 ng/L of OA, and was also validated with certified reference mussel samples.  相似文献   

19.
Polybutadiene (PB) can be easily halogenated by reaction with iodine chloride or bromine in tetrahydrofuran. The resulting glassy polymers were reacted with n-butyllithium, sec-butyl-lithium, and polystyryllithium in THF. Iodochlorinated PB gave a polybutadiene with a different cis/trans ratio with n-BuLi. The reformation of PB was accompanied by partial crosslinking. The reaction probably involved a halogen-metal exchange followed by intra- and intermolecular elimination of Li halide. With brominated PB, both coupling and elimination took place. With sec-BuLi, an allylic iodine derivative was obtained from iodochlorinated PB, probably by dehydrochlorination. The iodinated intermediate can easily undergo a coupling reaction with further sec-BuLi. Both iodochlorinated and brominated polybutadienes gave graft copolymers by reaction with polystyryllithium in THF. Grafting was always accompanied by gel formation.  相似文献   

20.
In the present research, molecular modeling methods were used to study a novel bioadhesive composed of gelatin (protein) and alginate (polysaccharides), crosslinked with N‐(3‐dimethylaminopropyl)‐N′‐ethylcarbodiimide hydrochloride (EDC) and N‐hydroxysuccinimide (NHS). Three antibiotic drugs were added to the bioadhesive: Vancomycin, Ofloxacin, and Clindamycin. Computational tools were applied to estimate the crosslinking degree and compare the effect of the antibiotics on the physical properties of the gelatin‐alginate conjugate. The crosslinking degree was estimated by calculating the enthalpy of mixing of gelatin with alginate and their interaction with the crosslinking agents. The calculations revealed that gelatin mixes well with alginate, which enables their crosslinking. Various ratios between EDC and NHS were examined, and an optimal ratio was found. The interaction of alginate‐gelatin conjugate with the antibiotics was correlated to the experimental results of bonding strength. The most significant interaction of the conjugate is with clindamycin. The gelatin part is responsible for the strong interaction with clindamycin, and alginate forms strong interaction with ofloxacin. Thus, the interaction of alginate‐gelatin conjugate with the antibiotics is governed by the proportion between gelatin and alginate in the conjugate. The degradation rate of gelatin‐alginate was related to its interaction with water. It was found that the conjugate is highly hydrophilic. Gelatin is more soluble in water than both alginate and alginate‐gelatin and is probably the part in the conjugate that governs the solubility and degradation rate. Therefore, the degradation rate of the conjugate can be controlled by changing the proportion between gelatin and alginate.  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