首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The adsorption of nonionic surfactants on hide powder previously treated with anionic surfactants has been studied. The adsorption of nonionic surfactants takes place through hydrophobic interactions. A mechanism has been proposed for this interaction, assuming that the nonionic surfactant has been fixed by means of secondary adsorption (hydrophobic interaction) after the primary adsorption of the anionic surfactant (ionic and hydrophobic interaction) which makes it possible.  相似文献   

2.
The displacement of the proteins (beta-lactoglobulin and beta-casein) from an air-water interface by the nonionic (Tween 20 and Tween 60) and ionic (sodium dodecyl sulfate, cetyltrimethylammonium bromide, and lyso-phosphatidylcholine-lauroyl) surfactants has been visualized by atomic force microscopy (AFM). The surface structure has been sampled by the use of Langmuir-Blodgett deposition onto mica substrates to allow imaging in the AFM. In all cases, the displacement process was found to occur through the recently proposed orogenic mechanism (Mackie et al. J. Colloid Interface Sci. 1999, 210, 157-166). In the case of the nonionic surfactants, the displacement involved nucleation and growth of surfactant domains leading to failure of the protein network and subsequent loss of protein into the bulk phase. The surface pressure dependence of the growth of surfactant domains and the failure of the network were found to be the same for both Tween 20 and Tween 60, demonstrating that the breakdown of the protein film was dominated by the mechanical properties of the network. The displacement of protein by ionic surfactants was found to be characterized by nucleation of surfactant domains with little domain growth prior to failure of the network. The size of the domains formed by ionic surfactants was found to be limited by the strong intersurfactant repulsive forces between the charged headgroups. Screening of these charges led to an increase in the size of the domains. The surface pressure at which the network continuity was lost was found to be dependent on the type of surfactant and, in all cases, to occur at higher surface pressures than that required for nonionic surfactants. This has been attributed to surfactant-protein binding that initially strengthens the protein network at low surfactant concentrations. Evidence obtained from surface shear rheology supports this assertion.  相似文献   

3.
Treatment of higher plant (Nicotiana tabacum L. var. Samsun) chloroplast thylakoid membranes with surfactants results in a shift of the chlorophyll a absorption maximum in the red spectral region from its in vivo value of 678.5 nanometers to shorter wavelengths. The magnitude of this shift is correlated with membrane disruption, and is not necessarily due to the release of pigment from pigment-protein complexes present in the membrane. Membrane disruption has been measured by the amount of pigment in the supernatant fraction after centrifugation of surfactant treated membranes. For an equivalent amount of disruption, the extent of the blue-shift is influenced by the ionic nature of the surfactant: anionic surfactants cause small shifts, cationic surfactants cause the largest (~10 nanometers) shifts, and nonionic surfactants produce intermediate shifts. The wavelength of maximum absorbance of chlorophyll a in the red region is a convenient criterion for assessing the potential utility of different surfactants for studies on the structure, composition and function of higher plant thylakoid membranes.  相似文献   

4.
Monteiro MJ  Hall G  Gee S  Xie L 《Biomacromolecules》2004,5(5):1637-1641
A way to control the average pore size in cross-linked polyacrylamide-based membranes is by altering the ratio of cross-linker to acylamide monomer. Larger pore sizes are prepared with a minimum amount of cross-linker, resulting in membranes that are mechanically weak and have short lifetimes. The aim of this study was to prepare cross-linked polyacrylamide membranes with large pore sizes and with good mechanical integrity. The methodology was to carry out the polymerization in a template, formed from the self-aggregation of surfactant. Two surfactant templates were used, and their pore size was examined with proteins of different sizes. The surfactants chosen for this study were sodium dodecyl sulfate (SDS, ionic surfactant) and TERIC BL8 (nonionic surfactant), both of which have very different aggregation properties. The data showed that at 10% and greater of TERIC BL8, a very different and open gel structure is formed, in which the pore size was significantly increased. SDS seemed to have little effect on the pore size. The data suggests that the gel structures for both surfactants up to 4% (w/v) are similar and micellular, because SDS is known to favor a micelle structure. Above 4% (w/v), TERIC BL8 then goes through a change in its lyotropic phase, thus, producing membranes of a large pore size. In conclusion, the pore size and gel structure of polyacrylamide hydrogel membranes can be significantly increased using TERIC BL8 (nonionic) surfactant. This allows large-pore-size membranes with a high cross-link density and consequently high mechanical strength to be prepared for the separation of large biomolecules.  相似文献   

5.
Disintegration by surfactants of egg yolk phosphatidylcholine vesicles stabilized with carboxymethylchitin was investigated by measuring the amount released of a marker dye from the vesicles. In solutions of pH around 7, anionic and nonionic surfactants caused vesicle disintegration at very low concentrations, while cationic surfactants produced a breakdown of the vesicles at rather high concentrations. Increase in the alkyl chain-length of surfactant molecules brought about decrease in the surfactant concentration at which vesicle disintegration starts. As the length of the polyoxyethylene chain in nonionic surfactant molecules increased, the tendency of vesicle disintegration to occur decreased. Both anionic and cationic surfactants gave clear solutions above their critical micelle concentrations when they acted on the phospholipid vesicles, whereas nonionic surfactants left ghost cell-like debris consisting of carboxymethylchitin molecules in their micellar solutions. The effect of pH on vesicle disintegration was notable for ionic surfactants but not for nonionic surfactants. Thus, anionic surfactants increased the degree of disintegration as pH increased, while cationic surfactants produced an identical vesicle disintegration curve below pH 8 above which the curve started to shift toward the lower concentration region of the agents. These findings were explained in terms of surfactant penetration into phospholipid bilayers and solubilization of phospholipid molecules by surfactant micelles.  相似文献   

6.
Nonionic surfactants are used in a number of different microbiological applications, including solubilization of cell membranes, washing bacterial cultures prior to experimentation, and enhancing biodegradation of low-solubility compounds. An important consideration in these applications is the potential for the surfactant to alter the cell membrane. One potential means to monitor the impact of surfactants on the bacterial cell membrane is through monitoring the absorbance spectrum of the bacterial suspension. This is due to the colloidal nature of bacteria, where the absorbance of a bacterial suspension is related to the size and refractive index of the bacterial cells. Through a systematic study it was shown that there can be a significant change in the bacterial absorbance spectrum due to the presence of nonionic surfactants, with the effect a function of surfactant structure and concentration, solution ionic strength and cation valence. The effects were most pronounced with Na(+) as the cation, with surfactants having mid-range hydrophile-lipophile balance (HLB) values, and with surfactant concentrations above the CMC. The results indicate that measurement of the absorbance spectrum of bacterial cultures can provide a means to monitor the effects of nonionic surfactants on the bacterial cell membrane. In addition, depending on the specific application, appropriate selection of surfactant structure and media composition can be made to enhance or minimize the effects.  相似文献   

7.
In this work, four nonionic surfactants based on sorbitan monolaurate (Span 20) were synthesized by introducing ethylene oxide gas (n?=?20, 40, 60, 80 ethylene oxide units) into Span 20 to give four new surfactants with different hydrophilic-lipophilic balance (HLB), namely, E(20), E(40), E(60), and E(80). The structures of the prepared nonionic surfactants were elucidated using Fourier-transform infrared (FT-IR) and (1)H-nuclear magnetic resonance (NMR) spectroscopy. The surface-tension measurements were recorded. The effects of the prepared nonionic surfactants on the simultaneous saccharification and fermentation (SSF) of microwave/alkali-pretreated rice straw to produce ethanol were investigated. From the obtained data, it was found that the addition of the nonionic surfactants at 2.5?g/L had a positive effect on SSF. The maximum ethanol yield (76 and 55%) was obtained after 72?hr for rice straw using Kluyveromyces marxianus and Saccharomyces cerevisiae, respectively. Also, it was found that the ethanol yield increases with increasing HLB of the prepared nonionic surfactants by increasing ethylene oxide units. The adsorption of nonionic surfactants on lignocelluloses is proposed to be due to hydrophobic and hydrogen bonding interactions between nonionic surfactants and the lignin part in the lignocelulose. It can be concluded that additions of surface-active compounds, such as nonionic surfactants, increase enzymatic conversion of rice straw for bioethanol purposes.  相似文献   

8.
In this work, four nonionic surfactants based on sorbitan monolaurate (Span 20) were synthesized by introducing ethylene oxide gas (n = 20, 40, 60, 80 ethylene oxide units) into Span 20 to give four new surfactants with different hydrophilic–lipophilic balance (HLB), namely, E(20), E(40), E(60), and E(80). The structures of the prepared nonionic surfactants were elucidated using Fourier-transform infrared (FT-IR) and 1H-nuclear magnetic resonance (NMR) spectroscopy. The surface-tension measurements were recorded. The effects of the prepared nonionic surfactants on the simultaneous saccharification and fermentation (SSF) of microwave/alkali-pretreated rice straw to produce ethanol were investigated. From the obtained data, it was found that the addition of the nonionic surfactants at 2.5 g/L had a positive effect on SSF. The maximum ethanol yield (76 and 55%) was obtained after 72 hr for rice straw using Kluyveromyces marxianus and Saccharomyces cerevisiae, respectively. Also, it was found that the ethanol yield increases with increasing HLB of the prepared nonionic surfactants by increasing ethylene oxide units. The adsorption of nonionic surfactants on lignocelluloses is proposed to be due to hydrophobic and hydrogen bonding interactions between nonionic surfactants and the lignin part in the lignocelulose. It can be concluded that additions of surface-active compounds, such as nonionic surfactants, increase enzymatic conversion of rice straw for bioethanol purposes.  相似文献   

9.
表面活性剂对分枝杆菌KR2菌株降解菲的影响   总被引:2,自引:0,他引:2  
采用同位素示踪方法,从表面活性剂的浓度、离子类型和直链长度三方面研究了表面活性剂对分枝杆菌KR2菌株降解菲的影响。结果表明,表面活性剂的存在不能促进KR2菌对菲的降解;高浓度表面活性剂(≥20mg·L-1)的存在,使菲的降解出现延迟期,非离子表面活性剂Tween80在低浓度时(≤10mg·L-1)可以优先作为营养基质被分枝杆菌KR2菌株利用,表面活性剂的离子类型对菲降解的抑制作用的顺序为阳离子表面活性剂TDTMA>阴离子表面活性剂LAS>非离子表面活性剂Tween80,表面活性剂的直链长度对菲降解的影响为直链越短,对微生物的毒性越大,菲降解得越不完全。  相似文献   

10.
Hjelm R  Schedin-Weiss S 《Biochemistry》2007,46(11):3378-3384
Idraparinux is a synthetic O-sulfated, O-methylated pentasaccharide that binds tightly to antithrombin (AT) and thereby specifically and efficiently induces the inactivation of the procoagulant protease, factor Xa. In this study, the affinity and kinetics of the interaction of this high-affinity pentasaccharide with alpha- and beta-AT were compared with those of a synthetic pentasaccharide comprising the natural AT-binding sequence of heparin. Dissociation equilibrium constants, Kd, for the interactions of Idraparinux with alpha- and beta-AT were approximately 0.4 and 0.1 nM, respectively, corresponding to an over 100-fold enhancement in affinity compared with that of the normal pentasaccharide. This large enhancement was due to a approximately 400-fold tighter conformationally activated complex formed in the second binding step, whereas the encounter complex established in the first step was approximately 4-fold weaker. The high-affinity and normal pentasaccharides both made a total of four ionic interactions with AT, although the high-affinity saccharide only established one ionic interaction in the first binding step and was compensated by three in the second step, whereas the normal pentasaccharide established two ionic interactions in each step. In contrast, the affinities of the nonionic interactions (Kd approximately 450 and 90 nM for the binding to alpha- and beta-AT, respectively) were considerably higher than those for the normal pentasaccharide and the highest of all AT-saccharide interactions reported so far. The nonionic contribution to the total free energy of the high-affinity pentasaccharide binding to AT thus amounted to approximately 70%. These findings show that nonionic interactions can play a predominant role in the binding of highly charged saccharide ligands to proteins and can be successfully exploited in the design of such biologically active ligands.  相似文献   

11.
Removal of nonionic surfactants from municipal wastewater using a constructed wetland with a horizontal subsurface flow was studied in 2009 and 2010. Extraction spectrophotometry with 3′,3″,5′,5″‐tetrabromophenolphthalein ethyl ester and KCl served to determine the analyte concentrations. Triton® X‐100 was used as a standard to express the nonionic‐surfactant concentrations. Anionic and cationic surfactants were shown not to interfere during the determination. Nonionic surfactants were degraded (to products undeterminable by the method) with a high average efficiency that reached 98.1% in 2009 and 99.1% in 2010, respectively. The average concentration of nonionic surfactants at the inflow was 0.978 mg/l, while it was close to the limit of quantification at the outflow (0.014 mg/l). A significant fraction of nonionic surfactants (38.7%) was already degraded during the pretreatment, and only 14.0% of the nonionic surfactants remained in the interstitial H2O taken in the vegetation bed at a distance of 1 m from the inflow zone at a 50‐cm depth. Nonionic surfactants were degraded both under aerobic and anaerobic conditions.  相似文献   

12.
Thoppil AA  Sharma R  Kishore N 《Biopolymers》2008,89(10):831-840
Binding of the antibiotic drug carbenicillin to bovine serum albumin (BSA) has been studied using isothermal titration calorimetry (ITC) in combination with fluorescence and circular dichroism (CD) spectroscopies. The thermodynamic parameters of binding have been evaluated as a function of temperature, ionic strength, and in the presence of anionic, cationic and nonionic surfactants, tetrabutylammonium bromide, and sucrose. The values of van't Hoff enthalpy do not agree with the calorimetric enthalpy indicating conformational changes in the protein upon drug binding. These observations are supported by the intrinsic fluorescence and CD spectroscopic measurements. A reduction in the binding affinity of carbenicillin to BSA is observed with increase in ionic strength of the solution, thereby suggesting, prevailing of electrostatic interactions in the binding process. The involvement of hydrophobic interactions in the binding of the drug to the protein is also indicated by a slight reduction in binding constant in the presence of tetrabutylammonium bromide. The experiments in the presence of sucrose suggest that hydrogen bonding is perhaps not dominant in the binding. The anionic surfactant sodium dodecyl sulphate (SDS) is observed to completely interfere in the ionic interactions in addition to its partial denaturing capacity. However, the presence of cationic surfactant hexadecyl trimethylammonium bromide (HTAB) and nonionic surfactant Triton-X 100 induce a slight reduction in the values of binding affinity. These calorimetric and spectroscopic results, provide quantitative information on the binding of carbenicillin to BSA and suggests that the binding is dominated by electrostatic interactions with contribution from hydrophobic interactions. (c) 2008 Wiley Periodicals, Inc. Biopolymers 89: 831-840, 2008.This article was originally published online as an accepted preprint. The "Published Online" date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com.  相似文献   

13.
Research within the field of colloidal liquid aphrons (CLAs) for enzyme immobilization has often used ionic surfactants for the retention of enzymes. Although these charged interactions allow for enhanced immobilization, they can often lead to denaturation of enzyme activity, and even release of the protein. Sodium alginate has been used in drug delivery applications due to its low toxicity and charged interactions that allow for encapsulation. Hence, alginate systems can be used as an alternative to ionic surfactants in CLA immobilization. This paper presents, for the first time, the use of sodium alginate as potential ligand for enhanced CLA immobilization. The use of five model proteins; lysozyme, bovine serum albumin, ovalbumin, insulin, and α-chymotrypsin, of various pIs and hydrophobicities, showed the relevance of electrostatic interactions in promoting binding with sodium alginate when the pH < pI, with 100% immobilization attributed to alginate incorporated CLAs over general nonionic formulations. Furthermore, above their pI, >80% protein recovery was observed, with activity and conformation comparable to their native counterparts. Finally, the use of proteolysis showed that as the degree of ionic bonding increased between the protein and sodium alginate, the degree of protease resistance decreased due to conformational changes experienced during binding.  相似文献   

14.
Effects of five homologous series of amphoteric, anionic, and nonionic surfactants on growth and anthocyanin production by callus cultures of Dimorphotheca sinuata were examined. The phytotoxicity of amphoteric sulfobetaines increased with alkyl chain length and reached a plateau at 14 to 16 carbon atoms. In the case of the anionic sodium alkyl sulfates, the dodecyl derivative caused the greatest inhibition. Higher molecular weight ionic detergents were less toxic, probably due to their high Krafft temperatures and concomitant reduced solubility. The nonionic higher alcohol ethoxylates were generally less inhibitory than the ionic detergents and had little or no effect below their critical micelle concentrations. Fresh weights of tissues and anthocyanin production by Dimorphotheca callus cultures declined with increasing surfactant concentrations.  相似文献   

15.
S R Stone  S Dennis  J Hofsteenge 《Biochemistry》1989,28(17):6857-6863
The effect of ionic strength on the kinetics of inhibition of human alpha-thrombin has been examined by using genetically engineered forms of hirudin that differed only in the number of negatively charged residues in the carboxyl-terminal region of the molecule. Analysis of the data obtained allowed the binding energy for the thrombin-hirudin complex to be divided into contributions from ionic and nonionic interactions. The contribution of nonionic interactions to the binding energy was the same for each of the forms whereas the ionic contribution varied with the charge of the molecule. Each of the negatively charged residues made an approximately equal contribution of -4kJ mol-1 to the binding energy. For native hirudin, ionic interactions accounted for 32% of the binding energy at an ionic strength of zero.  相似文献   

16.
The hemolytic action of a number of homologous series of cationic surfactants on human erythrocytes was measured. The hemolytic effects of anionic, nonionic and cationic surface-active agents are compared. The relationship which exists between the key physicochemical properties of surfactants (critical micelle concentration, hydrophile-lipophile balance) and their hemolytic capacities is discussed. The parameters required to compare the actions of various surfactants on different cellular membranes are considered in relation to the study of the correlation between the surfactant lytic effects and the features of the membrane molecular organization.  相似文献   

17.
We analyzed the binding and fusogenic properties of surfactant vesicles (SVs), composed of ionic and nonionic surfactants and cholesterol, with the surface of different human lymphoid cells. The influence of charge on SVs-cell interaction was evaluated by monitoring the presence of fluorescent sodium calcein artificially entrapped in the vesicles using optical fluorescence microscopy and laser scanning confocal microscopy. Our results clearly indicate that only negatively charged vesicles bind and fuse with the plasma membrane of human lymphoid cells, and the number of SVs bound to the cell surface was variable among the positive cells. Thin section electron microscopy illustrated that the fusogenic events of SVs with the cell plasma membrane mostly occurred at smooth and nonvillous regions of the cell surface. Taken together, our results suggest that binding and fusion of SVs with the cell plasma membrane might be dependent on interactions with specific membrane components that preferentially recognize negatively charged SVs.  相似文献   

18.
Zwitterionic detergents interfere with the salt-induced phase separation for nonionic detergents in a concentration-dependent manner by shifting the normal cloud point of nonionic detergents to a higher ionic strength at room temperature. This phenomenon was used to determine the concentration of the zwitterionic detergents CHAPS, CHAPSO, and sulfobetaine SB-12 in solution by titration with ammonium sulfate in the presence of Triton X-100. Among the ionic detergents tested, the method was only applicable to sodium cholate. The assay can be used to control the removal of zwitterionic detergents during the reconstitution of membrane proteins in liposomes. However, it cannot be used to determine the specific binding of zwitterionic detergents to highly diluted, pure membrane proteins because of the limited sensitivity. Neither proteins nor phospholipids interfered with this method at concentrations up to 20 mg/ml of test solution (human serum albumin) or 10 mg/ml (phospholipids), respectively. Since the assay is based on the competition between salts and nonionic detergents for water molecules, it is important to equalize the ionic strength of samples and calibration standards.  相似文献   

19.
Nonionic and ionic surfactants diminish the initial rate of proteolysis of aqueous bovine serum albumin (BSA) by subtilisin Carlsberg. Surfactants studied include: nonionic tetraethylene glycol monododecyl ether (C12E4); anionic sodium dodecyl sulfate (SDS), anionic sodium dodecylbenzenesulfonate (SDBS), and cationic dodecyltrimethylamonium bromide (DTAB). Kinetic data are obtained using fluorescence emission. Special attention is given to enzyme kinetic specificity determined by fitting initial-rate data to the Michaelis-Menten model. All surfactants reduce the rate of proteolysis, most strongly at concentrations near and above the critical micelle concentration (CMC). Circular dichroism (CD), tryptophan/tyrosine fluorescence spectra, and tryptophan fluorescence thermograms indicate that BSA partially unfolds at ionic surfactant concentrations near and above the CMC. Changes in BSA conformation are less apparent at ionic surfactant concentrations below the CMC and for the nonionic surfactant C12E4. Subtilisin Carlsberg activity against the polypeptide, succinyl-Ala-Ala-Pro-Phe-p-nitroanilide, decreased due to enzyme-surfactant interaction. At the concentrations and time frames studied, there was no enzyme autolysis. Importantly, aqueous proteolysis rates are significantly reduced at high surfactant concentrations where protein-micellar-surfactant aggregates occur. To explain the negative effect of surfactant on subtilisin Carlsberg proteolytic activity against BSA, we propose that micelle/protein complexes hinder enzyme access.  相似文献   

20.
不同类型表面活性剂在土壤上的吸附特征比较研究   总被引:18,自引:0,他引:18  
应用平衡振荡法,研究了阴、阳和非离子表面活性剂在土壤上的吸附.结果表明,阳离子表面活性剂十六烷基三甲基溴化铵能强列吸附在6种不同性质的土壤上,吸附等温线为L型,分配常数Kd,为3.0×10^2~48×10^2L·kg^-1;阴离子表面活性剂十二烷基苯磺酸钠、非离子表面活性剂OP及Tween-20的吸附等温线随土壤类型不同而不同,有L、S等型,吸附强度远弱于阳离子表面活性剂,Kd分别大体处于5.3~39、0.13~0.44(Tween-20)和4.4~22.4L·kg^-1(OP).阳离子表面活性剂的土壤最大吸附量与土壤阳离子交换容量呈线性相关.低浓度范围内,阴离子表面活性剂的土壤分配常数与土壤粘粒含量呈正相关.同时土壤颗粒表面的电荷特性也影响吸附.非离子表面活性剂的Kd与土壤粘粒、砂粒、粉沙含量及表面积存在经验函数关系.  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