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 562 毫秒
1.
1. The interaction of a wide range of surfactants with isolated gill epithelial cells of rainbow trout (Oncorhynchus mykiss) was investigated as a function of the surfactant concentration up to and above the critical micelle concentration (cmc). The surfactants included a homologous series of n-alkyl sulphates, single and double chain tri and dimethylammonium bromides (TABs and DABs), cholates and the nonionics n-octylglucoside and Triton X-100.2. With the exception of the C22 alkyl chain TAB and the double chain [(C12)2] DAB, the surfactants solubilized between 84 and 100% of the cell protein at high concentrations (>cmc).3. At low concentrations n-dodecyltrimethylammonium bromide and, to a lesser extent, Triton X-100 and sodium n-dodecylsulphate release a larger proportion of cell protein than they solubilize lipid in contrast to sodium cholate which initially preferentially solubilizes cell lipid. This differential pattern of solubilization is similar to that observed for other plasma membranes such as those of human erythrocytes and platelets.4. The surfactant concentration required to solubilize 50% (S50%) of cell protein increases with the cmc. There is an approximately linear relationship between log(S50%) and log cmc.5. Light microscopy shows that the surfactants at high concentrations (>cmc) fragment the secondary lamellae of the gill filaments.  相似文献   

2.
The insolubility of carbon nanotubes (CNTs) in aqueous media has been a limitation for the practical application of this unique material. Recent studies have demonstrated that the suspend ability of CNT can be substantially improved by employing appropriate surfactants. Although various surfactants have been tested, the exact mechanism by which carbon nanotubes and the different surfactants interact is not fully understood. To deepen the understanding of molecular interaction between CNT and surfactants, as well as to investigate the influence of the surfactant tail length on the adsorption process, we report here the first detailed large-scale all-atomistic molecular dynamics simulation study of the adsorption and morphology of aggregates of the cationic surfactants containing trimethylammonium headgroups (C12TAB and C16TAB) on single-walled carbon nanotube (SWNT) surfaces. We find that the aggregation morphology of both C12TAB and C16TAB on the SWNT is dependent upon the number of the surfactants in the simulation box. As the number of the surfactants increases the random monolayer structure gradually changes to the cylinder-like monolayer structure. Moreover, we make a comparison between the C12TAB and C16TAB adsorption onto SWNTs to clarify the role of the surfactant tail length on the adsorption process. This comparison indicates that by increasing the number of surfactant molecules, the larger number of the C16TAB molecules tend to adsorb onto SWNTs. Further, our results show that a longer chain yields the higher packed aggregates in which the surfactant heads are extended far into the aqueous phase, which in turn may increase the SWNTs stabilization in aqueous suspensions.  相似文献   

3.
The effect of a homologue series of nonionic surfactants, namely poly(ethylene glycol) (PEG) fatty acid esters, differing in oxyethylene (PEG 8, PEG 12, and PEG 40) and fatty acid (stearate, mono and di-laurate, and mono and di-oleate) chain lengths, on in vitro skin permeability of ketoprofen (KTP) vehicled in plasters was investigated. The drug diffusion through hairless mouse skin as well as the effect of the surfactant type and strength was studied by Franz diffusion cells and ATR-FTIR spectroscopy. The use of PEG stearate series revealed that the surfactant with the largest polar head, namely PEG 40, was ineffective in enhancing the skin permeation of KTP, independently of the plaster concentrations. The effect of the hydrophobic chain was investigated only by using the shortest oxyethylene chains. The experimental results revealed that the oxyethylene chain length of surfactants appeared to be more influent than the alkyl chain. The prediction of the absorption enhancing capability of these PEG derivatives appeared related to the vehicle other than the proper combination of the number of ethylene oxide groups and alkyl groups.  相似文献   

4.
The effects of ethoxy (EO) chain length on surfactant-induced ethylene production for selected octylphenoxy (OP) and linear alcohol (LA) surfactants were established using primary leaves of cowpea (Vigna unguiculata (L.) Walp. subsp. unguiculata Dixielee) seedlings. OP surfactant-induced ethylene production was concentration dependent and decreased log linearly with increasing EO content. C12–15 LA-induced ethylene production decreased log linearly with increasing EO content at 0.1%; however, at 1.0% the relationship was curvilinear with maximum response at 7 EO. Relationships for the C9–11 and C9 LA series were nonlinear with greatest biological activity at intermediate (8–12) EO content. Short EO chain length OP surfactants were only slightly water soluble, and induced low levels of ethylene production and phytotoxicity. Addition of OP+1EO to a long chain, water soluble, non-ethylene inducing surfactant (OP+40EO) solution significantly increased ethylene production by OP+1EO in cowpea. A similar response was found for surfactant-induced phytotoxicity and EO chain length as between ethylene production and EO content. Similar EO chain length and ethylene production relationships were found for germinating mung bean (Vigna radiata (L.) R. Wilcz) seeds as for ethylene production and phytotoxicity in cowpea. Radicle growth was markedly inhibited by OP surfactants with an EO chain length of 10 or less and, in some cases, radicles were irreversibly damaged by ethylene inducing surfactants.  相似文献   

5.
Quaternary ammonium salts substituted with a long alkyl chain exemplify a trustworthy group of medicinal compounds frequently employed as antifungal and antibacterial agents. A great asset of these surfactants underlying their widespread use is low local and system toxicity in humans. In this Letter, a series of novel quaternary 6-hydroxyquinolinium salts with varying length of N-alkyl chain (from C10 to C18) was synthesized and tested for in vitro activity against pathogenic bacterial and fungal strains. 6-Hydroxyquinolinium salt with C12 alkyl chain seems to be very interesting candidate due to a high antimicrobial efficacy and cytotoxic safety.  相似文献   

6.
The crystal structures of alkyl 2-deoxy-α-d-arabino-hexopyranosides, with the alkyl chain lengths from C8 to C18, are established by the single crystal X-ray structural determination. The even-alkyl chain length derivatives crystallized orthorhombic, with space group P212121, whereas the odd-alkyl chain length derivatives crystallized monoclinic, with space group P21. The sugar moieties retained a 4C1 chair conformation and the conformation of the alkyl chains was all-trans. The molecules formed a bilayer structure, in which alkyl chains were interdigitated. The hydrogen bonds, originating from the sugar moieties, were observed in adjacent layers and also within the same layer, resulting in the formation of infinite chains. The alkyl chains arranged parallel to each other and formed planar structures. The thermal properties of the alkyl 2-deoxy glucosides were analyzed further. It was observed that none of the derivatives exhibited mesomorphism. This study establishes that the absence of the hydroxyl group at C-2 of the sugar moiety results in a non-mesogenic nature of the alkyl 2-deoxy-α-d-glycosides, as opposed to the profound mesogenic nature of the normal alkyl glycosides.  相似文献   

7.
Five non-ionic dialkylglycerol poly(oxyethylene) ether surfactants, designated 2CmEn (where m, the number of carbons in each alkyl chain = 16 or 18, and n, the number of oxyethylene units = 12, 16 or 17) have been examined for their ability to form vesicles when dispersed in water or in an aqueous solution of 154 mM NaCl, alone or in the presence of 50 mol% cholesterol. Freeze fracture electron microscopy and light scattering showed that regardless of the hydrating fluid, all the non-ionic surfactants, with the exception of 2C16E17 and 2C18E17, formed vesicles in the absence of cholesterol – 2C16E17 and 2C18E17 instead formed micellar aggregates. All surfactants, however, formed vesicles in the presence of 50 mol% cholesterol. Small angle neutron scattering studies of the surfactant vesicles enabled the bilayer thickness and repeat distance (d-spacing) to be determined. The bilayers formed by all the non-ionic surfactants in the absence of cholesterol were surprisingly thin (∼50 Å for the E12 containing surfactants and ∼64 Å for 2C18E16) most likely due to the intrusion of oxyethylene groups into the hydrophobic core of the bilayers. In contrast, however, the non-ionic surfactants exhibited a relatively large d-spacing of around ∼130–150 Å. The addition of 50 mol% cholesterol had a dramatic effect on the thickness of the vesicle bilayer, increasing its size by 10–20 Å, most probably because of an extrusion of oxyethylene from the hydrophobic region of the bilayer and/or a reduction in the tilt on the surfactant alkyl chains. Additionally the presence of cholesterol in a vesicle tended to reduce slightly both the d-spacing and the thickness of the water layer separating the bilayers. The presence of NaCl, even at the low concentrations used in the study, did affect the properties of the bilayer such that it reduced the d-spacing and, in the case of cholesterol-containing systems, also reduced bilayer thickness.  相似文献   

8.
In in vitro tests, amines were screened for the inhibition of polar filament extrusion by spores ofPlistophora anguillarum, a microsporidian parasite of the eel. Primary amines having C8 C18 alkyl chains were effective, irrespective of the anions bound. Secondary amines having a branched C8 chain and tertiary amines having C14−C18 chains were also effective. Asymmetry in side chains seemed to be required for inhibitory action. Quaternary ammonium salts having C12−C16 alkyl chains and some germicides or disinfectants such as cetyl pyridinium chloride and benzethonium chloride were also effective. Diamines and amides were ineffective regardless of the length of their alkyl chins. In in vivo tests, eel larvae-fed for 4 days with a commercial feed supplemented with 5×108 cells of spores and 4g of laurylamine aspartate/100 g feed, and thereafter with the usual commercial feed for 26 days—were not infected.  相似文献   

9.
Abstract

Simulation studies of the friction between layers of dialkyl surfactants have been performed using a non-equilibrium molecular dynamics method. The model layers at a temperature of 298 K and a normal pressure of 210 MPa are sheared at a relative velocity of 1 ms?1. The friction coefficient has been studied as a function of the molecular geometry by simulating bilayers of C10C18 (asymmetrical) and C18C18 and C10C10 (symmetrical) surfactants. In all cases the hydrocarbon chains are attached to a positively charged dimethylammonium head-group which interacts with a negatively charged surface. At a head-group area of 50 Å2 per molecule, the friction between the layers of asymmetrical surfactants is greater than that between layers of symmetrical surfactants at approximately the same normal pressure. At 77 Å2 the friction between the C18C10 layers remains higher than that of the C18C18 layers but is now lower than that of the C10C10, where the surface structure is highly disordered and the two layers are separated by only 15.8 Å. The friction between the layers correlates well with the amount of layer overlap as defined by the common area under the chain density profiles. These observation, which are in broad aggreement with the experimental measurement on similar dichain surfactants are rationalised in terms of the translational, orientational and conformational structures of the layers.  相似文献   

10.
Abstract

The effect of C12E8 (polyoxyethylene 8 lauryl ether) and sodium laurate on the structural properties of soybean lecithin vesicles was studied in different concentrations of surfactants, using the fluorescent probes NBD-PE (N-7-nitro-2,1,3-benzodiazoyl phosphatidylethanolamine) and N-Rh-PE (N-lissamine rhodamine B sulfonyl phosphatidylethanolamine). Direct energy transfer studies were carried out in labelled vesicles with addition of surfactants. Rhodamine emission with maximum at β585 nm was detected by excitation of NBD at λ = 473 nm. This fact is caused by direct energy transfer process from NBD to rhodamine. The yield of this process decreases with increasing amounts of surfactants, indicating that the average spatial separation of the lipid probes increases, as a result of the enlargement of the vesicle size and due to alteration of its structure to mixed micelles.  相似文献   

11.
Intraspecific variation in four New Zealand species of Chionochloa, C. flavescens, C. pallens, C. rigida; and C. rubra, was investigated by examining the major carbon chain lengths of fatty acids, alcohols, aldehydes, wax esters and alkanes of the epicuticular waxes. The major even-carbon chain lengths ranged generally from C24 to C32 in the acids, alcohols and aldehydes; C29 to C33 in the alkanes; and even-carbon chains between C36 and C52 in the wax esters. A computer program was used to calculate the degree of similarity between samples in terms of chain length distribution. In C. rigida eastern and western South Island localities were identified; in C. flavescens Canterbury and Nelson, western South Island and southern North Island regions were recognized; and C. pallens and C. rubra were divisible into four regions; Canterbury, Nelson, western South Island and southern North Island. The possible elongation-decarboxylation pathways and the specificity of the enzymes in the biosynthetic pathways of epicuticular wax synthesis suggest the possibility that the northwest Nelson region could be a biogenetic centre from which wax synthesis has diversified along three routes, one to the western South Island, another to eastern South Island and the third to southern North Island. Identification of each of the four species based on the distribution of the carbon chain lengths in the individual lipid fractions is impossible unless the locality of collection is known. Intraspecific variation in lipid composition is not coincident with patterns of variation already reported.  相似文献   

12.
Interactions between melittin and a variety of negatively-charged lipid bilayers have been investigated by intrinsic fluorescence, fluorescence polarization of 1,6-diphenylhexatriene and differential scanning calorimetry. (1) Intrinsic fluorescence of the single tryptophan residue of melittin shows that binding of this peptide to negatively-charged phospholipids is directly related to the surface charge density, but is unaffected by the physical state of lipids, fluid or gel, single-shell vesicles or unsonicated dispersions. (2) Changes in the thermotropic properties of negatively-charged lipids upon melittin binding allow to differentiate two groups of lipids: (i) A progressive disappearance of the transition, without any shift in temperature, is observed with monoacid C14 lipids such as dimyristoylphosphatidylglycerol and -serine (group 1). (ii) With a second group of lipids (group 2), a transition occurs even at melittin saturation, and two transitions are detected at intermediate melittin content, one corresponding to remaining unperturbed lipids, the other shifted downward by 10–20°C. This second group of lipids is constituted by monoacid C16 lipids, dipalmitoylphosphatidylglycerol and -serine. Phosphatidic acids also enter this classification, but it is the net charge of the phosphate group which allows to discriminate: singly charged phosphatidic acids belong to group 2, whereas totally ionized ones behave like group 1 lipids, whatever the chain length. (3) It is concluded that melittin induces phase separations between unperturbed lipid regions which give a transition at the same temperature as pure lipid, and peptide rich domains in which the stoichiometry is 1 toxin per 8 phospholipids. The properties of such domains depend on the bilayer stability: in the case of C16 aliphatic chains and singly charged polar heads, the lipid-peptide domains have a transition at a lower temperature than the pure lipid. With shorter C14 chains or with two net charges by polar group, the bilayer structure is probably totally disrupted, and the new resulting phase can no longer lead to a cooperative transition.  相似文献   

13.
Changes in physico-chemical properties of dimyristoyl phosphatidylcholine (DMPC) lipid bilayers caused by the addition of 9.4 mol% nonionic surfactant decaoxyethylene monododecyl ethers (C12E10) have been investigated by molecular dynamics calculations. In spite of addition of single chain C12E10, the lipid bilayers showed an increase of membrane area. Isothermal area compressibility, which is a measure of membrane softness in lateral direction, also increased by 50% for DMPC/C12E10 mixed bilayers. Furthermore, the order parameter of C–H vector for DMPC acyl tails decreased. We found that these changes are caused by the hydrophilic head groups of C12E10 which are located near the glycerol backbone of the DMPC molecules and have bulky random coil conformation without any preferential ordered structures.  相似文献   

14.
Various alkyltrimethylammonium bromides ATABs, R = C1 to C6 were added to a mixed liquor of sewage activated sludge and the effects on settling properties were investigated. The 30-min settled sludge volume decreased from 28% to 24, 20, 17, and 15% on addition of 1 mM C10, C12, C14, and C16 compounds to mixed liquor (sludge concentration, 1.28 g/l), respectively. With increasing alkyl chain length, the amounts of ATABs adsorbed by the sludge, and metal ions released from the sludge, increased. However, a quantitative correlation was not observed between the two. In addition, the adsorption of ATABs resulted in a small decrease in sludge wet weight. However, such a change was not observed on addition of ATABs bearing short alkyl chains. A good correlation was observed between the amounts of ATABs adsorbed and their organic solvent/water partition coefficients. These results indicated that the effectiveness of ATABs as sludge conditioners depends on their hydrophobicity.  相似文献   

15.
Aims: To investigate if Burkholderia glumae can produce rhamnolipids, define a culture medium for good production yields, analyse their composition and determine their tensioactive properties. Methods and Results: Burkholderia glumae AU6208 produces a large spectrum of mono‐ and di‐rhamnolipid congeners with side chains varying between C12‐C12 and C16‐C16, the most abundant being Rha‐Rha‐C14‐C14.The effects on rhamnolipid production of the cultivation temperature, nitrogen and carbon source were investigated. With urea as the nitrogen source and canola oil as the carbon source, a production of 1000·7 mg l?1 was reached after 6 days. These rhamnolipids display a critical micelle concentration of 25–27 mg l?1 and decrease the interfacial tension against hexadecane from 40 to 1·8 mN m?1. They also have excellent emulsifying properties against long chain alkanes. Conclusions: Burkholderia glumae AU6208 can produce considerable amounts of rhamnolipids. They are produced as diversified mixtures of congeners. Their side chains are longer than those normally produced by those of Pseudomonas aeruginosa. They also present excellent tensioactive properties. Significance and Impact of the Study: In contrast with the classical rhamnolipid producer Ps. aeruginosa, B. glumae is not a pathogen to humans. This work shows that the industrial production of rhamnolipids with this species could be easier than with Ps. aeruginosa.  相似文献   

16.
The membrane lipid environment and lipid signaling pathways are potentially involved in the modulation of the activity of the cardiac Na+-Ca2+ exchanger (NCX). In the present study biophysical mechanisms of interactions of amphiphiles with the NCX and the functional consequences were examined. For this purpose, intracellular Ca2+ concentration jumps were generated by laser-flash photolysis of caged Ca2+ in guinea-pig ventricular myocytes and Na+-Ca2+ exchange currents (INa/Ca) were recorded in the whole-cell configuration of the patch-clamp technique. The inhibitory effect of amphiphiles increased with the length of the aliphatic chain between C7 and C10 and was more potent with cationic or anionic head groups than with uncharged head groups. Long-chain cationic amines (C12) exhibited a cut-off in their efficacy in INa/Ca inhibition. Analysis of the time-course, comparison with the Ni2+-induced INa/Ca block and confocal laser scanning microscopy experiments with fluorescent lipid analogs (C6- and C12-NBD-labeled analogs) suggested that amphiphiles need to be incorporated into the membrane. Furthermore, NCX block appears to require transbilayer movement of the amphiphile to the inner leaflet (“flip”). We conclude that both, hydrophobic and electrostatic interactions between the lipids and the NCX may be important factors for the modulation by lipids and could be relevant in cardiac diseases where the lipid metabolism is altered.This revised version was published online in August 2005 with a corrected cover date.  相似文献   

17.
In the preceding paper on the interrelation between sucrose ester of fatty acid and biotin, the fatty acid being a mixture of C10 to C18 acid, it was described that carbon chain length of fatty acid has a great influence on the accumulation of l-glutamic acid. Fatty acids with C12 to C18 chain length, particularly myristic, palmitic and margaric acids were effective on the accumulation of l-glutamic acid in the culture medium containing sufficient biotin, whereas lower and higher length acids were ineffective. In the form of polyoxyethylene sorbitan or polyethylene glycol ester, C16 and C18 acids were remarkably effective. However, the ester of C12 acid and polyoxyethylene ethers of C12 to C18 alcohols had little or no effect.  相似文献   

18.
When n-alkanes with 13 to 20 carbon atoms were fed to a Nocardia closely related to N. salmonicolor, the produced cellular triglycerides and aliphatic waxes invariably contained fatty acids with an even or an odd number of carbon atoms subject to this feature of the n-alkane substrate. Beta-oxidation and C2 addition are both operative, as evidenced by the spectra of fatty acids incorporated into the cellular lipid components. There is no distinction in the rate of microbial incorporation of the odd-or even-numbered carbon chains. The fatty acids are apparently directly derived from the long chain n-alkanes, rather than synthesized via the classic C2-condensation route. The alcohol component of waxes produced by the Nocardia is invariably of the same chain length as the n-alkane substrate.  相似文献   

19.
PEGylation induced changes in molecular volume and solution properties of HbA have been implicated as potential modulators of its vasoconstrictive activity. However, our recent studies with PEGylated Hbs carrying two PEG chains/Hb, have demonstrated that the modulation of the vasoconstrictive activity of Hb is not a direct correlate of the molecular volume and solution properties of the PEGylated Hb and implicated a role for the surface charge and/or the pattern of surface decoration of Hb with PEG. HbA has now been modified by thiolation mediated maleimide chemistry based PEGylation that does not alter its surface charge and conjugates multiple copies of PEG5K chains. This protocol has been optimized to generate a PEGylated Hb, (SP-PEG5K)6-Hb, that carries ~six PEG5K chains/Hb – HexaPEGylated Hb. PEGylation increased the O2 affinity of Hb and desensitized the molecule for the influence of ionic strength, pH, and allosteric effectors, presumably a consequence of the hydrated PEG-shell generated around the protein. The total PEG mass in (SP-PEG5K)6-Hb, its molecular volume, O2 affinity and solution properties are similar to that of another PEGylated Hb, (SP-PEG20K)2-Hb, that carries two PEG20K chains/Hb. However, (SP-PEG5K)6-Hb exhibited significantly reduced vasoconstriction mediated response than (SP-PEG20K)2-Hb. These results demonstrate that the enhanced molecular size and solution properties achieved through the conjugation of multiple copies of small PEG chains to Hb is more effective in decreasing its vasoconstrictive activity than that achieved through the conjugation of a comparable PEG mass using a small number of large PEG chains.  相似文献   

20.
In this study, we show that 5α‐reductase derived from rat fresh liver was inhibited by certain aliphatic free fatty acids. The influences of chain length, unsaturation, oxidation, and esterification on the potency to inhibit 5α‐reductase activity were studied. Among the fatty acids we tested, inhibitory saturated fatty acids had C12–C16 chains, and the presence of a C?C bond enhanced the inhibitory activity. Esterification and hydroxy compounds were totally inactive. Finally, we tested the prostate cancer cell proliferation effect of free fatty acids. In keeping with the results of the 5α‐reductase assay, saturated fatty acids with a C12 chain (lauric acid) and unsaturated fatty acids (oleic acid and α‐linolenic acid) showed a proliferation inhibitory effect on lymph‐node carcinoma of the prostate (LNCaP) cells. At the same time, the testosterone‐induced prostate‐specific antigen (PSA) mRNA expression was down‐regulated. These results suggested that fatty acids with 5α‐reductase inhibitory activity block the conversion of testosterone to 5α‐dihydrotestosterone (DHT) and then inhibit the proliferation of prostate cancer cells.  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