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 15 毫秒
1.
Sliding of STOP proteins on microtubules   总被引:5,自引:0,他引:5  
M Pabion  D Job  R L Margolis 《Biochemistry》1984,23(26):6642-6648
Microtubules are stabilized against cold temperature disassembly by 145-kilodalton proteins [stable tubule only polypeptides (STOPs)] that block the end-wise dissociation of subunits from the polymers. We describe here several kinetic parameters of the interaction of STOPs with microtubules. STOPs will bind to microtubules either during assembly of the polymer or at steady state. The addition appears random on the polymers and does not require the mediation of tubulin subunits. Tubulin subunits compete with microtubules for STOP binding, but binding to the polymers is apparently irreversible. We demonstrate that STOPs do not exchange measurably between polymers at steady state. Nonetheless, a displacement of STOPs within a single polymer is readily demonstrable. We have determined that the displacement is apparently due to a surface translocation, or "sliding", of STOPs on microtubules.  相似文献   

2.
In this paper, one nanoporous graphene grafting several zwitterionic polymer chains was designed as the osmosis membrane for seawater desalination. Using molecular dynamics simulation, the efficiency and mechanism of salt rejection were discussed. The simulated results showed that the zwitterionic polymer chains on nanoporous graphene can form the charge channel to block Na+ and Cl? ions pass through, and the slat rejection efficiency of functionalised graphene can reach to about 90%. In the simulation, the steric hindrance and electrostatic interaction are the main factors for the salt rejection. With time evolution, the charge channel formed by the soft polymer chains can decrease the effective pore area of membrane, leading to the increase of steric hindrance; the positive and negative centres of polymer chains can adsorb Na+ and Cl? ions under electrostatic interaction in the solution, contributing into the increase of charge density above the membrane. These conclusions are consistent with experimental report. Our designed osmosis membrane about the graphene is helpful for improving the potential application of defect graphene in water desalination and reducing the trouble of obtaining appropriate graphene sheet with small aperture.  相似文献   

3.
The ability of several surface-active agents to stimulate the humoral immune response in mice against haptenated liposomes was tested. The surfactants were block copolymers of hydrophilic polyoxyethylene (POE) and hydrophobic polyoxypropylene (POP) that differed in m.w., percentage of POE, and mode of linkage of POP to POE. The liposomes were haptenated with tripeptide-enlarged dinitrophenyl coupled to phosphatidylethanolamine, which was incorporated into the liposomal membrane. Additional injection of mice with surfactant stimulated serum hemagglutination titers and splenic plaque-forming cell (PFC) numbers to varying extents. Block polymers with POP chains flanking a POE center, as well as polymers with POE chains flanking a POP center, displayed high adjuvant activity. These block polymers stimulated the antibody response in a dose-dependent manner. They stimulated the antibody response with both high and low antigen doses. Furthermore, the addition of one of these adjuvants (25R1) reduced the amount of carrier lipid required in the liposome in order to obtain an optimal antibody response. The surfactants, which displayed high adjuvant activity, did not interfere with liposome stability as measured with a liposome lysis assay. Moreover, in vitro preincubation of liposomes with a block polymer did not affect their immunogenicity. Optimal adjuvant activity was observed when both adjuvant and liposomes were administered by the same route. Simultaneous injection of both components, however, is not a prerequisite. Conclusively, it can be stated that nonionic block polymer surfactants are potent adjuvants for stimulation of the antibody response against haptenated liposomes.  相似文献   

4.
BACKGROUND: The theoretical state diagram for semi-flexible macromolecules such as DNA predicts that a tightly wound toroid can be a stable structure. Experimentally, toroids roughly 100 nm in diameter are routinely observed for DNA in the presence of multivalent cations at low DNA concentration. Theory also predicts toroids can form between non-DNA semi-flexible polymers and multivalent counterions. This phenomenon provides a means to co-package DNA with functionalized anionic polymers to create gene delivery systems. METHODS AND RESULTS: We show using electron microscopy that non-DNA polymers (polylysine, polyglutamic acid, and dextran sulfate) form toroids when mixed with multi- or polyvalent ions of opposite charge. The non-DNA toroids are similar in diameter to ones made with DNA. The results using dextran sulfate, a semi-flexible polymer, are explained by current theory. However, theory predicts that high flexibility in polypeptides should discourage their incorporation into stable toroids. To explain these latter observations we propose that charge neutralization facilitates secondary structure formation, which confers stiffness, thereby allowing stable toroids for the polypeptides studied. We measured the secondary structure of the toroid-forming polypeptides using circular dichroism (CD). The CD spectrum indicates the polypeptides undergo transitions from non-ordered structures (random coil) to ordered secondary structures (either alpha-helix or beta-sheet) upon charge neutralization which supports the hypothesis. The type of secondary structure is dependent on the type of multivalent counterion used to form the toroids. Formation of the polypeptide toroids confers resistance to heat denaturation of the resulting polypeptide secondary structure. The CD spectrum of DNA in a toroid also is changed from that of uncomplexed DNA, but all of the counterions used to form DNA toroids created structures with similar CD spectra in the DNA region (250-290 nm). CONCLUSIONS: The toroid structure obtained using DNA is observed in other semi-flexible non-DNA polymers such as dextran sulfate, and also in flexible polymers such as polylysine and polyglutamic acid upon charge neutralization with multivalent counterions. In the flexible polymers we propose that this phenomenon is due to induction of secondary structure upon charge neutralization, which decreases polymer flexibility, i.e. increases polymer stiffness, to enable toroid formation. These results have significant implications for the co-assembly of non-DNA anionic polymers with DNA to create nanoscopic gene carriers.  相似文献   

5.
Poly(L-lactide)-based microspheres having cationic or anionic surfaces were prepared using polydepsipeptide-block-poly(L-lactide)s as surfactants. Polydepsipeptide-block-poly(L-lactide)s having amino or carboxylic acid groups on their side chains were synthesized through anionic ring-opening polymerizations of L-lactide using the corresponding protected polydepsipeptides as macroinitiators and consequent deprotections. Since these amphiphilic copolymers consisting of hydrophobic segments and hydrophilic segments with amino or carboxylic acid groups could be converted to cationic or anionic block copolymers, they could act as surfactants preparing poly(L-lactide)-based microspheres by an oil-in-water emulsion method. The amount of ionic groups located on the surfaces of the obtained microspheres was found to increase with increasing the feed of charged polydepsipeptide-block-poly(L-lactide)s in the blend of poly(L-lactide) and block copolymers. The average diameters of the dried microspheres estimated by scanning electron microscopy were found to decrease with an increase in feed of block copolymers in polymer blends.  相似文献   

6.
The interaction between DNA and different types of amino acid-based cationic surfactants was investigated. Particular attention was directed to determine the extent of influence of surfactant head-group geometry toward tuning the interaction behavior of these surfactants with DNA. An overview is obtained by gel retardation assay, isothermal titration calorimetry, fluorescence spectroscopy, and circular dichroism at different mole ratios of surfactant/DNA; also, cell viability was assessed. The studies show that the surfactants with more complex/bulkier hydrophobic head group interact more strongly with DNA but exclude ethidium bromide less efficiently; thus, the accessibility of DNA to small molecules is preserved to a certain extent. The presence of more hydrophobic groups surrounding the positive amino charge also gave rise to a significantly lower cytotoxicity. The surfactant self-assembly pattern is quite different without and with DNA, illustrating the roles of electrostatic and steric effects in determining the effective shape of a surfactant molecule.  相似文献   

7.
Protein aggregation into oligomers and mature fibrils are associated with more than 20 diseases in humans. The interactions between cationic surfactants dodecyltrimethylammonium bromide (DTAB) and tetradecyltrimethylammonium bromide (TTAB) with varying alkyl chain lengths and bovine liver catalase (BLC) were examined by various biophysical approaches. The delicate coordination of electrostatic and hydrophobic interactions with protein, play imperative role in aggregation. In this article, we have reconnoitered the relation between charge, hydrophobicity and cationic surfactants DTAB and TTAB on BLC at pH 7.4 and 9.4 which are two and four units above pI, respectively. We have used techniques like turbidity, Rayleigh light scattering, far-UV CD, ThT, ANS, Congo red binding assay, DLS, and transmission electron microscopy. The low concentration ranges of DTAB (0–600 μM) and TTAB (0–250 μM) were observed to increase aggregation at pH 9.4. Nevertheless, at pH 7.4 only TTAB was capable of inducing aggregate. DTAB did not produce any significant change in secondary structure at pH 7.4 suggestive of the role of respective charges on surfactants and protein according to the pI and alkyl chain length. The morphology of aggregates was further determined by TEM, which proved the existence of a fibrillar structure. The surfactants interaction with BLC was primarily electrostatic as examined by ITC. Our work demystifies the critical role of charge as well as hydrophobicity in amyloid formation.  相似文献   

8.
Adsorption of the polylysine and of the copolypeptides: L-lysine/L-serine and L-lysine/L-phenylalanine on phospholipid monolayers has been investigated. The charge density of the monolayers was varied by using the negatively charged phosphatidyl serine and the neutral phosphatidyl choline at different ratios. The surface concentrations of the adsorbed polypeptides was determined by measuring the surface radiation of their radioactive label.The adsorbing capacity of the monolayer surfaces increases with their negative charge, however with respect to polypeptides the surface activity sequence is pL < pLS < pLφ. From the dependence of adsorption on the ionic strength it was concluded that it is controlled by three types of interaction: (1) electrostatic attraction to the negatively charged surface; (2) electrostatic repulsion between adsorbed polybases; (3) hydrophobic interactions involving specific structural arrangements. This is true even of the apparently neutral PC monolayer where the fixed phosphate groups form an electrical double layer with the more mobile choline groups which can be interpenetrated by the charged groups of the basic polypeptides.  相似文献   

9.
The study includes partitioning of proteins in aqueous two-phase systems consisting of the polymer dextran and the non-ionic surfactant C12E5 (pentaethylene glycol mono-n-dodecyl ether). In this system a micelle-enriched phase is in equilibrium with a polymer-enriched phase. Charges can be introduced into the micelles by the addition of charged surfactants. The charge of the mixed micelles is easily varied in sign and magnitude independently of pH, by the addition of different amounts of negatively charged surfactant, sodium dodecyl sulphate (SDS), or positively charged surfactant dodecyl trimethyl ammonium chloride (DoTAC). A series of water-soluble model proteins (BSA, β-lactoglobulin, myoglobin, cytochrome c and lysozyme), with different net charges at pH 7.1, have been partitioned in non-charged systems and in systems with charged mixed micelles or charged polymer (dextran sulphate). It is shown that partition coefficients for charged proteins in dextran-C12E5 systems can be strongly affected by addition of charged surfactants (SDS, DoTAC) or polymer (dextran sulphate) and that the effects are directly correlated to protein net charge.  相似文献   

10.
The interaction of alpha-amylase with n-alkylammonium bromides above and below their critical micellar concentrations (cmc) has been studied in buffer at pH 7 and 10 by UV spectrophotometry, photon correlation spectroscopy and Doppler microelectrophoresis. This interaction produces a complex that is dependent on pH of the medium. This complex appears at surfactant concentrations below the cmc, which means that individual surfactant molecules can bind tightly to native alpha-amylase. The complex maintains its aggregation state when the concentration of surfactants with a hydrocarbon chain of 16 carbons increases, but not for surfactants of 12 and 14 carbons. Measurements of zeta-potential indicate the influence of electrostatic and hydrophobic forces. When the size of the aggregate is maximal, proteins are at their point of zero charge. In such conditions, Van der Waals forces and contacts between the alkyl chain and the hydrophobic core of the protein favour the formation of a larger aggregate.  相似文献   

11.
The structural polypeptides of foot-and-mouth disease virus were analyzed by electrofocusing in a polyacrylamide gel containing 9 M urea. Three versions of the technique were used to accomodate the widely differing isoelectric points of the four polypeptides. VP2 was identified by comparing mature virus with procapsids. The selective actions of proteases on virions of two serotypes and on their 12S particles were examined. From this emerged a simple test for distinguishing the similarly sized polypeptides: VP1, VP2, and VP3. The effects of carbamylation and succinylation on the charge of the polypeptides were investigated. From the properties of polypeptides modified either chemically or by mutation, it was concluded that all amino acid substitutions that might be expected to cause a charge change would be detected except for neutral-to-histidine substitutions in the most basic polypeptide, VP1. In a sample of 73 temperature-sensitive mutants, 11 classes of variant polypeptides were distinguished on the basis of charge. Their molecular weights were unchanged. Alterations were found in all structural polypeptides except VP4. Mutations affecting VP2 caused similar shifts in the precursor, VP0.  相似文献   

12.
Cell surface charge and hydrophobicity of Bacillus subtilis, Escherichia coli O157:H7, Listeria monocytogenes, Salmonella typhimurium, Serratia marcescens, Staphylococcus aureus, and Staphylococcus epidermidis were determined by hydrocarbon adherence, hydrophobic interaction, and electrostatic interaction chromatography. Surface charge and hydrophobicity were compared with the initial attachment values and rates of attachment of the bacteria to meat surfaces. There was a linear correlation between the relative negative charge on the bacterial cell surface and initial attachment to lean beef muscle (r2 = 0.885) and fat tissue (r2 = 0.777). Hydrophobicity correlated well with attachment to fat tissue only. The relative hydrophobicity of each bacterium was dependent on the specific method of determination, with wide variations noted between methods.  相似文献   

13.
Cell surface charge and hydrophobicity of Bacillus subtilis, Escherichia coli O157:H7, Listeria monocytogenes, Salmonella typhimurium, Serratia marcescens, Staphylococcus aureus, and Staphylococcus epidermidis were determined by hydrocarbon adherence, hydrophobic interaction, and electrostatic interaction chromatography. Surface charge and hydrophobicity were compared with the initial attachment values and rates of attachment of the bacteria to meat surfaces. There was a linear correlation between the relative negative charge on the bacterial cell surface and initial attachment to lean beef muscle (r2 = 0.885) and fat tissue (r2 = 0.777). Hydrophobicity correlated well with attachment to fat tissue only. The relative hydrophobicity of each bacterium was dependent on the specific method of determination, with wide variations noted between methods.  相似文献   

14.
To establish correlation between structural properties (charge, composition, and conformation) and membrane penetration capability, the interaction of epitope peptide-carrier constructs with phospholipid model membranes was studied. For this we have conjugated a linear epitope peptide, (110)FWRGDLVFDFQV(121) (110-121), from VP3 capside protein of the Hepatitis A virus with polylysine-based branched polypeptides with different chemical characteristics. The epitope peptide elongated by one Cys residue at the N-terminal [C(110-121)] was attached to poly[Lys-(DL-Ala(m)()-X(i)())] (i < 1, m approximately 3), where x = ?(AK), Ser (SAK), or Glu (EAK) by the amide-thiol heterobifunctional reagent, 3-(2-pyridyldithio)propionic acid N-hydroxy-succinimide ester. The interaction of these polymer-[C(110-121)] conjugates with phospholipid monolayers and bilayers was studied using DPPC and DPPC/PG (95/5 mol/mol) mixture. Changes in the fluidity of liposomes induced by these conjugates were detected by using two fluorescent probes 1,6-diphenyl-1,3, 5-hexatriene (DPH) and sodium anilino naphthalene sulfonate (ANS). The binding of conjugates to the model membranes was compared and the contribution of the polymer component to these interactions were evaluated. We found that conjugates with polyanionic/EAK-[C(110-121)] or polycationic/SAK-[C(110-121)], AK-[C(110-121)]/character were capable to form monomolecular layers at the air/water interface with structure dependent stability in the following order: EAK-[C(110-121)] > SAK-[C(110-121)] > AK-[C(110-121)]. Data obtained from penetration studies into phospholipid monolayers indicated that conjugate insertion is more pronounced for EAK-[C(110-121)] than for AK-[C(110-121)] or SAK-[C(110-121)]. Changes in the fluorescence intensity and in polarization of fluorescent probes either at the polar surface (ANS) or within the hydrophobic core (DPH) of the DPPC/PG liposomes suggested that all three conjugates interact with the outer surface of the bilayer. Marked penetration was documented by a significant increase of the transition temperature only with the polyanionic compound/EAK-[C(110-121)]. Taken together, we found that the binding/penetration of conjugates to phospholipid model membranes is dependent on the charge properties of the constructs. Considering that the orientation and number of VP3 epitope peptides attached to branched polypeptides were almost identical, we can conclude that the structural characteristics (amino acid composition, charge, and surface activity) of the carrier have a pronounced effect on the conjugate-phospholipid membrane interaction. These observations suggest that the selection of polymer carrier for epitope attachment might significantly influence the membrane activity of the conjugate and provide guidelines for adequate presentation of immunogenic peptides to the cells.  相似文献   

15.
Using X-ray diffraction we have studied fibres obtained from complexes of DNA with lysine-rich polypeptides and with proteins that have different conformations, to ascertain whether the conformations of the polypeptides and the DNA are maintained upon interaction. Substances investigated include N-acetyl-Lys-Ala-Tyr-Ala-Lys-ethylamide, random poly(Leu50, Lys50), sequential poly(Leu-Lys), poly(Val-Lys), poly(Ala-Lys), poly(Lys-Ala-Ala-Lys), poly(Lys-Ala-Ala), poly(Lys-Leu-Ala), poly(Lys-Ala-Gly), protein phi 0 from sea cucumber spermatozoa, histone H1 and two fragments of this protein obtained by chemical cleavage. In general, the B form of DNA with ten base-pairs per helical turn is maintained upon interaction at high levels of humidity. The A form is never observed; it appears to be forbidden in a protein environment. No evidence for transition into any novel DNA conformation has been observed, although the B form is altered in some cases, in particular upon dehydration. Such alteration occurs always in the sense of tightening the double helix, so that the number of base-pairs per helical turn diminishes. The polypeptides may interact with DNA in both the alpha and beta conformations. We have found different types of complexes in which either a monolayer or a double layer of beta-pleated sheets is intercalated between layers of DNA molecules. Alternatively, the polypeptide chain may be wrapped around the DNA, following one of the grooves. The polypeptide conformation may be either maintained or changed upon interaction. The charge density of the polypeptide is an important parameter of the interaction. When it matches the charge density of the DNA, the polypeptide conformation is maintained in most cases; otherwise it is modified. The globular part of histone H1 gives a unique X-ray pattern upon interaction, indicative of a loss of order of DNA in the complex. On the other hand, the C-terminal part of histone H1 gives a very well-ordered complex, similar to a nucleoprotamine, in spite of its lower charge density.  相似文献   

16.
The cooperative nature of interaction of cationic surfactants with short oligonucleotides leading to eventual stabilization of DNA duplexes is demonstrated. At submicellar concentrations and DNA:surfactant charge ratios of 0.2 to 0.8, the association of single chain (CTAB) and double chain (DOTAP) surfactants to oligonucleotides is initiated by electrostatic interaction of cationic ligands with polyanionic DNA that aligns the surfactant molecules on the DNA template. This is followed by binding of new surfactant ligands to the initial complex, driven cooperatively by the hydrophobic forces, leading to in situ formation of surfactant-bound and bare duplexes as separate species. These exhibit independent melting behaviour characterised by double transition in thermal UV profiles, with a higher T(m) for surfactant-DNA complexes. Understanding the cooperative binding of the cationic surfactants to the DNA described here may have implications for rational design of DNA binding drugs and DNA delivery systems.  相似文献   

17.
Branched polypeptides with polylysine backbone are promising candidates for selective delivery of drugs, epitopes. or reporter molecules. We reported earlier that polylysine-based polypeptides with polyanionic character were internalized by murine bone marrow derived macrophages via class A scavenger receptor. In the present studies, our investigations were extended to seven polypeptides with different amino acid composition and charge properties. We report on our findings on the concentration-dependent influence of these compounds on survival and chemotaxis of the murine macrophage-like cell line J774 and internalization properties of the polypeptides by J774 cells. Our observations indicate that the polypeptides regardless of their charge properties were essentially nontoxic and did not alter significantly the chemotaxis of J774 cells; therefore, the polypeptides suit the requirements for nontoxic and "neutral" carrier molecules. We also demonstrated that the polypeptides were internalized efficiently by J774 cells, depending on their chemical structure and charge properties. Using the scavenger receptor-ligand fucoidan as inhibitor, we established that the scavenger receptor played a role-in accordance with findings on murine bone marrow derived macrophages in the internalization only of the polyanionic polypeptides.  相似文献   

18.
We studied the interaction of poly‐l ‐lysine (PLL) and poly‐l ‐arginine (PLAG) with sodium dodecyl sulfate (SDS) surfactant and the interaction of poly‐l‐ glutamic acid (PLGA) and poly‐l ‐aspartic acid (PLAA) with tetradecyltrimethylammonium bromide (TTAB) surfactant using vibrational circular dichroism (VCD) spectroscopy in the region of C‐H stretching vibration and in the Amide I region both in solution and in mulls. A chirality transfer from polypeptides to achiral surfactants was observed in the C‐H stretching region, where measurements in solution were impossible. This observation was enabled by a special sample treatment technique using lyophilization and the preparation of mulls. This technique demonstrated itself as an interesting and beneficial tool for VCD measurements. In addition, we observed that SDS changed the secondary structure of PLL to the β‐sheet and of PLAG to the α‐helix. TTAB disrupted the PLGA and PLAA structure. These results were obtained in the mull but were confirmed by the VCD spectra measured in solution and by electronic circular dichroism. The chirality transfer from the polypeptides to SDS was caused by polypeptides ordered into a specific conformation during the interaction, while in the TTBA system it was induced primarily by the chirality of the amino acid residues. Chirality 27:965–972, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

19.
The effect of surfactants on membrane-attached biofilms (MABs) was studied in a lab-scale extractive membrane bioreactor (EMB). Twenty-two surfactants were screened for their potential of increasing the cell wall negative charge (i.e. the electrostatic repulsion between bacteria) of Burkholderia sp. JS150 bacterial strain. Surfactants resulting in increased bacterial negative charge were further investigated for their effects on MAB population morphology and MAB attachment behaviour. Microscopic investigation of the bacterial population in MABs showed that surfactants affect the development of flagella, suggesting changes in the attachment capability of the JS150 strain in the presence of different surfactants. Among the screened surfactants, teepol showed the best characteristics in relation to the reduction of MAB accumulation, and it was tested in an EMB system for the extraction of monochlorobenzene from a synthetic wastewater. Comparison with a control EMB, operated without surfactants under the same conditions, proved that teepol effectively reduces MAB accumulation on the membrane walls. As a result, the overall mass transfer coefficient in the presence of teepol was 53% higher than in the control EMB.  相似文献   

20.
This investigation aimed to develop nimesulide (NMS)-loaded poly(lactic-co-glycolic acid) (PLGA)-based nanoparticulate formulations as a biodegradable polymeric drug carrier to treat rheumatoid arthritis. Polymeric nanoparticles (NPs) were prepared with two different nonionic surfactants, vitamin E d-α-tocopheryl polyethylene glycol 1000 succinate (vitamin E TPGS) and poly(vinyl alcohol) (PVA), using an ultrasonication solvent evaporation technique. Nine batches were formulated for each surfactant using a 32 factorial design for optimal concentration of the emulsifying agents, 0.03–0.09% for vitamin E TPGS and 2–4% for PVA. The surfactant percentage and the drug/polymer ratio (1:10, 1:15, 1:20) of the NMS-loaded NPs were investigated based on four responses: encapsulation efficiency, particle size, the polydispersity index, and the surface charge. The response surface plots and linearity curves indicated a relationship between the experiment’s responses and a set of independent variables. The NPs produced with both surfactants exhibited a negative surface charge, and scanning electron micrographs revealed that all of the NPs were spherical in shape. A narrower size distribution and higher drug loadings were achieved in PVA-emulsified PLGA NPs than in the vitamin E TPGS emulsified. Decreasing amounts of both nonionic surfactants resulted in a reduction in the emulsion’s viscosity, which led to a decrease in the particle size of NPs. According to the ANOVA results obtained in this present research, vitamin E TPGS exhibited the best correlation between the independent variables, namely drug/polymer ratio and the surfactant percentage, and the dependent variables (encapsulation efficiency R2 = 0.9603, particle size R2 = 0.9965, size distribution R2 = 0.9899, and surface charge R2 = 0.8969) compared with PVA.KEY WORDS: ANOVA, factorial design, nanoparticles, nimesulide, PLGA, PVA, vitamin E TPGS  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