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.
Dynamic light scattering (DLS) measurements were performed to study the binding of anionic surfactant alpha olefin sulfonate (AOS) to gelatin chains at various NaCl concentrations at 30 degrees C in aqueous sodium phosphate buffer (pH = 6.8) solutions. The surfactant concentration was varied from 0 to 80 mM and the NaCl concentrations chosen were 0.025, 0.05, and 0.1 M. AOS exhibited electrostatic binding to the positively charged sites of the polypeptide chain resulting in considerable reduction in its hydrodynamic radius up to critical micellar concentration (cmc = 8 mM for no salt, 0.01 and 0.025 M, and 5 mM for 0.05 M and 2 mM for 0.1 M solutions). The correlation function revealed the presence of two types of structures above cmc; namely the micelles of AOS and gelatin-AOS micelle complexes. The micellar radii (Rm), the effective gelatin-surfactant complex radii (Rc), have been determined as a function of salt concentration. No critical aggregation concentration (cac) was observed. The inter-gelatin-surfactant complex (kD1) and inter-micellar interactions (kD2), were determined by fitting the concentration dependence of Rm and Rc to a virial expansion in reduced concentration (c - cmc), which are compared. While kD1 showed strong ionic strength dependence, kD2 remained invariant of the same. The protein to surfactant binding ratio was found to be smaller than normal. Results have been discussed within the framework of the necklace-bead model of polymer-surfactant interactions.  相似文献   

2.
The binding of the fluorescent probes 1-anilino-8-naphthalene sulfonate and dansyl cadaverine to the sodium salts of cholic, deoxycholic and dehydrocholic acids has been investigated. Enhanced probe solubilisation accompanies aggregation. Monitoring of fluorescence intensities as a function of bile salt concentration permits the detection of primary micelle formation, as well as secondary association. The transition concentrations obtained by fluorescence are in good agreement with values determined for the critical micelle concentrations, by other methods. Differences in the behaviour of cholate and deoxycholate have been noted. Fluorescence polarisation studies of 1,6-diphenyl-1,3,5-hexatriene solubilised in bile salt micelles suggest a higher microviscosity for the interior of the deoxycholate micelle as compared to cholate. 1H NMR studies of deoxycholate over the range 1–100 mg/ml suggest that micelle formation leads to a greater immobilisation of the C18 and C19 methyl groups as compared to the C21 methyl group. Well resolved 13C resonances are observed for all three steroids even at high concentration. Both fluorescence and NMR studies confirm that dehydrocholate does not aggregate.  相似文献   

3.
We have investigated the conformational transition and aggregation process of recombinant Syrian hamster prion protein (SHaPrP90-232) by Fourier transform infrared spectroscopy, circular dichroism spectroscopy, light scattering, and electron microscopy under equilibrium and kinetic conditions. SHaPrP90-232 showed an infrared absorbance spectrum typical of proteins with a predominant alpha-helical structure both at pH 7.0 and at pH 4.2 in the absence of guanidine hydrochloride. At pH 4.2 and destabilizing conditions (0.3-2 m guanidine hydrochloride), the secondary structure of SHaPrP90-232 was transformed to a strongly hydrogen-bonded, most probably intermolecularly arranged antiparallel beta-sheet structure as indicated by dominant amide I band components at 1620 and 1691 cm-1. Kinetic analysis of the transition process showed that the decrease in alpha-helical structures and the increase in beta-sheet structures occurred concomitantly according to a bimolecular reaction. However, the concentration dependence of the corresponding rate constant pointed to an apparent third order reaction. No beta-sheet structure was formed within the dead time (190 ms) of the infrared experiments. Light scattering measurements revealed that the structural transition of SHaPrP90-232 was accompanied by formation of oligomers, whose size was linearly dependent on protein concentration. Extrapolation to zero protein concentration yielded octamers as the smallest oligomers, which are considered as "critical oligomers." The small oligomers showed spherical and annular shapes in electron micrographs. Critical oligomers seem to play a key role during the transition and aggregation process of SHaPrP90-232. A new model for the structural transition and aggregation process of the prion protein is described.  相似文献   

4.
Micellization of sodium deoxycholate (NaDC) and sodium ursodeoxycholate (NaUDC) was studied for the critical micelle concentration (CMC), the micelle aggregation number, and the degree of counterion binding to micelle, where sodium cholate (NaC) was used as a reference. The fluorescence probe technique of pyrene was employed to determine accurately the CMC values for the bile salts, which indicated that a certain concentration range of CMC and a stepwise aggregation for micellization were reasonable. The temperature dependences of micellization for NaDC and NaUDC were studied at 288.2, 298.2, 308.2, and 318.2 K by aqueous solubility change with solution pH. Aggregations of the bile salt anions were analyzed using the stepwise association model and found to grow in size with increasing concentration, which confirmed that the mass action model worked quite well. The average aggregation number was found to be 2.5 (NaUDC) and 10.5 (NaDC) at the concentration of 20 mM and at 308.2 K. The aggregation number determined by static light scattering also agreed well with those by the solubility method in the order of size: NaUDC相似文献   

5.
The effect of low concentrations of nonionic detergents with different critical micelle concentrations such as Triton X-100, Brij 35 and octylglucoside on rabbit liver microsomes is studied by means of 31P-NMR, 1H-NMR, dynamic light scattering and functional investigations. Hexane phosphonic acid diethyl ester was used as a phosphorus membrane probe molecule to monitor the interaction of detergent molecules with microsomal phospholipids by 31P-NMR. This method is more sensitive than 31P-NMR of phospholipids alone and permitted the estimation of the maximum number of detergent molecules which can be incorporated in microsomes without the formation of mixed micelles outside the membrane. These membrane saturation concentrations were determined to be 0.07 (Brij 35), 0.1 (Triton X-100) and 0.4 (octylglucoside) (molar ratio of detergent/total phospholipids). Above these detergent concentrations, mixed micelles consisting of detergent and membrane constituents are formed, coexisting with the microsomes up to the membrane solubilization concentration. The results indicate a dependence of the membrane saturation concentration on the critical micelle concentration of the detergent and a preferential removal of phosphatidylcholine over phosphatidylethanolamine from the microsomes by all detergents studied.  相似文献   

6.
Human islet amyloid polypeptide (hIAPP) is the major component of amyloid plaques found in the pancreatic islets of persons with type 2 diabetes mellitus. HIAPP belongs to the group of amyloidogenic proteins, characterized by their aggregation and deposition as fibrillar amyloid in various body tissues. The aggregation of amyloidogenic proteins is thought to occur via a common pathway, but currently no unifying kinetic model exists. In previous work, we presented a model of amyloid fibril formation formulated from our observations of the aggregation of an amyloidogenic fragment of hIAPP, amino acids 20-29. Our model is based on nucleation-dependent aggregation, modified by the formation of off-pathway hIAPP micelles. In the present study we confirm the presence of peptide micelles, and experimentally determine the critical micelle concentration in solutions of hIAPP fragments using three different techniques: conductivity, pH, and fluorescence. All three techniques yield a critical micelle concentration of 3-3.5 micro M peptide. Furthermore, based on changes in the fluorescence intensity of a labeled peptide fragment as well as a decrease in solution pH as a result of deprotonation of the amino terminus, we conclude that the amino terminus of the fragment undergoes a significant change of environment upon micellization.  相似文献   

7.
The interactions between the diblock copolymer S(15)E(63) and the surfactants sodium dodecyl sulfate (SDS), sodium decyl sulfate (SDeS), and sodium octyl sulfate (SOS) have been investigated by dynamic light scattering (DLS), transmission electron microscopy (TEM), and isothermal titration calorimetry (ITC). The surfactants with the same headgroup differentiate in their chain length. At 20 degrees C, the block copolymer is associated into micelles with a hydrodynamic radius of 11.6 nm, which is composed of a hydrophobic styrene oxide (S) core and a water-swollen oxypolyethylene (PEO or E) corona. The different copolymer/surfactant systems have been studied at a constant copolymer concentration of 2.5 g dm(-3) and in a vast range of surfactant concentrations, from 7.5 x 10(-6) up to 0.75 M. When SDS and SDeS are added to the block copolymer solution, different regions are observed in the DLS data: at low surfactant concentrations (c < 1.0 x 10(-4) M), single surfactant molecules associate with the copolymer micelle, probably the former being solubilized in the micelle core, leading to a certain disruption of the mixed micelle due to repulsive electrostatic interactions between surfactant headgroups followed by a stabilization of the mixed micelle. At higher concentrations (1.0 x 10(-4) < c < 0.1 M), two types of copolymer-surfactant complexes coexist: one large copolymer-rich/surfactant complex and one small complex consisting of one or a few copolymer chains and rich in surfactants. At higher SDS and SDeS concentrations, complete disintegration of mixed micelles takes place. In contrast, SOS-S(15)E(63) interactions are less important up to surfactant concentrations of 0.05 M due to its higher hydrophilicity, reducing the hydrophobic interactions between surfactant alkyl chains and copolymer micelles. At concentration larger than the critical aggregation concentration (cac) of the system, 0.05 M, disruption of copolymer micelles occurs. These regions have been confirmed by transmission electron microscopy. On the other hand, the titration calorimetric data for SDS and SDeS present an endothermic increase indicating the formation of mixed copolymer-rich-surfactant micelles. From that point, important differences in the ITC plot for both surfactants are present. However, the ITC curve obtained after titration of a SOS solution in the copolymer solution is quite similar to that of its titration in water.  相似文献   

8.
Micelle formations of sodium glyco- and taurochenodeoxycholate (NaGCDC and NaTCDC) and sodium glyco- and tauroursodeoxycholates (NaGUDC and NaTUDC) was studied at 308.2 K for their critical micelle concentrations at various NaCl concentrations by pyrene fluorescence probe, and the degree of counterion binding to micelle was determined using the Corrin-Harkins plots. The degree of counterion binding was found to be 0.37-0.38 for chenodeoxycholate conjugates, while the determination of the degree was quite difficult for ursodeoxycholate conjugates. The change of I1/I3 values on the fluorescence spectrum with the conjugate bile salt concentration suggested two steps for their bile salt aggregation. The first step is a commencement of smaller aggregates, the first cmc, and the second one is a starting of stable aggregates, the second cmc. The aggregation number was determined at 308.2 K and 0.15 M NaCl concentration by static light scattering: 16.3 and 11.9 for sodium NaGCDC and NaTCDC, and 7.9 and 7.1 for NaGUDC and NaTUDC, respectively. The solubilization of cholesterol into the bile salt micelles in the presence of coexisting cholesterol phase and the maximum additive concentration (MAC) of cholesterol was determined against the bile salt concentration. The standard Gibbs energy change for the solubilization was evaluated, where the micelles were regarded as a chemical species. The solubilization was stabilized in the order of NaGUDC approximately = NaTUDC < NaTC < NaGC < NaTCDC < NaGCDC < NaTDC < NaGDC, where the preceding results were taken into the order.  相似文献   

9.
A method has been developed to obtain micelles based on amphiphilic chitosan derivatives which were synthesized by grafting hydrophobic stearoyl, palmitoyl and octanoyl aliphatic chains onto molecules of chitosan with degrees of substitution from 0.9% to 29.6%. The N-fatty acylations were carried out by reacting carboxylic anhydride with chitosan in dimethyl sulfoxide. The chitosan derivative-based micelles were spherical as observed by transmission electron microscope (TEM). Their sizes were in the range of 140–278 nm as measured by dynamic light scattering (DLS). The micellar critical aggregation concentration (CAC) can reach 1.99 × 10−3 mg/mL, indicating that they are more stable upon dilution than micelles based on other chitosan derivatives such as deoxycholic acid-modified chitosan reported previously.  相似文献   

10.
Rao J  Luo Z  Ge Z  Liu H  Liu S 《Biomacromolecules》2007,8(12):3871-3878
A polypeptide hybrid double hydrophilic diblock copolymer (DHBC), poly( N-isopropylacrylamide)- b-poly( l-glutamic acid) (PNIPAM- b-PLGA), was synthesized via the ring-opening polymerization of gamma-benzyl- l-glutamate N-carboxyanhydride (BLG-NCA) using monoamino-terminated PNIPAM as the macroinitiator, followed by deprotection of benzyl groups under alkaline conditions. Containing a thermoresponsive PNIPAM block and a pH-responsive PLGA block, the obtained polypeptide hybrid diblock copolymer molecularly dissolves in aqueous solution at alkaline pH and room temperature but supramolecularly self-assembles into PNIPAM-core micelles at alkaline pH and elevated temperatures and PLGA-core micelles at acidic pH and room temperature accompanied with coil-to-helix transition of the PLGA sequence. The pH- and thermoresponsive "schizophrenic" micellization behavior of PNIPAM- b-PLGA diblock copolymer has been investigated by (1)H NMR, optical transmittance, fluorescence probe measurement, transmission electron microscopy (TEM), dynamic and static laser light scattering (LLS), and circular dichroism (CD) spectroscopy. Moreover, the micellization process was investigated employing stopped-flow light scattering technique. The pH-induced micelle growth of PNIPAM- b-PLGA in aqueous solution exhibits drastically different kinetics compared to that of conventional pH-responsive DHBCs, probably due to the stabilization effects exerted by the formed alpha-helix secondary structures within the PLGA core at low pH. Exhibiting "schizophrenic" micellization, the polypeptide sequence of PNIPAM- b-PLGA can either locate within micelle cores or stabilizing coronas. The incorporation of polypeptide block into DHBCs can endow them with structural versatility, tunable spatial arrangement of chain segments within self-assembled nanostructures, and broader applications in the field of biomedicines.  相似文献   

11.
The aggregation of poly(γ-benzyl-α,L -glutamate) and its enantiomer in toluene has been investigated by following the viscosity as a function of temperature, concentration, molecular weight, molecular-weight distribution, helix chirality, and shear rate. The temperature and concentration data for a 138,000-molecular-weight sample was fitted to an open, reversible end-to-end aggregation model. The aggregation numbers resulting from this fit were consistent with the sudden onset in non-Newtonian flow resulting from only a 0.2-wt% increase in concentration. The association equilibrium constant was then used to predict viscosity for comparison with other data, in particular, the effect of molecular weight and molecular-weight distribution. A mixture of right-and left-handed helices showed the aggregation was not chiral selective. The stiffness of end-to-end aggregated (hydrogen-bonded) molecules differed little from their covalent counterparts, at least below a molecular weight of ~106. We conclude that polybenzylglutamate aggregation in toluene can be described by an open end-to-end aggregation model.  相似文献   

12.
The critical micelle concentrations of aqueous solutions of Nα-acyl-L-histidine have been determined by the spectral shift method with Rhodamine 6G and by the light scattering method. With the spectral shift method critical micelle concentrations of 40, 9.0, 1.0, 0.11, and 0.012 mM were obtained for Nα-acyl-L-histidine containing saturated acids of 8, 10, 12, 14, and 16 carbons respectively, at 45°C and pH 8.6 in the absence of added salt. For the homologs containing 10, 12, and 14 carbon acids, critical micelle concentration of 9.0, 1.0, and 0.11 mM were determined by the light scattering method.The light scattering studies yield micelle weights of 60, 66, and 84 thousand for the C-10, C-12, and C-14 homologs, respectively.Nα-acyl-L-histidine is an unusual surfactant in that the hydrophilic portion of the molecule is relatively large and contains both an ionic group (carboxylate group) and a nonionic group (imidazole side-chain). The bulky hydrophilic group of Nα-acyl-L-histidine causes this molecule to exhibit physico-chemical behavior which is not typical of that exhibited by most ionic surfactants. In particular, the dependence of the critical micelle concentration on the acyl chain length and on the concentration of added salt is atypical.Chemical shift measurements (by NMR) on the C-2 and C-5 protons of imidazole in micellar Nα-dodecanoyl-L-histidine indicate that the imidazole group is, indeed, positioned at the water-micelle interface.  相似文献   

13.
Earlier neutron small-angle scattering experiments had revealed the low resolution structure of the complex between sodium dodecyl sulfate (SDS) and the single polypeptide (452 amino acid residues) of a water-soluble enzyme. The saturated complex consists of three globular micelles which are connected by short flexible polypeptide segments. New experiments, described here, were performed at subsaturating concentrations of free SDS in equilibrium with the complex. The data show a decrease in stoichiometry from one bound dodecyl sulfate (DS) anion per two amino acid residues near the critical micelle concentration (CMC) to one per four residues at half the CMC. At 0.3 CMC, a two-micelle complex is formed by the recombination of the small amino-terminal micelle with the middle one; and the center-to-center distance between the carboxyl-terminal micelle and the middle one decreases from 7.5 to 6.2 nm. These structural data allow us to better understand earlier results obtained with high-performance agarose gel chromatography of the same SDS-protein complexes.  相似文献   

14.
Assignments have been obtained for most of the 1H-NMR lines of melittin bound to fully deuterated dodecylphosphocholine micelles by combined use of two-dimensional spin echo correlated spectroscopy and one-dimensional NMR methods. Nuclear Overhauser enhancement measurements showed that the mobility of the entire polypeptide chain is reduced by binding of melittin to the detergent micelle and that the amino-terminal and carboxy-terminal halves of the primary structure constitute separate, compact domains within the conformation of micelle-bound melittin. p2H titration experiments showed that the presence of positive charges on the four amino groups of melittin had little influence on the conformation of the micelle-bound polypeptide. Titration of tetrameric melittin with detergent provided evidence that melittin assumes similar conformations as a self-aggregated tetramer and as a monomer bound to micelles.  相似文献   

15.
The spin labels, 5-doxylstearate, 12-doxylstearate, 16-doxylstearate and 1-oxyl-2,2,6,6-tetramethyl-4-dodecylphospiperidine, have been incorporated into dodecylphospocholine micelles and mixed dodecylphosphocholine/ glucagon micelles. The EPR spectral parameters for the different spin labels and the 1H- and 13C-NMR relaxation rates for nuclei of the detergent molecules indicated that inclusion of up to one spin label molecule per micelle had little influence on the spatial organization of the micelles. Furthermore, the location and environment of the spin labels in the dodecylphosphocholine micelles were not noticeably affected by the addition of glucagon and the 1H-NMR spectra observed for glucagon in mixed spin label/deuterated dodecylphosphocholine/glucagon micelles showed that the different spin labels had essentially no effect on the conformation of glucagon. Approximate spatial locations within the micelle for the nitroxide moieties of the different spin labels were determined from the NMR relaxation rates observed for different nuclei of dodecylphosphocholine. On this basis, the line broadening of individually assigned glucagon 1H-NMR lines by the different spin labels was used to determine the approximate orientation of the polypeptide chain with respect to the micelle surface. Overall, the data indicate that the glucagon backbone runs roughly parallel to the micelle surface, with the depth of immersion adjusted so that polar and apolar side chains can be oriented towards the surface or interior of the micelle, respectively.  相似文献   

16.
The structure of the complex between sodium dodecyl sulfate (SDS) and a deuterated bifunctional enzyme, N-5'-phosphoribosylanthranilate isomerase/indole-3-glycerol-phosphate synthase (Mr 49,484), has been studied in dilute solution by small-angle neutron scattering. The complex nearly acquired its final size, as shown by molecular-sieve chromatography, at the chosen SDS concentration of 1.6 mM, which is slightly below the critical micelle concentration of 1.8 mM (at the ionic strength of 0.1 M). The 452 amino-acid residues of the bifunctional enzyme were combined with 216 detergent molecules. The complex was found to be composed of three protein-decorated SDS micelles of unequal size, connected by short flexible polypeptide segments. The largest of the three micelles was the middle one. The SDS-protein complex contained the dodecyl hydrocarbon moieties in three globular cores. Each core was surrounded by a hydrophilic shell, formed by the hydrophilic and amphiphilic stretches of the polypeptide chain, and by the sulfate head groups of the detergent. The average thickness of these shells was 0.7-0.8 nm. The three-micelle complex was cleaved with trypsin at a single site, possibly in a micelle-connecting segment, into a single-micelle fragment at the carboxyl-terminal which comprised 73 SDS molecules and 163 amino-acid residues, and a dual-micelle fragment. One of the micelles within this larger fragment contained 42 SDS molecules and about 90 amino-acid residues; the other micelle contained 101 SDS molecules and about 190 amino-acid residues. The individual micelle sizes seemed to be determined by the amino-acid sequence.  相似文献   

17.
Interaction of sodium dodecyl sulfate (SDS) with a globular protein, trypsin, has been physicochemically studied in aqueous medium in detail using tensiometric, conductometric, calorimetric, fluorimetric, viscometric, and circular dichroism techniques. The results indicate that SDS-trypsin aggregates start to form at a surfactant concentration higher than the critical micelle concentration of pure SDS micelle. In contrast, the counterion binding decreases in the presence of trypsin. The free energies and enthalpies of micellization, interfacial adsorption, and entropy of micellization associated with the interaction have also been calculated. The values show that the interaction phenomenon is entropy controlled and endothermic in nature. The increase in viscosity is observed for the system of SDS-trypsin cluster above the critical micelle concentration of SDS micelle only. The aggregation number and interface polarity decrease compared to the values of micelles without protein. Circular dichroism spectra show the high alpha-helical content and unfolded structure of trypsin in the presence of SDS due to strong electrostatic repulsion leading to a probable "necklace and bead" model in the case of biopolymer-surfactant complexes.  相似文献   

18.
Dispersed from a dry film into buffer (5 mM phosphate, 0.15 M NaCl, pH 7.4), the liponucleotide 1-β-d-arabinofuranosylcytosine 5′-diphosphate l-1,2-diacylglycerol (ara-CDPdiacylglycerol) spontaneously forms vesicles which are several microns in diameter and probably unilamellar. Their average size immediately begins to decrease, and after 2 h none can be seen in the light microscope. During 1–2 days in unstirred solutions at 25°C, the vesicles are transformed to spherical or nearly spherical micelles having an apparent partial specific volume of 0.835 ml·g?1, a maximum possible aggregation number of about 150, and an anhydrous radius of about 37 Å. The critical micelle concentration (CMC) is about 10 μM in buffer and 20 μM in distilled water, but micelle-monomer equilibration requires at least 1 week at a total concentration of 66 μM. This exceedingly slow equilibration is unique among reported detergents. The standard enthalpy and entropy of micellization are ?13 kJ·mol?1 and 87 J·mol?1·K?1, respectively. These values are within the range reported for other detergents. Sonication accelerates the vesicle-micelle transformation to 30 min.  相似文献   

19.
Reduced viscosity and infrared spectra of low-molecular-weight poly-γ-benzyl-L -glutamate (which was prepared by polymerization of the N-carboxyanhydride with n-hexylamine initiation at [A]/[I] 3, 4, and 8) have been measured in various organic solvents. Infrared spectra indicate that the polypeptide molecules consist of a series of residues of two forms, the solvated σ-form and the hydrogen-bonded β-form, and relative abundance of the two forms depends on solvent species and polypeptide concentration. An approximate method is developed for estimating the content of β-structure from a single spectrum of dissolved polypeptide. The reduced viscosity of some solutions is scarcely dependent in polypeptide concentration, in which a single conformation is predominantly kept over the concentration range. In the other solutions the reduced viscosity displays a strong concentration dependence or some anomalous behavior. The observed viscosity behavior has been attributed to the changes in size and shape of aggregates, which are determined by the number of hydrogen bonds in the aggregate. This unusual behavior is exhibited by solutions of the polypeptides which have a moderate content of β-structure at a finite concentration. Both the content of β-structure and the extent of association increase in the following solvents, ranked in order of effectiveness: dimethylformamide, trifluoroethanol < trimethyl phosphate < chloroform < dioxane < ethylene dichloride < ethylene dibromide. Infrared spectra suggest that the conformation of the polypeptide in dichloroacetic acid differs from either the σ- or the β-conformation.  相似文献   

20.
Biotinylated and non-biotinylated copolymers of poly(ethylene oxide) (PEO) and poly(2-(diethylamino)ethyl methacrylate) (PDEAEMA) were synthesized by the atom transfer radical polymerization technique. The chemical compositions of the copolymers as determined by NMR are represented by PEO(113)PDEAEMA(70) and biotin-PEO(104)PDEAEMA(93), respectively. The aggregation behavior of these polymers in aqueous solutions at different pHs and ionic strengths was studied using a combination of potentiometric titration, dynamic light scattering, static light scattering, and transmission electron microscopy. Both PEO-b-PDEAEMA and biotin-PEO-b-PDEAEMA diblock copolymers form micelles at high pH with hydrodynamic radii (R(h)) of about 19 and 23 nm, respectively. At low pH, the copolymers are dispersed as unimers in solution with R(h) values of about 6-7 nm. However, at a physiological salt concentration (c(s)) of about 0.16 M NaCl and a pH of 7-8, the copolymers form large loosely packed Gaussian chains, which were not present at the low c(s) of 0.001 M NaCl. The critical micelle concentrations (cmc's) and the cytotoxicities of the copolymers were investigated to determine a suitable polymer concentration range for future biological applications. Both PEO-b-PDEAEMA and biotin-PEO-b-PDEAEMA diblock copolymers possess identical cmc values of about 0.0023 mg/g, while the cytotoxicity test indicated that the copolymers are not toxic up to 0.05 mg/g (>83% cell survival at this concentration).  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