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 0 毫秒
1.
The temperature dependencies of the infrared spectra of pure and cholesterol-containing multibilayers of 1,2-dipalmitoyl-sn-glycero-3-phosphocholine were studied using Fourier transform infrared techniques. A comparison of the spectroscopic data showed the retention of a melting phenomenon at 60 mol% cholesterol content, and the retention of some all-trans conformations in the liquid-crystalline phase. It is also demonstrated that at temperatures less than 30 degrees C, the cholesterol-containing 1,2-dipalmitoyl-sn-glycero-3-phosphocholine multibilayers still contain a small amount of pure 1,2-dipalmitoyl-sn-glycero-3-phosphocholine, packed in an orthorhombic subcell lattice. Spectral changes were found in the absorptions characteristic of the phospholipid head groups. The addition of cholesterol results in changes in the ester bands, and demonstrates the induction by cholesterol of non-equivalent ester conformations.  相似文献   

2.
Enkephalins are pentapeptides found in the human nervous system, where they are involved in the relief of pain. The interaction of these neuropeptides with the nerve cell membranes would be a key-step in the receptor binding. We have used both Fourier-transform infrared and solid-state NMR spectroscopies to shed light on the interactions responsible for the association of enkephalins with negatively charged membranes. More specifically, we have investigated the interaction of methionine-enkephalin (Menk) with DMPG and DMPS vesicles. Our results suggest that Menk interacts electrostatically with both model membranes via its terminal NH3+ group. However, the peptide induced the formation of elongated DMPG vesicles in the magnetic field. On the other hand, the association of Menk with DMPS bilayers was concentration-dependent and disrupted the membrane at high peptide concentrations. The different effect of methionine-enkephalin with the two types of anionic membranes is most likely related to the different fluidity of these systems.  相似文献   

3.
Structural studies of retinochrome, and its photoproduct, lumiretinochrome, were done by Fourier transform infrared difference spectroscopy. The absorption bands in the carbonyl stretching region which shift in D2O show the changes in the protein part during the photoreaction. Strong absorption bands in the finger-print region show that the all-trans-retinal chromophore in retinochrome isomerizes to the 11-cis-retinal chromophore in lumiretinochrome upon illumination with yellow-green light at 83K.  相似文献   

4.
Ca2+-ATPase from rabbit sarcoplasmic reticulum has been isolated, purified, and reconstituted into lipid environments containing as primary components 1,2-dielaidoylphosphatidylcholine (DEPC) and acyl-chain perdeuterated 1,2-dimyristoylphosphatidylcholine (DMPC-d54). Differential scanning calorimetry (DSC) has been used to elucidate the phase behavior of this lipid pair while Fourier transform infrared spectroscopy (FT-IR) has been used to monitor the state of each lipid component in the presence of protein. The lipid mixture shows gel state miscibility over at least most of the composition range, a result in good accord with Van Dijck et al. (Biochim. Biophys. Acta 470, 58-69 (1977)), for the binary mixture with proteated DMPC. Acyl chain perdeuteration thus does not greatly alter the miscibility properties of the lipid pair. Reconstitution of Ca2+-ATPase with this lipid pair proceeds with moderate efficiency. Up to 80% of the endogenous lipid can be replaced depending on the lipid composition. Unusual composition-dependent protein-induced effects on lipid melting properties are noticed. At low levels of DMPC-d54, both the DEPC and DMPC-d54 components have their melting processes broadened and shifted to lower temperatures, compared with binary lipid mixtures of the same composition. This suggests that protein perturbs both lipids in similar fashion. At high levels of DMPC-d54, the DEPC component exhibits a highly cooperative melting process at temperatures close to that for pure DEPC. This strongly indicates that domains of DEPC are present (at least at low temperatures) in the bilayer, and that Ca2+-ATPase is excluded from these domains. The protein thus exhibits preferential interaction with the DMPC-d54 component. This work demonstrates the utility of FT-IR for identification of the molecular origin of particular domains in reasonably complex lipid mixtures. The relevance of this work to native membrane systems where lipid domains have been observed by several groups is discussed.  相似文献   

5.
Fourier transform infrared spectroscopy has been used to monitor lipid-protein interaction and protein secondary structure in native and reconstituted sarcoplasmic reticulum vesicles. Studies of the temperature dependence of the CH2 symmetric stretching frequency reveal no cooperative phase transitions in purified sarcoplasmic reticulum or in vesicles reconstituted with dioleoylphosphatidylcholine, although a continuous introduction of disorder into the lipid acyl chains is observed as the temperature is raised. In addition, temperature-dependent changes are observed in the Amide I and Amide II vibrations arising from protein peptide bonds. A comparison of lipid order in native sarcoplasmic reticulum and its lipid extract showed that the introduction of protein is accompanied by a slight increase in lipid order. Reconstitution of Ca2+-ATPase from sarcoplasmic reticulum with dipalmitoylphosphatidylcholine (lipid/protein ratio 30:1), reveals a perturbed lipid melting event broadened and reduced in midpoint temperature from multilamellar lipid vesicles. The onset of melting (27–28°C) correlates well with the onset of ATPase activity and confirms a suggestion (Hesketh, T.R., Smith G.A., Houslay M.D., McGill, K.A., Birdsall, N.J.M., Metcalfe, J.C. and Warren, G.B. (1976) Biochemistry 15, 4145–4151) that a liquid crystalline environment is a requirement for optimal protein function. Finally, Ca2+-ATPase has been reconstituted into binary lipid mixtures of DOPC and acyl-chain perdeuterated DPPC. The effect of protein on the structure and melting behavior of each lipid component was monitored. The protein appears to preferentially interact with the DOPC component.  相似文献   

6.
To obtain cationic liposomes of which affinity to negatively charged membranes can be controlled by temperature, cationic liposomes consisting of 3beta-[N-(N', N'-dimethylaminoethane)carbamoyl]cholesterol and dioleoylphosphatidylethanolamine were modified with poly(N-acryloylpyrrolidine), which is a thermosensitive polymer exhibiting a lower critical solution temperature (LCST) at ca. 52 degrees C. The unmodified cationic liposomes did not change its zeta potential between 20-60 degrees C. The polymer-modified cationic liposomes revealed much lower zeta potential values below the LCST of the polymer than the unmodified cationic liposomes. However, their zeta potential increased significantly above this temperature. The unmodified cationic liposomes formed aggregates and fused intensively with anionic liposomes consisting of egg yolk phosphatidylcholine and phosphatidic acid in the region of 20-60 degrees C, due to the electrostatic interaction. In contrast, aggregation and fusion of the polymer-modified cationic liposomes with the anionic liposomes were strongly suppressed below the LCST. However, these interactions were enhanced remarkably above the LCST. In addition, the polymer-modified cationic liposomes did not cause leakage of calcein from the anionic liposomes below the LCST, but promoted the leakage above this temperature as the unmodified cationic liposomes did. Temperature-induced conformational change of the polymer chains from a hydrated coil to a dehydrated globule might affect the affinity of the polymer-modified cationic liposomes to the anionic liposomes.  相似文献   

7.
The interaction of the 36 amino acid neuropeptide Y (NPY) with liposomes was studied using the intrinsic tyrosine fluorescence of NPY and an NPY fragment comprising amino acids 18–36. The vesicular membranes were composed of phosphatidylcholine and phosphatidylserine at varying mixing ratios. From the experimentally measured binding curves, the standard Gibbs free energy for the peptide transfer from aqueous solution to the lipid membrane was calculated to be around ?30 kJ/mol for membrane mixtures containing physiological amounts of acidic lipids at pH 5. The effective charge of the peptide depends on the pH of the buffer and is about half of its theoretical net charge. The results were confirmed using the fluorescence of the NPY analogue [Trp32]-NPY. Further, the position of NPY’s α-helix in the membrane was estimated from the intrinsic tyrosine fluorescence of NPY, from quenching experiments with spin-labelled phospholipids using [Trp32]-NPY, and from 1H magic-angle spinning NMR relaxation measurements using spin-labelled [Ala31, TOAC32]-NPY. The results suggest that the immersion depth of NPY into the membrane is triggered by the membrane composition. The α-helix of NPY is located in the upper chain region of zwitterionic membranes but its position is shifted to the glycerol region in negatively charged membranes. For membranes composed of phosphatidylcholine and phosphatidylserine, an intermediate position of the α-helix is observed.  相似文献   

8.
Fourier transform infrared spectroscopic studies of Ca(2+)-binding proteins   总被引:2,自引:0,他引:2  
M Jackson  P I Haris  D Chapman 《Biochemistry》1991,30(40):9681-9686
The secondary structures of calmodulin and parvalbumin are well established from X-ray diffraction and nuclear magnetic resonance spectroscopic studies, which indicate that these proteins are predominantly alpha-helical in character. Recent infrared studies have nevertheless suggested that the helical structures present in these proteins in solution are not the standard alpha-helix but rather some kind of distorted helices [Trewhella, J., et al. (1989) Biochemistry 28, 1294]. The evidence for this was the unusually low amide I frequency for calmodulin and troponin C in 2H2O solution. The studies presented here, however, suggest that the helical structures in these proteins are not significantly distorted, for two reasons. First, distorted helical structures have weaker hydrogen bonds than the standard alpha-helix and would therefore be expected to absorb at a higher rather than a lower frequency. Second, distorted helical structures would absorb at an unusual frequency in H2O solutions which is not the case for the proteins studied here. The band frequency of these proteins is observed to occur at a frequency observed with other proteins known to contain predominantly alpha-helical structures. Quantitative analysis of the FT-IR spectra of calmodulin (67% alpha-helix) and parvalbumin (68% alpha-helix) in H2O in the presence of Ca2+ gives helical contents similar to those reported by X-ray studies. This raises the question as to why these proteins in H2O show a normal frequency for the presence of alpha-helical structures and an abnormal frequency in 2H2O. Addition of deuterated glycerol to the proteins in 2H2O solutions results in a significant shift of absorbance to higher frequency.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
Fourier transform infrared (FT-IR) spectroscopy and differential scanning calorimetry (DSC) have been used to elucidate the phase behavior of two binary lipid mixtures, acyl chain perdeuterated 1,2-dipalmitoylphosphatidylethanolamine (DPPE-d62)/1,2-dielaidoylphosphatidylcholine (DEPC) and acyl chain perdeuterated 1,2-dipalmitoylphosphatidylcholine (DPPC-d62)/1,2-dimyristoylphosphatidylethanolamine (DMPE). The former shows gel state immiscibility over most of the composition range. The FT-IR data indicate that one of the solid phases is essentially pure DEPC, while the other solid phase contains both lipids. The DPPC-d62/DMPE pair are miscible over the entire composition range. The use of deuterated lipids as one component in the mixture permits the melting characteristics of each component to be separately determined in the FT-IR experiment. The FT-IR data are used to assign the endotherms observed in the DSC to particular molecular components. For the DPPE-d62/DEPC system, two endotherms are observed at compositions between 10 and 67 mol% DPPE-d62. The lower transition is assigned to the DEPC component, while the higher event contains contributions to the enthalpy from both lipids in the mixture. The midpoint of the DEPC melting occurs substantially below that for DPPE-d62. For the miscible pair, each of the lipids melt over approximately the same temperature range. The complementary and consistent nature of the information available from FT-IR and from DSC is demonstrated from the current work.  相似文献   

10.
In an attempt to understand the multifunctional involvement of beta(2)-glycoprotein I (beta(2)GPI) in autoimmune diseases, thrombosis, atherosclerosis, and inflammatory processes, substantial interest is focused on the interaction of beta(2)GPI with negatively charged ligands, in particular, with acidic phospholipids. In this study, unilamellar vesicles composed of cardiolipin were used as in vitro membrane system to test and further refine a model of interaction based on the crystal structure of beta(2)GPI. The data suggest that beta(2)GPI anchors to the membrane surface with its hydrophobic loop adjacent to the positively charged lysine rich region in domain V. Subsequently, beta(2)GPI penetrates the membrane interfacial headgroup region as indicated by a restriction of the lipid side chain mobility, but without formation of a nonbilayer lipid phase. A structural rearrangement of beta(2)GPI upon lipid binding was detected by microcalorimetry and may result in the exposure of cryptic epitopes located in the complement control protein domains. This lipid-dependent conformational change may induce oligomerization of beta(2)GPI and promote intermolecular associations. Thus, the aggregation tendency of beta(2)GPI may serve as the basis for the formation of a molecular link between cells but may also be an essential feature for binding of autoantibodies and hence determine the role of beta(2)GPI in autoimmune diseases.  相似文献   

11.
Glycophorin has been isolated in pure form from human erythrocyte membranes and reconstituted into lipid vesicles composed of binary mixtures of bovine brain phosphatidylserine (PS) and acyl-chain perdeuterated dipalmitoylphosphatidylcholine (DPPC-d62). The effect of protein on lipid melting behavior and order has been monitored with differential scanning calorimetry and Fourier transform infrared spectroscopy (FT-IR). The phase diagram for PS/DPPC-d62 is consistent with that previously reported for PS/DPPC (Stewart et al. (1979) Biochim. Biophys. Acta 556, 1-16) and indicates that acyl chain perdeuteration does not greatly alter the lipid mixing characteristics. The use of deuterated lipid allows the examination of lipid order by FT-IR of each lipid component in the binary mixtures as well as in the ternary (lipid/lipid/protein) systems. Addition of glycophorin to a 30:70 PS/DPPC-d62 binary lipid mixture results in a preferential glycophorin/PS interaction leading to bulk lipid enriched in DPPC-d62. This is revealed in two ways: first, through cooperative calorimetric transitions increased in temperature from the binary lipid system and second, through FT-IR melting curves of the DPPC-d62 component which shows transitions increased in both onset and completion temperatures in the presence of protein. In addition, non-cooperative melting events are observed at temperatures below the onset of phase separation. The FT-IR data are used to assign these non-cooperative events to the melting of the PS component. For the 50:50 lipid mixture with protein, two transitions are observed in the DSC experiments. The IR results indicate that both lipid components are involved with the lower temperature event.  相似文献   

12.
The effect of divalent ion binding to deionized bacteriorhodopsin (dI-bR) on the thermal transitions of the protein secondary structure have been studied by using temperature-dependent Fourier transform infrared (FT-IR) spectroscopy. The native metal ions in bR, Ca(2+), and Mg(2+), which we studied previously, are compared with Mn(2+), Hg(2+), and a large, synthesized divalent organic cation, ((Et)(3)N)(2)Bu(2+). It was found that in all cases of ion regeneration, there is a pre-melting, reversible conformational transition in which the amide frequency shifts from 1665 to 1652 cm(-1). This always occurs at approximately 80 degrees C, independent of which cation is used for the regeneration. The irreversible thermal transition (melting), monitored by the appearance of the band at 1623 cm(-1), is found to occur at a lower temperature than that for the native bR but higher than that for acid blue bR in all cases. However, the temperature for this transition is dependent on the identity of the cation. Furthermore, it is shown that the mechanism of melting of the organic cation regenerated bR is different than for the metal cations, suggesting a difference in the type of binding to the protein (either to different sites or different binding to the same site). These results are used to propose specific direct binding mechanisms of the ions to the protein of deionized bR.  相似文献   

13.
Fourier transform infrared spectroscopy has been used to investigate the secondary structure of porcine and bovine pancreatic phospholipase A2 (PLA2) and the zymogen of porcine PLA2, prophospholipase A2 (proPLA2), in both H2O and D2O media. Detailed qualitative analysis was made of these proteins using second derivative and deconvolution techniques. Quantitative studies of the proteins in solution made using Factor Analysis gave average values of 54% alpha-helix, 15% beta-sheet and 23% beta-turns. These values agree well with the secondary structures deduced from previous studies of single crystals using X-ray techniques. No significant differences in secondary structure were observed for porcine pancreatic (pro)phospholipase A2 in the presence or absence of Ca2+ ions, or in the temperature range 10-45 degrees C. The binding of the non-degradable phospholipid analogue, n-alkylphosphocholine, in monomeric form produced no significant difference in the secondary structure of either enzyme. Conformational differences were, however, observed between the enzyme lyophilised in a solid film and in aqueous solution. The change is probably due to the formation of beta-sheet upon hydration, coupled with a loss of random structures. Conformational differences in both porcine and bovine pancreatic PLA2 were observed on binding to n-alkylphosphocholine micelles. This change may be due to a small increase in the alpha-helical structure and a decrease in the beta-sheet, and/or possibly beta-turn content. Similar conformational changes were observed for the interaction of porcine and bovine PLA2 with the substrate analogue inhibitor 1-heptanoyl-2-heptanoylamino-2-deoxy-sn-glycero-3-phospho glycol in micellar form.  相似文献   

14.
Ion binding and lipid ionization of the acidic phospholipid 1,2-dimyristoyl-sn-glycero-3-phosphoglycerol (DMPG) in monolayers was studied by measuring the lateral pressure Pi as a function of the molecular area A at the air/water interface at different temperatures. The pH of the subphase (pH 2 and 7) and the ionic strength (NaCl) was varied. In addition, different divalent cations (1mM MgCl2, CaCl2 and SrCl2, pH 7) were added. DMPG is partly protonated on pure water at pH 7. An increase in the NaCl concentration in the subphase leads to film expansion. This effect is caused by an ionization of the headgroup of DMPG, i.e. a shift of the apparent pK. More condensed films are obtained on pure water at pH 2, due to the reduction of electrostatic repulsion by headgroup protonation and the possibility for the formation of a hydrogen bonding network. The divalent cations Mg2+, Ca2+ and Sr2+ interact differently with a DMPG monolayer in pure water at pH 7. In the presence of 1mM CaCl2 a condensation of the DMPG film is induced, whereas an expansion of the monolayer is observed in the presence of Mg2+ and Sr2+. Two counteracting effects are operative: (a) ionization of the headgroup due to electrostatic screening leads to film expansion and (b) binding of the divalent cations to the lipid headgroups leads to condensation. The latter effect is more pronounced in the case of Ca2+, whereas the binding of Mg2+ and Sr2+ to DMPG is weaker. Site-specific cation binding has to be assumed in addition to electrostatic effects.  相似文献   

15.
Fourier transform infrared (FTIR) attenuated total reflection (ATR) spectroscopy was used to elucidate the hydration behavior and molecular order of phospholipid/ganglioside bilayers. We examined dry and hydrated films of the gangliosides GM1, deacetyl-GM1, lyso-GM1, deacetyllyso-GM1, and GM3 and oriented mixed films of these gangliosides with 1,2-dimyristoyl-sn-glycero-3-phosphorylcholine (DMPC) using polarized light. Analysis of the amide I frequencies reveals that the amide groups are involved in intermolecular interactions via hydrogen bonds of varying strengths. The tilt angle of the acyl chains of the lipids in mixed films was determined as a function of ganglioside structure. Deacetylation of the sialic acid in the headgroup has a stronger influence on the tilt angle than the removal of the ganglioside fatty acid. The phase behavior was examined by FTIR ATR spectroscopy and by differential scanning calorimetry (DSC) measurements on lipid suspensions. At the same molar concentration, lyso-gangliosides have less effect on changes of transition temperature compared to the double-chain analogs. Distinct differences in the amide band shapes were observed between mixtures with lyso-gangliosides and normal double-chain gangliosides. Determined from the dicroic ratio RATR, the orientation of the COO- group in all DMPC/ganglioside mixtures was found to be relatively fixed with respect to the membrane normal. In 4:1 mixtures of DMPC with GM1 and deacetyl-GM1, the binding of Ca2+ leads to a slight decrease in chain tilt in the gel phase, probably caused by a dehydration of the membrane-water interface. In mixtures of DMPC with GM3 and deacetyl-lyso-GM1, a slight increase in chain tilt is observed. The chain tilt in DMPC/lyso-GM1 mixtures is unchanged. Analysis of the COO- band reveals that Ca2+ does not bind to the carboxylate group of the sialic acid of GM1 and deacetyl-GM1, the mixtures in which a decrease in chain tilt was observed. Binding to the sialic acid was only observed for mixtures of DMPC with GM3, lyso-GM1, and deacetyl-lyso-GM1. Ca2+ obviously accumulates at the bilayer-water interface and leads to partial dehydration of the headgroup region in the gel as well as in the liquid-crystalline phase. This can be concluded from the changes in the amide I band shapes. With the exception of DMPC/deacetyl-GM1, the effects on the ester C==O bands are small. The addition of Ca2+ has minor effects on the phase behavior, with the exception of the DMPC/GM1 mixture.  相似文献   

16.
Deconvolved and second derivative Fourier transform infrared spectra of the proteins flavodoxin and triosephosphate isomerase have been obtained in the 1600 to 1700 cm-1 (amide I) region. To our knowledge these results provide the first experimental infrared data on proteins with parallel beta-chains. Characteristic absorption bands for the parallel beta-segments are observed at 1626-1639 cm-1 (strong) and close to 1675 cm-1 (weak). Previous theoretical studies based on hypothetical models with large, regular beta-sheets had suggested bands close to 1650 and 1666 cm-1. Our new assignments were confirmed by band area measurements, which yield conformational information in good agreement with results from X-ray diffraction data. The spectra were compared with corresponding spectra of concanavalin A and carboxypeptidase A. The first contains only antiparallel beta-segments, the second "mixed" beta-segments, with some strands lying antiparallel and others parallel. None of the observed amide I band frequencies assigned to parallel beta-chains occurs in the 1650 cm-1 region associated with helical segments.  相似文献   

17.
Polarization-modulated infrared reflection absorption spectroscopy (PM-IRRAS) was used to follow the hydrolysis of phospholipid monolayers at the air-water interface by phospholipase A2 (PLA2). The decrease in the intensity of the nuC=O ester band of dipalmitoylphosphatidylcholine at 1733 cm(-1) and the appearance of two new infrared bands in the 1530-1580 cm(-1) region allowed to monitor phospholipid hydrolysis by PLA2. Indeed, the decrease in the intensity of the band at 1733 cm(-1) was attributed to the enzymatic hydrolysis of the acyl ester linkage of the sn-2 fatty acid on the glycerol backbone whereas the doublet appearing at 1537 and 1575 cm(-1) was attributed to the nu(a) COO- vibration of the newly formed calcium-palmitate. The presence of this band as a doublet indicates the formation of a crystalline-like calcium-palmitate monolayer. This observation supports our previously postulated mechanism for the formation of PLA2 domains at the air-water interface. Definitive assignment of the infrared bands has been possible by measuring PM-IRRAS spectra of the individual hydrolysis products (palmitic acid and lysopalmitoylphosphatidylcholine) as well as of 1-caproyl-2-palmitoyl-phosphatidylcholine and 1-palmitoyl-2-caproylphosphatidylcholine monolayers before and after hydrolysis by PLA2.  相似文献   

18.
Fourier transform infrared spectroscopy has been used to study the secondary structure of the human erythrocyte glucose transporter after purification and reconstitution in erythrocyte lipids. The spectra indicate that the glucose transporter contains, in addition to the predominant alpha-helical structure, an appreciable amount of beta-structure and random coil conformation. A study of the time dependency of H-2H exchange revealed that more than 80% of the polypeptide backbone is readily accessible to the solvent. This result indicates that a portion of the intramembrane-spanning region of the membrane protein is exposed to the solvent, suggesting the existence of an intraprotein aqueous channel. The residual (10-20%) portion of the protein which exchanges slowly includes some alpha-helical structure, probably situated in a hydrophobic environment inside the membrane. The infrared spectra of transporter preparations were also examined after incubation with substrate and substrate analogues. Compared with the spectra recorded under conditions in which the "inward-facing" form predominates, a small but reproducible shift in the bands assigned to alpha-helical and beta-strand structures is observed after incubation with 4,6-O-ethylidene-D-glucose, which largely fixes the transporter in the "outward-facing" conformation. An increase of temperature, which is known to increase the proportion of transporter in the outward-facing conformation, results in a similar shift in this alpha-helical absorption band.  相似文献   

19.
The secondary structure of streptokinase (Sk) in aqueous solution was quantitatively examined by using Fourier transform infrared (FT-IR) spectroscopy. Resolution enhancement techniques, including Fourier deconvolution and derivative spectroscopy, were combined with band curve-fitting procedures to quantitate the spectral information from the amide I bands. Nine component bands were found under the broad, nearly featureless amide I bands which reflect the presence of various substructures. The relative areas of these component bands indicate an amount of beta-sheet between 30 and 37% and an alpha-helix content of only 12-13% in Sk. Further conformational substructures are assigned to turns (25-26%) and to "random" structures (15-16%). Additionally, the correlation of a pronounced component band near 1640 cm-1 (10-16% fractional area) with the possible presence of 3(10)-helices is discussed.  相似文献   

20.
Fourier transform infrared and laser Raman spectroscopies were used to study the effects of dodecylpyridinium bromide on the conformation of haemoglobin, myoglobin, bovine serum albumin, ribonuclease, ovalbumin, lysozyme, trypsin and beta-lactoglobulin in aqueous solution. Addition of the cationic detergent caused a decrease in alpha-helix conformation in highly helical proteins. At low detergent concentrations stabilization of beta-sheet conformation was observed.  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