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 531 毫秒
1.
To investigate the role of the N-terminal region in the lytic mechanism of the pore-forming toxin sticholysin II (St II), we studied the conformational and functional properties of peptides encompassing the first 30 residues of the protein. Peptides containing residues 1-30 (P1-30) and 11-30 (P11-30) were synthesized and their conformational properties were examined in aqueous solution as a function of peptide concentration, pH, ionic strength, and addition of the secondary structure-inducing solvent trifluoroethanol (TFE). CD spectra showed that increasing concentration, pH, and ionic strength led to aggregation of P1-30; as a consequence, the peptide acquired beta-sheet conformation. In contrast, P11-30 exhibited practically no conformational changes under the same conditions, remaining essentially structureless. Moreover, this peptide did not undergo aggregation. These differences clearly point to the modulating effect of the first 10 hydrophobic residues on the peptides aggregation and conformational properties. In TFE both the first ten hydrophobic peptides acquired alpha-helical conformation, albeit to a different extent, P11-30 displayed lower alpha-helical content. P1-30 presented a larger fraction of residues in alpha-helical conformation in TFE than that found in St II's crystal structure for that portion of the protein. Since TFE mimics the membrane environment, such increase in helical content could also occur upon toxin binding to membranes and represent a step in the mechanism of pore formation. The peptides conformational properties correlated well with their functional behavior. Thus, P1-30 exhibited much higher hemolytic activity than P11-30. In addition, P11-30 was able to block the toxin's hemolytic activity. The size of pores formed in red blood cells by P1-30 was estimated by measuring the permeability to PEGs of different molecular mass. The pore radius (0.95 +/- 0.01 nm) was very similar to that of the pore formed by the toxin. The results demonstrate that the synthetic peptide P1-30 is a good model of St II conformation and function and emphasize the contribution of the toxin's N-terminal region, and, in particular, the hydrophobic residues 1-10 to pore formation.  相似文献   

2.
Laser Raman spectra of cobrotoxin under various conditions have been obtained. Comparison of the spectra of native cobrotoxin in lyophilized form and in aqueous solution indicates that the secondary structures of cobrotoxin are not significantly affected by the removal of the aqueous solvent. On going from the native to the partially reduced and the completely reduced, carboxy-methylated forms, characteristic peaks of the C-S-S-C and tyrosine ring in the region of 500--900 cm-1 showed definite changes in structure. The partially reduced form gave two peaks at 502 and 524 cm-1, suggesting difference in the conformation of the remaining disulfide bonds. As indicated by the present work, the conformation of the main chain of cobrotoxin in the native unperturbed state, in the partially reduced and in the completely reduced forms are the coexistence of beta-pleated sheet with random-coil structure, predominantly random coil, and predominantly random coil with the existence of an alpha-helix type structure, respectively. The effect of pH on the conformation of cobrotoxin in solution appeared to give rise to the change of the local structure of two aromatic residues common to all snake neurotoxins.  相似文献   

3.
The alpha-subunit of the nicotinic acetylcholine receptor (alphaAChR) contains a binding site for alpha-bungarotoxin (alpha-BTX), a snake-venom-derived alpha-neurotoxin. Previous studies have established that the segment comprising residues 173-204 of alphaAChR contains the major determinant interacting with the toxin, but the precise boundaries of this determinant have not been clearly defined to date. In this study, we applied NMR dynamic filtering to determine the exact sequence constituting the major alphaAChR determinant interacting with alpha-BTX. Two overlapping synthetic peptides corresponding to segments 179-200 and 182-202 of the alphaAChR were complexed with alpha-BTX. HOHAHA and ROESY spectra of these complexes acquired with long mixing times highlight the residues of the peptide that do not interact with the toxin and retain considerable mobility upon binding to alpha-BTX. These results, together with changes in the chemical shifts of the peptide protons upon complex formation, suggest that residues 184-200 form the contact region. At pH 4, the molecular mass of the complex determined by dynamic light scattering (DLS) was found to be 11.2 kDa, in excellent agreement with the expected molecular mass of a 1:1 complex, while at pH >5 the DLS measurement of 20 kDa molecular mass indicated dimerization of the complex. These results were supported by T(2) measurements. Complete resonance assignment of the 11.2 kDa complex of alpha-BTX bound to the alphaAChR peptide comprising residues 182-202 was obtained at pH 4 using homonuclear 2D NMR spectra measured at 800 MHz. The secondary structures of both alpha-BTX and the bound alphaAChR peptide were determined using 2D (1)H NMR experiments. The peptide folds into a beta-hairpin conformation, in which residues (R)H186-(R)V188 and (R)Y198-(R)D200 form the two beta-strands. Residues (R)Y189-(R)T191 form an intermolecular beta-sheet with residues (B)K38-(B)V40 of the second finger of alpha-BTX. These results accurately pinpoint the alpha-BTX-binding site on the alphaAChR and pave the way to structure determination of this important alphaAChR determinant involved in binding acetylcholine and cholinergic agonists and antagonists.  相似文献   

4.
Clostridium difficile toxin B (269 kDa), which is one of the causative agents of antibiotic-associated diarrhea and pseudomembranous colitis, inactivates Rho GTPases by glucosylation. Here we studied the uptake and membrane interaction of the toxin with eukaryotic target cells. Bafilomycin A1, which prevents acidification of endosomal compartments, blocked the cellular uptake of toxin B in Chinese hamster ovary cells cells. Extracellular acidification (pH 相似文献   

5.
A Fourier-transform infrared spectroscopic study of hen egg phosvitin and ovalbumin has been carried out. Bands arising from monoanionic and dianionic phosphate monoester [Shimanouchi, T., Tsuboi, M., & Kyogoku, Y. (1964) Adv. Chem. Phys. 8, 435-498] can be identified easily in the 1300-930 cm-1 region in spectra of solutions of O-phosphoserine and phosvitin, a highly phosphorylated protein. On the other hand, spectra of ovalbumin show a relatively strong absorption above 1000 cm-1 arising from the protein moiety. Below 1000 cm-1, a single band at 979 cm-1 is observed; this band is not present in spectra of dephosphorylated ovalbumin, and therefore, it has been assigned to the symmetric stretching of the phosphorylated Ser-68 and Ser-344 in the dianionic ionization state. In addition, bands arising from symmetric and antisymmetric stretchings of the monoanionic ionization state, and from the antisymmetric stretching of the dianionic state, can be detected above 1000 cm-1 in difference spectra of ovalbumin minus dephosphorylated ovalbumin. The effect of pH on the infrared spectra of O-phosphoserine, phosvitin, and ovalbumin is consistent with the phosphoserine residues undergoing ionization with pK values about 6. This study demonstrates that Fourier-transform infrared spectroscopy can be a useful technique to assess the ionization state of phosphoserine residues in proteins in solution.  相似文献   

6.
The Z-DNA crystal structures of d(CGCGTG) and d(CGCGCG) are compared by laser Raman spectroscopy. Raman bands originating from vibrations of the phosphodiester groups and sensitive to the DNA backbone conformation are similar for the two structures, indicating no significant perturbation to the Z-DNA backbone as a result of the incorporation of G.T mismatches. Both Z structures also exhibit Raman markers at 625 and 670 cm-1, assigned respectively to C3'-endo/syn-dG (internal) and C2'-endo/syn-dG conformers (3' terminus). Additional Raman intensity near 620 and 670 cm-1 in the spectrum of the d(CGCGTG) crystal is assigned to C4'-exo/syn-dG conformers at the mismatch sites (penultimate from the 5' terminus). A Raman band at 1680 cm-1, detected only in the d(CGCGTG) crystal, is assigned to the hydrogen-bonded dT residues and is proposed as a definitive marker of the Z-DNA wobble G.T pair. For aqueous solutions, the Raman spectra of d(CGCGTG) and d(CGCGCG) are those of B-DNA, but with significant differences between them. For example, the usual B-form marker band at 832 cm-1 in the spectrum of d(CGCGTG) is about 40% less intense than the corresponding band in the spectrum of d(CGCGCG), and the former structure exhibits a companion band at 864 cm-1 not observed for d(CGCGCG). The simplest interpretation of these results is that the conventional B-form OPO geometry occurs for only 6 of the 10 OPO groups of d(CGCGTG). The remaining four OPO groups, believed to be those at or near the mismatch site, are in an "unusual B" conformation which generates the 864 cm-1 band.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
We report Raman spectra of various cholinesterases: lytic tetrameric forms (G4) obtained by tryptic digestion of asymmetric acetylcholinesterase (AChE) from Torpedo californica and Electrophorus electricus, a PI-PLC-treated dimeric form (G2) of AChE from T marmorata, and the soluble tetrameric form (G4) of butyrylcholinesterase (BuChE) from human plasma. The contribution of different types of secondary structure was estimated by analyzing the amide I band, using the method of Williams. The spectra of cholinesterases in 10 mM Tris-HCl (pH 7.0) indicate the presence of both alpha-helices (about 50%) and beta-sheets (about 25%), together with 15% turns and 10% undefined structures. In 20 mM phosphate buffer (pH 7.0), the spectra indicated a smaller contribution of alpha-helical structure (about 35%) and an increased beta-sheet content (from 25 to 35%). This shows that the ionic milieu profoundly affects either the conformation of the protein (AChE activity is known to be sensitive to ionic strength), or the evaluation of secondary structure, or both. In addition, we analyzed vibrations corresponding to the side chains of aromatic and aliphatic amino acids. In particular, the analyses of the tyrosine doublet (830-850 cm-1) and of the tryptophan vibration at 880 cm-1 indicated that these residues are predominantly 'exposed' on the surface of the molecules.  相似文献   

8.
The degradation of bovine myelin basic protein by bovine brain cathepsin D (ED 3.4.23.5) was studied over a pH range of 2.75 - 6.0. Throughout this pH range pepstatin, an inhibitor of cathepsin D, prevented the degradation. The degradation at a pH away from the optimum of pH 3.5 was predictably slower, but also resulted in more restricted cleavage. Above pH 4.5 bovine basic protein peptide 1 - 42 was not degraded further to peptide 1 - 36 as occurs at pH 3.5. Additionally, at pH 5.5 another fragment of basic protein, peptide 1 - 91, persisted indicating that under certain basic protein as well as basic protein peptide 43 - 169 may be cleaved in the molecular region of basic protein around the phenylalanyl-phenylalanine residues at position 88 - 89. The small amount of peptides 1 - 91 and 92 - 169 detected at pH 5.5 suggests that the bond between residues 91 and 92 in intact basic protein is a minor cleavage site. The options and variation in cleavage around residues 88 - 92 of basic protein presumably result from pH-dependent changes in conformation in the is region but could also be due to changes in conformation of cathepsin D. These results indicate that local tissue changes such a pH amy affect not only the velocity of the reaction but also the nature of th product formed by the degradation of basic protein by brain cathepsin D  相似文献   

9.
The solution structure of a synthetic peptide, Cn2(1-15)NH2-S-S-acetyl-Cn2(52-66)NH2 from toxin 2 (Cn2) of the New World scorpion Centruroides noxius was determined using nmr and molecular dynamics calculations. The peptide has no significant secondary structure such as an alpha-helix or a beta-sheet, yet it has a fixed conformation for the first chain. The backbone secondary structure involving residues 6-12 in this peptide shows an excellent overlap with the structures of natural neurotoxins from Centruroides sculpturatus Ewing. Residues 6-9 form a distorted type I beta-turn and residues 10-12 form a gamma-turn. As residues 7-10 in the Centruroides toxins correspond to one of the regions of highest sequence variability, it may account for the species specificity and/or selectivity of toxic action. The conformation of this region evidently plays an important role in receptor recognition and in binding to the neutralizing monoclonal antibody BCF2 raised against the intact toxin.  相似文献   

10.
The solution conformation of endothelium-derived vasoconstrictor peptide, endothelin, has been determined by two-dimensional 1H-NMR spectroscopy and distance geometry. Conformation in the N-terminal core region (residues 1-15) is well-defined and a characteristic is the helix-like conformation in the segment from Lys9 to Cys15. Contrarily, the C-terminal tail region (residues 16-21) does not assume a defined conformation and there are no specific interactions between the core and the tail regions.  相似文献   

11.
Fourier-transform infrared spectroscopy is a valuable method for the study of protein conformation in solution primarily because of the sensitivity to conformation of the amide I band (1700-1620 cm-1) which arises from the backbone C = O stretching vibration. Combined with resolution-enhancement techniques such as derivative spectroscopy and self-deconvolution, plus the application of iterative curve-fitting techniques, this method provides a wealth of information concerning protein secondary structure. Further extraction of conformational information from the amide I band is dependent upon discerning the correlations between specific conformational types and component bands in the amide I region. In this paper, we report spectra-structure correlations derived from conformational perturbations in bovine trypsin which arise from autolytic processing, zymogen activation, and active-site inhibition. IR spectra were collected for the single-chain (beta-trypsin) and once-cleaved, double-chain (alpha-trypsin) forms as well as at various times during the course of autolysis and also for zymogen, trypsinogen, and beta-trypsin inhibited with diisopropyl fluorophosphate. Spectral differences among the various molecular forms were interpreted in light of previous biochemical studies of autolysis and the known three-dimensional structures of the zymogen, the active enzyme, and the DIP-inhibited form. Our spectroscopic results from these proteins in D2O imply that certain loop structures may absorb in the region of 1655 cm-1. Previously, amide I' infrared bands near 1655 cm-1 have been interpreted as arising solely from alpha-helices. These new data suggest caution in interpreting this band. We have also proposed that regions of protein molecules which are known from crystallographic experiments to be disordered absorb in the 1645 cm-1 region and that type II beta-turns absorb in the region of 1672-1685 cm-1. Our results also corroborate assignment of the low-frequency component of extended strands to bands below 1636 cm-1. Additionally, the results of multiple measurements have allowed us to estimate the variability present in component band areas calculated by curve fitting the resolution-enhanced IR spectra. We estimate that this approach to data analysis and interpretation is sensitive to changes of 0.01 unit or less in the relative integrated intensities of component bands in spectra whose peaks are well resolved.  相似文献   

12.
Y Nishimura  M Tsuboi    T Sato 《Nucleic acids research》1984,12(17):6901-6908
Raman spectra of nine crystals of known structures which involve guanosine moieties with various conformations have been observed. It has been established that a guanosine residue with the C3'endo-anti conformation gives a strong Raman line at 666 +/- 2 cm-1. It has also been found that the residue with 04'endo-anti gives a strong Raman line at 682 cm-1, and C3'exo-syn at 616 cm-1. The usefulness of these structure-spectrum correlations in the conformation studies of polynucleotides are shown.  相似文献   

13.
Polygalacturonases hydrolyze the alpha-(1-4) glycosidic bonds of de-esterified pectate in the smooth region of the plant cell wall. Crystal structures of polygalacturonase from Aspergillus aculeatus were determined at pH 4.5 and 8.5 both to 2.0 A resolution. A. aculeatus polygalacturonase is a glycoprotein with one N and ten O-glycosylation sites and folds into a right-handed parallel beta-helix. The structures of the three independent molecules are essentially the same, showing no dependency on pH or crystal packing, and are very similar to that of Aspergillus niger polygalacturonase. However, the structures of the long T1 loop containing a catalytic tyrosine residue are significantly different in the two proteins. A three-dimensional model showing the substrate binding mode for a family 28 hydrolase was obtained by a combined approach of flexible docking, molecular dynamics simulations, and energy minimization. The octagalacturonate substrate was modeled as an unbent irregular helix with the -1 ring in a half-chair ((4)H(3)) form that approaches the transition state conformation. A comparative modeling of the three polygalacturonases with known structure shows that six subsites ranging from -4 to +2 are clearly defined but subsites -5 and +3 may or may not be shaped depending on the nearby amino acid residues. Both distal subsites are mostly exposed to the solvent region and have weak binding affinity even if they exist. The complex model provides a clear explanation for the functions, either in catalysis or in substrate binding, of all conserved amino acid residues in the polygalacturonase family of proteins. Modeling suggests that the role of the conserved Asn157 and Tyr270, which had previously been unidentified, may be in transition state stabilization. In A. niger polygalacturonase, the long T1 loop may have to undergo conformational change upon binding of the substrate to bring the tyrosine residue close to subsite -1.  相似文献   

14.
The substitution of iron for cobalt in the monomeric insect hemoglobin CTT (Chironomus thummi thummi) III does not alter the Bohr effect for O2-binding. The cobalt substitution in this hemoglobin allows us to identify not only the O-O and Co-O2 stretching mode but also the Co-O-O bending mode by resonance Raman spectroscopy. The assignments were made via 16O2/18O2 isotope exchange. The modes associated with the Co-O-O moiety are pH-dependent. These pH-induced changes of the resonance Raman spectra are correlated with the t = r conformation transition. At high pH (high-affinity state) two unperturbed O-O stretching modes are observed at 1,068 cm-1 (major component) and 1,093 cm-1 (minor component) for the 18O2 complex. These frequencies correspond to split modes at 1,107 cm-1 and 1,136 cm-1 and an unperturbed mode at approximately 1,153 cm-1 for the 16O2 complex. At low pH (low-affinity state) the minor component becomes the major component and vice versa. The Co-O2 stretching frequency varies for approximately 520 cm-1 (pH 5.5) to 537 cm-1 (pH 9.5) indicating a stronger (hence shorter) Co-O2 bond in the high-affinity state. On the other hand, the O-O bond is weakened upon the conversion of the low- to the high-affinity state. The Co-O-O bending mode changes from 390 cm-1 (pH 9.5) to 374 cm-1 (pH 5.5). In the deoxy form the resonance Raman spectra are essentially pH-insensitive except for a vinyl mode at 414 cm-1 (pH 5.5), which is shifted to 416 cm-1 (pH 5.5).  相似文献   

15.
Raman spectra have been measured for intact rat lens nuclei at various stages of aging in an attempt to gain further insight into age-related structural changes in the lens proteins, especially changes concerning protein sulfhydryl groups. Two Raman bands at 2579 and 2561 cm-1 were observed to be assignable to SH stretching modes of the cysteine residues. These bands have been attributed to "exposed" and "buried" sulfhydryl groups of the lens proteins, respectively, on the basis of a model compound study. The relative intensities of both SH stretching modes decreased with lens aging, and concurrently the intensity of a S-S stretching mode at 509 cm-1 due to disulfide bridges increased, suggesting that not only exposed but also buried protein sulfhydryl groups are converted to disulfide groups as a result of aging. The rate of the intensity decrease in the 2561 cm-1 band was similar to that in the 2579 cm-1 band. Therefore, it seems likely that the sulfhydryl groups in the two distinct environments are nearly equally subjected to the oxidation. Cysteine and cystine residues of the lens proteins gave their C-S stretching modes at 708 cm-1, indicating that they predominantly assume PC and/or PN conformers. The intensity ratio of a tyrosine doublet near 840 cm-1 (I832/I855) changed from approximately 0.86 to approximately 0.81 with the aging of the rat lens. This result implies that some tyrosine residues undergo a change in their hydrogen bonding environments during the course of aging. Of particular importance is that the relative intensity change of the tyrosine doublet with normal aging and that with cataract formation are in opposite directions.  相似文献   

16.
The further characterization of toxin I from venom of the scorpion Centruroides sculpturatus Ewing (region, Southwestern United States) is reported. Toxin I is a single polypeptide chain of 64 amino acid residues crosslinked by four disulfide bridges. The complete amino acid sequence of toxin I was deduced from the sequence of its tryptic peptides and overlaps provided by its chymotryptic peptides. Toxin I has an amino terminal lysyl residue and a carboxyl terminal threonyl residue.The amino acid sequences of toxin I and neurotoxic variants 1, 2, and 3, likewise isolated from C. sculpturatus venom, differ at 26 positions.The sequences of toxin I from C. sculpturatus and toxins I and II from the North African scorpion, Androctonus australis Hector, are also compared.  相似文献   

17.
Studies on a single component human cryoimmunoglobulin (cryo-IgG) (gamma 1 : lambda, Gm 4) were undertaken to gain a better understanding of the conformational stability of macromolecular interfaces essential for self-association of cryo-IgG leading to the formation of visible gel mass. Changes in the gross and localized conformation of cryo-IgG and a monoclonal IgG (gamma 1 : lambda, Gm 4) isolated from a myeloma patient (Hy) (Hy IgG) (gamma 1 : lambda, Gm 4) in alkaline media were determined by analytical ultracentrifugation, fluorescence characteristics, tyrosine ionization and H+ titration. Ultracentrifugal studies revealed that major transition in gross conformation took place at pH 11.4 for cryo-IgG and pH 11.7 for Hy IgG, whereby the number of charges and tyrosine residues exposed to aqueous environment was 110 and 26 for cryo-IgG, and 111 and 48 for Hy IgG, respectively. Beyond this transition pH fragmentation of both the proteins occurred and cryo-IgG lost its capacity for gel formation. Self-association of cryo-IgG was observed upto pH 11.4 in decreasing order with increase in denaturation pH. Cryo-IgG renatured from exposure to higher alkaline pH upto pH 11.4, showed the capability for forming gel, in spite of the irreversible local conformational changes as established by direct and reverse fluorimetric titration and tyrosine ionization studies. Cryo-IgG could be maintained in the optically clear sol phase at pH 10.5, at which pH 12 out of 62 tyrosine residues became exposed to aqueous media. There are distinct differences in the accessibility of tyrosine residues of cryo-IgG and Hy IgG as reflected in their tyrosine ionization profiles.  相似文献   

18.
We have undertaken a new and more detailed Fourier-transform infrared (FTIR) spectroscopic study of alpha-lactalbumin (in D2O solution) aimed at correlating its secondary structures to observed Amide I' infrared bands. The spectra reported here were interpreted in light of the recently determined crystal structure of alpha-lactalbumin and by comparison with the spectra and structure of the homologous protein lysozyme. Of particular importance is the new evidence supporting the assignment of the band at 1639 cm-1 to 3(10)-helices. This assignment is in excellent agreement with one based on theoretical and experimental studies of 3(10)-helical polypeptides. The frequency observed for 3(10)-helices is distinctly different from that at which alpha-helices are typically found (viz., around 1655 cm-1). In the present study, two bands are clearly resolved in the latter region at 1651 and 1659 cm-1. Both are apparently associated with alpha-helices. These results suggest that for D2O solutions of globular proteins. FTIR spectroscopy can be a facile method for detecting the presence of these two different types of helical conformation and distinguishing between them. This provides a distinct advantage over ultraviolet circular dichroism spectroscopy (UV-CD). This work also provides a basis for future studies of alpha-lactalbumin which examine the effects of environment (e.g., pH, temperature) and ligands (e.g., Ca2+, Mn2+) on its conformation.  相似文献   

19.
Conformational changes in the prion protein (PrP) seem to be responsible for prion diseases. We have used conformation-dependent chemical-shift measurements and rotational-resonance distance measurements to analyze the conformation of solid-state peptides lacking long-range order, corresponding to a region of PrP designated H1. This region is predicted to undergo a transformation of secondary structure in generating the infectious form of the protein. Solid-state NMR spectra of specifically 13C-enriched samples of H1, residues 109-122 (MKHMAGAAAAGAVV) of Syrian hamster PrP, have been acquired under cross-polarization and magic-angle spinning conditions. Samples lyophilized from 50% acetonitrile/50% water show chemical shifts characteristic of a beta-sheet conformation in the region corresponding to residues 112-121, whereas samples lyophilized from hexafluoroisopropanol display shifts indicative of alpha-helical secondary structure in the region corresponding to residues 113-117. Complete conversion to the helical conformation was not observed and conversion from alpha-helix back to beta-sheet, as inferred from the solid-state NMR spectra, occurred when samples were exposed to water. Rotational-resonance experiments were performed on seven doubly 13C-labeled H1 samples dried from water. Measured distances suggest that the peptide is in an extended, possibly beta-strand, conformation. These results are consistent with the experimental observation that PrP can exist in different conformational states and with structural predictions based on biological data and theoretical modeling that suggest that H1 may play a key role in the conformational transition involved in the development of prion diseases.  相似文献   

20.
Vibrational Raman spectra of the solid and gel phases of bovine brain cerebrosides and the component fractions, kerasin and phrenosin, provide conformational information for these glycosphingolipids in bilayer systems. The carbon-carbon stretching mode profiles (1,150-1,000 cm-1) indicate that at 22 degrees C the alkyl chains assume an almost all-trans arrangement. These spectral data, combined with those from the C-H stretching region (3,050-2,800 cm-1), show that phrenosin forms the most highly ordered polycrystalline solid and kerasin the most ordered gel phase. The conformation of the unsaturated, 24-carbon acyl chains is monitored independently by a skeletal stretching mode at 1,112 cm-1. The alkyl chains in the kerasin and phrenosin gels are sufficiently extended to allow interdigitation of the 24-carbon acyl chains across the midplane of the bilayer. The amide I vibrational mode occurs at a lower frequency in solid phrenosin than kerasin, a shift consistent with stronger hydrogen bounding. This band is broadened and shifted to higher frequencies, however, in the phrenosin gel phase. In both the solid and gel phases natural cerebroside exhibits a composite amide I mode. The disruptive effects on cerebroside chain packing and headgroup orientation arising from mixing with dimyristoyl phosphatidylcholine are examined. Vibrational data for cerebroside are also compared to those for ceramide, sphingosine, and distearoyl phosphatidylcholine structures. Spectral interpretations are discussed in terms of calorimetric and X-ray structural data.  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