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 31 毫秒
1.
Microtubules are an important cytoskeletal component involved in cell motility and morphogenesis. They are unique polymers because they are highly dynamic in vivo and in vitro, displaying spontaneous transitions between phases of elongation and rapid shortening. This property has been termed microtubule dynamic instability. Here we describe the application of negative-stain electron microscopy to examine the morphology of microtubules. The purpose was to provide insight into the structural basis of dynamic instability. Highly purified porcine brain tubulin was seeded from isolated axoneme seeds and the morphologies of the tubulin polymers were examined. As previously reported, tubulin polymer sheets in addition to intact MTs were observed during elongation. Cross-sections of identical preparations displayed intact circles (MTs) and c-shaped polymers. The fraction of sheets (28%) observed by negative staining was identical to the fraction of c-shaped polymers in cross-sections. These results suggest that tubulin polymer elongation is not strictly helical due to the lack of helical symmetry of tubulin sheets. Finally, we document the novel effect of glutaraldehyde fixation on MTs. Fixation of MTs in 1% glutaraldehyde for longer than 2 min severely disrupted MT protofilament structure and induced MT curvature at a gross level. Even at 1 min, thin, thread-like structures were observed that were only present in glutaraldehyde-fixed samples. It is therefore an advantage to minimize the extent of glutaraldehyde fixation.  相似文献   

2.
Using a potato amylose fraction of 8 × 105, molecular-weight viscosity studies were carried out at 25°C on solutions containing 0.176–0.042% polymer, 8.67 mM KI, 1% ethanol, and different concentrations of iodine. By a novel extrapolation method, the intrinsic viscosities of the amylose/iodine complex were determined under various conditions of iodine binding (0–0.133 g I2/g amylose). Contrary to the view long held in this research area, it was found that the intrinsic viscosity of amylose solutions decreases significantly upon complex formation with iodine. Taking into account the results of our previous kinetic studies, the present findings are interpreted in terms of an amylose model characterized by loose, extended helical regions which are interrupted by short disordered regions. It is proposed that the intrinsic viscosity decrease observed is due to a shortening of the linear dimension of the polymer chain. This conformation change is apparently caused by the contraction of loose helical regions of the amylose macromolecule due to the entrapment of iodine (and perhaps other) atoms inside the helical cavities.  相似文献   

3.
R H Yun  A Anderson  J Hermans 《Proteins》1991,10(3):219-228
Free-energy simulations have been used to estimate the change in the conformational stability of short polyalanine alpha-helices when one of the alanines is replaced by a proline residue. For substituting proline in the middle of the helix the change in free energy of folding (delta delta G degrees) was calculated as 14 kJ/mol (3.4 kcal/mol), in excellent agreement with the one available experimental value. The helix containing proline was found to be strongly kinked; the free energy for reducing the angle of the kink from 40 degrees to 15 degrees was calculated, and found to be small. A tendency to alternate hydrogen bonding schemes was observed in the proline-containing helix. These observations for the oligopeptide agree well with the observation of a range of kink angles (18-35 degrees) and variety of hydrogen bonding schemes, in the rare instances where proline occurs in helices in globular proteins. For substituting proline at the N-terminus of the helix the change in free energy of folding (delta delta G degrees) was calculated as -4 kJ/mol in the first helical position (N1) and +6 kJ/mol in the second helical position (N2). The observed frequent occurrence of proline in position N1 in alpha-helices in proteins therefore has its origin in stability differences of secondary structure. The conclusion reached here that proline may be a better helix former in position N1 than (even) alanine, and thus be a helix initiator may be testable experimentally by measurements of fraction helical conformation of individual residues in oligopeptides of appropriate sequence. The relevance of these results in regards to the frequent occurrence of proline-containing helices in certain membrane proteins is discussed.  相似文献   

4.
The immunologically important (1 → 6) comb‐like branched (1 → 3)‐β‐D ‐glucans scleroglucan, schizophyllan, lentinan, and others, exist mainly as linear triple‐helical structures in aqueous solution. Partial interconversion from linear to circular topology has been reported to take place following conformational transition of the triple‐helical structure and subsequent regeneration of the triplex conformation. We here report on experimental data indicating that complete strand separation of the triple‐helical structure is required for this interconversion. NaOH or dimethylsulfoxide was used to induce dissociation of the triplex at combinations of concentrations and temperatures shown by calorimetry to yield a conformational transition of the triplex structures. For the alkaline treatment at 55°C, it is found that up to about 30% of the material readily can be converted to the cyclic topology. This fraction increased to about 60% when the subsequent annealing of the scleroglucan in aqueous solution at pH 7 was carried out at 100°C. Further increase of the annealing temperature yielded a smaller relative amount of cyclic species. The data indicate that the lower molecular weight fraction of the molecular weight distributions can be converted selectively to the macrocyclic topology by conditions that do not yield complete strand separation of the whole sample. These findings add to previous reports by providing more details about how the conditions required for the linear triplex to macrocycle interconversion relate to the conformational properties of the triple‐helical structure. © 1999 John Wiley & Sons, Inc. Biopoly 50: 496–512, 1999  相似文献   

5.
Conformational transitions induced by pH changes in random copolymers of leucine and glutamic acid have been studied. Significant differences were observed in the potentiometric titration curves of copolymers with small (up to 4%) and large leucine contents. The helical stability of copolymers with small leucine content, determined from titration curves by the Zimm and Rice method, decreases slightly with an increase in the leucine content, whereas the helical stability of copolymers with large leucine content increases sharply with an increase of the leucine content. It is shown that copolymers with large leucine content aggregate in the region of transition into the helical state, but the increase of their helical state stability is not connected with intermolecular aggregation, as it was also observed for a nonaggregating fraction isolated from one of the copolymers by gel chromatography. A conclusion is made that the helix–coil equilibrium constant s for leucine does not itself exceed the s constant for uncharged polyglutamic acid. The stabilization of the helical state in copolymers with large leucine content is due to intramolecular aggregation of helices in these copolymers. The analysis of the leucine residue distribution between helical and nonhelical regions in globular proteins also gives no real arguments to ascribe special helix-forming properties to leucine.  相似文献   

6.
M Hatano  M Yoneyama  Y Sato 《Biopolymers》1973,12(4):895-903
Interaction of poly-α,L -glutamic acid (PGLA) with acridine orange (AO) was studied with circular dichroism and absorption spectra measurements. The following results were observed: (1) the addition of a comparable amount of AO with the glutamyl residue to the PLGA solution at pH of 4.5 reduced the fraction of helix of the polymer; (2) when AO was added to the PGLA solution, the pH range of the helix-coil transition of the polymer shifted toward higher pH regions; and (3) when the mixture of the same amount of AO and the glutamyl residue was brought to the neutral and alkaline pH region, some induced circular dichroism bands were observed. In this case, it was assumed that PLGA in the system takes a helical form due to the neutralization of the anionized side chains by the cationic species of AO. We concluded that AO molecules bound to the carboxylate groups of the side chains of PLGA arrange to from a righthanded super-helix which surrounds the core of the righthanded α-helix of PLGA.  相似文献   

7.
The hexapositive complex cation ruthenium red very effectively stabilizes DNA and RNA double helices against thermal denaturation. In the presence of nucleic acid helices, this symmetric cation acquires an extrinsic CD spectrum near the wavelength of the dye's maximum absorbance. Competition experiments with single-stranded polyd(T) show this induced CD to be the result of selective binding to helical sites. The preferential affinity of ruthenium red for double helical binding sites is so great that it brings about biphasic absorbance- temperature profiles of polyd(A-T) at low [cation]: [polynucleotide phosphate]. The visible CD signal and fraction of helix melting at the upper transition increases with ruthenium red concentration until approximate charge neutrality is reached. These interactions, which have been studied in detail with the poly(U-U) helix as well as polyd(A-T), are likely largely electrostatic, since sufficient [NaCl] eliminates the bipliasic melting of polyd(A-T), renders the ultraviolet absorbance of poly (U) insensitive to ruthenium red, and abolishes the induced CD effects. The bipliasic melting of polyd(A-T) at intermediate [dye] is attributed to saturation of remaining double helical segments by cation migration from newly melted regions- Furthermore, virtually no change was observed in the induced CD upon melting through the first transition, whereas the effect is destroyed upon inciting through the second transition. A quantitative treatment of the data is used to obtain binding site size and association constant for the complex. The induced effect may prove useful in the exploration of exposed nucleic acid helical structure in such complex particles as nucleosomes or ribosomes.  相似文献   

8.
A significant fraction of the amino acids in proteins are alpha helical in conformation. Alpha helices in globular proteins are short, with an average length of about twelve residues, so that residues at the ends of helices make up an important fraction of all helical residues. In the middle of a helix, H-bonds connect the NH and CO groups of each residue to partners four residues along the chain. At the ends of a helix, the H-bond potential of the main chain remains unfulfilled, and helix capping interactions involving bonds from polar side chains to the NH or CO of the backbone have been proposed and detected. In a study of synthetic helical peptides, we have found that the sequence Ser-Glu-Asp-Glu stabilizes the alpha helix in a series of helical peptides with consensus sequences. Following the report by Harper and Rose, which identifies SerXaaXaaGlu as a member of a class of common motifs at the N termini of alpha helices in proteins that they refer to as “capping boxes,” we have reexamined the side chain–main chain interactions in a varient sequence using 1H NMR, and find that the postulated reciprocal side chain-backbone bonding between the first Ser and last Glu side chains and their peptide NH partners can be resolved: Deletion of two residues N terminal to the Ser-Glu-Asp-Glu sequence in these peptides has no effect on the initiation of helical structure, as defined by two-dimensional (2D) NMR experiments on this variant. Thus the capping box sequence Ser-Glu-Asp-Glu inhibits N terminal fraying of the N terminus of alpha helix in these peptides, and shows the side chain–main chain interactions proposed by Harper and Rose. It thus acts as a helix initiating signal. Since normal a helix cannot propagate beyond the N terminus of this structure, the box acts as a termination signal in this direction as well. © 1994 John Wiley & Sons, Inc.  相似文献   

9.
P G Scott  A Veis  G Mechanic 《Biochemistry》1976,15(15):3191-3198
A peptide fraction isolated from a cyanogen bromide digest of bovine dentin collagen had a molecular weight of 46000. Its size and amino acid composition indicated that it could not consist of peptides derived from the cleavage of a single alpha chain. On reduction with tritiated sodium borohydride, radioactivity was incorporated primarily into 5, 5'-dihydroxylysinonorleucine without degradation at the peptide backbone. Periodate cleavage of the reduced or nonreduced peptide fraction generated one fragment of molecular weight 28000 and one of 18000 completely accounting for the size of the parent peptide. On amino acid analysis the constituent single-chain peptides were determined to be alpha2CB4 and alpha1CB6. Both peptides isolated after periodate oxidation of the tritiated borohydride reduced cross-link peptide were found to contain (3H)hydroxynorvaline. These data show that some hydroxylysine of alpha2CB4, a helical region peptide, was present in aldehyde form and could act as the aldehyde donor icross-link, Schiff's base formation. The only cross-linkage of this alpha2CB4 acting as an aldehyde donor peptide to alpha1CB6 would be a helical region to helical region bond, perhaps accounting for the unusual stability and low solubility of dentin collagen.  相似文献   

10.
11.
Oriented circular dichroism (OCD) was used to characterize and compare in a quantitative manner the secondary structure and concentration dependent realignment of the antimicrobial peptides PGLa and MSI-103, and of the structurally related cell-penetrating peptide MAP in aligned phospholipid bilayers. All these peptides adopt an amphiphilic α-helical conformation, and from solid-state NMR analysis they are known to bind to membranes in two distinct orientations depending on their concentration. At low peptide/lipid (P/L) ratio the helices are aligned parallel to membrane surface (S-state), but with increasing concentration they realign to a tilted orientation (T-state), getting immersed into the membrane with an oblique angle supposedly as a result of dimer-formation. In macroscopically aligned liquid crystalline 1,2-dimyristoyl-sn-glycero-3-phosphatidylcholine bilayers the two limiting states are represented by distinct OCD spectra, and all spectra at intermediate peptide concentrations can be described by a linear combination of these two line shapes. The corresponding fraction of molecules occupying the T-state was determined by fitting the intermediate spectra with a superposition of the two extreme line shapes. By plotting this fraction versus 1/(P/L), the threshold P/L* ratio for realignment was extracted for each of the three related peptides. Despite their structural similarity distinctly different thresholds were obtained, namely for MSI-103 realignment starts already at a low P/L of ∼1:236, for a MAP derivative (using a nonaggregating analog containing a D-amino acid) the transition begins at P/L ∼1:156, whereas PGLa needs the highest concentration to flip into T-state at P/L ∼1:85. Analysis of the original MAP sequence (containing only L-amino acids) gave OCD spectra compatible with β-pleated conformation, suggesting that this peptide starts to aggregate with increasing concentration, unlike the other helical peptides. All these changes in peptide conformation and membrane alignment observed here by OCD seem to be functionally relevant, as they can be correlated with the membrane perturbing activities of the three antimicrobial and cell-penetrating sequences.  相似文献   

12.
The protozoan parasite Entamoeba histolytica is an ancient eukaryotic cell that shows morphologically atypical organelles and differs metabolically from higher eukaryotic cells. The aim of this study was to determine the subcellular localization of ameba NAD+-dependent alcohol dehydrogenase (ADH2). The enzyme activity was present in soluble and mainly in particulate material whose density was 1.105 in a sucrose gradient. By differential centrifugation, most of the ADH activity sedimented at 160,000 g (160,000-g pellet), similar to the Escherichia coli polymeric ADHE. In the Coomassie staining of the 160,000-g pellet analyzed by electrophoresis, a 96-kDa protein was more prominent than in other fractions; this band was recognized by antibodies against Lactococcus lactis ADHE. By gold labeling, the antibodies recognized the granular material that mainly constitutes the 160,000-g pellet and a material that sedimented along with the internal membrane vesicles. By negative staining, the 160,000-g fraction showed helical rodlike structures with an average length of 103 nm; almost no membrane vesicles were observed in this pellet. In internal membrane fractions, no rodlike structures were found, but protomerlike round structures were observed. These results indicate that the main amebic NAD+-dependent ADH2 activity is naturally organized as rodlike helical particles, similar to bacterial ADHE. Detection of ADH2 in membrane fractions might be explained by cosedimentation of the multimeric ADH during membrane purification.  相似文献   

13.
Employing high-resolution (13)C solution NMR and circular dichroism (CD) spectroscopic techniques, the distinctive influence of two intimately related hexafluoro solvents, 1,1,1,3,3,3-hexafluoro-2-propanol (HFIP) and hexafluoroacetone trihydrate (HFA), on the structural characteristics of Bombyx mori (B. mori) silk fibroin, the chymotrypsin precipitate (C(p)) fraction, and two synthetic peptides, (AGSGAG)(5) and (AG)(15), is described. The observed (13)C solution NMR and CD spectra of these polypeptides in HFIP and HFA revealed a distinctive influence on their conformational characteristics. The (13)C NMR spectra, as analyzed from the unique chemical shifts of C(alpha) and C(beta) resonances of constituent residues revealed that fibroin largely assumes helical conformation(s) in both solvents. However, the peak shifts were greater for the samples in HFIP, indicating that the types of helical structure(s) may be different from the one populated in HFA. Similar structural tendencies of these polypeptides were reflected in CD spectra. The observed CD patterns, i.e., a strong positive band at approximately 190 nm and negative bands at approximately 206 and 222 nm, have been attributed to the preponderance of helical structures. Of the two prevalent helical structures, alpha-helix and 3(10)-helix, the evidence emerged for the fibroin protein in favor of 3(10)-helical structure stabilization in HFIP and its significant disruption in HFA, as deduced from the characteristic R1 (=[theta](190)/[theta](202)) and R2 (=[theta](222)/[theta](206)) ratios, determined from the CD data. Conversely, the native polypeptides and synthetic peptide fragments derived from highly crystalline regions of the silk fibroin protein sustained predominantly an unordered structure in HFA solvent.  相似文献   

14.
A number of highly curved membranes in vivo, such as epithelial cell microvilli, have the relatively high sphingolipid content associated with “raft-like” composition. Given the much lower bending energy measured for bilayers with “nonraft” low sphingomyelin and low cholesterol content, observing high curvature for presumably more rigid compositions seems counterintuitive. To understand this behavior, we measured membrane rigidity by fluctuation analysis of giant unilamellar vesicles. We found that including a transmembrane helical GWALP peptide increases the membrane bending modulus of the liquid-disordered (Ld) phase. We observed this increase at both low-cholesterol fraction and higher, more physiological cholesterol fraction. We find that simplified, commonly used Ld and liquid-ordered (Lo) phases are not representative of those that coexist. When Ld and Lo phases coexist, GWALP peptide favors the Ld phase with a partition coefficient of 3–10 depending on mixture composition. In model membranes at high cholesterol fractions, Ld phases with GWALP have greater bending moduli than the Lo phase that would coexist.  相似文献   

15.
E W Small  W L Peticolas 《Biopolymers》1971,10(8):1377-1418
Raman spectra are presented on ordered and presumably helical structures of DNA and RNA as well as the poly A·poly U helical complex, polydAT, and the helical aggregates of 5′-GMP and 3′-GMP. The changes in the frequency and the intensity of the Raman bands as these structures undergo order-disorder transitions have been measured. In general the changes we have found can be placed into three categories: (1) A reduction in the intensities of certain ring vibrations of the polynucleotide bases is observed when stacking or ordering occurs (Raman hypochromism). Since the ring vibrational frequencies are different for each type of base, we have been able to obtain some estimate of average amount of order of each type of base in partially ordered helical systems. (2) A very large increase in the intensity of a sharp, strongly polarized band at about 815 cm?1 is observed when polyriboA and polyriboU are formed into a helical complex. Although this band is not present in the separated chains at high temperature, a broad diffuse band at about 800 cm?1 is present. The 815 cm?1 band undoubtedly arises from the vibrations of the phosphate-sugar portions of the molecule and provides a sensitive handle to the back-bone conformation of the polymer. This band also appears upon ordering of RNA, formation of the helical aggregate of 5′-riboGMP, and to some extent in the selfstacking of the polyribonucleotides polyA, polyU in the presence of Mg++, PolyC, and polyG. No such intense, polarized band is found, however, in ordered DNA, polydAT, or the 3′-riboGMP aggregate, although there is a conformationally independent band at about 795 cm?1 in DNA and polydAT. (3) Numerous frequency changes occur during Conformational changes. In particular the 1600–1700 cm?1 region in D2O shows significant conformationally dependent changes in the C?O stretching region analogous to the changes in this region which have been observed in these substances in the infrared. Thus, Raman scattering appears to provide a technique for simultaneously observing the effects of base stacking, backbone conformation and carbonyl hydrogen bonding in nucleic acids in moderately dilute (10–25 mg/ml) aqueous solutions.  相似文献   

16.
In amyloidosis associated with apolipoprotein A-I (ApoA-I), heart amyloid deposits are mainly constituted by the 93-residue ApoA-I N-terminal region. A recombinant form of the amyloidogenic polypeptide, named [1-93]ApoA-I, shares conformational properties and aggregation propensity with its natural counterpart. The polypeptide, predominantly in a random coil state at pH 8.0, following acidification to pH 4.0 adopts a helical/molten globule transient state, which leads to formation of aggregates. Here we provide evidence that fibrillogenesis occurs also in physiologic-like conditions. At pH 6.4, [1-93]ApoA-I was found to assume predominantly an α-helical state, which undergoes aggregation at 37°C over time at a lower rate than at pH 4.0. After 7 days at pH 6.4, protofibrils were observed by atomic force microscopy (AFM). Using a multidisciplinary approach, including circular dichroism (CD), fluorescence, electrophoretic, and AFM analyses, we investigated the effects of a lipid environment on the conformational state and aggregation propensity of [1-93]ApoA-I. Following addition of the lipid-mimicking detergent Triton X-100, the polypeptide was found to be in a helical state at both pH 8.0 and 6.4, with no conformational transition occurring upon acidification. These helical conformers are stable and do not generate aggregated species, as observed by AFM after 21 days. Similarly, analyses of the effects of cholesterol demonstrated that this natural ApoA-I ligand induces formation of α-helix at physiological concentrations at both pH 8.0 and 6.4. Zwitterionic, positively charged, and negatively charged liposomes were found to affect [1-93]ApoA-I conformation, inducing helical species. Our data support the idea that lipids play a key role in [1-93]ApoA-I aggregation in vivo.  相似文献   

17.
Domain 1 of the cell adhesion protein CD2 (CD2-1) has an all β-structure typical of proteins belonging to the immunoglobin superfamily. It has a remarkable, ability to fold as a native monomer or a metastable intertwined dimer. To understand the origin of structural rearrangements of CD2-1, we have studied equilibrium unfolding of the protein using various biophysical spectroscopic techniques. At temperatures above approx 68°C, a partially folded state of CD2-1 (H state) with a distinct secondary structure, involving largely exposed aromatic and hydrophobic residues and a substantially perturbed tertiary structure, is observed. In contrast, an unfolded state (D state) of CD2-1 with random-coil-like secondary and tertiary structures is observed in 6 M GuHCl. This partially folded high-temperature state has increased negative molar ellipticity at 222 nm in far-ultraviolet CD spectra, implying formation of a non-native helical conformation. The existence of this non-native high-temperature intermediate is consistent with relatively high intrinsic helical propensities in the primary sequence of CD2-1. This conformation flexibility may be important in the observed domain swapping of CD2-1.  相似文献   

18.
The addition of small amounts of multivalent cations to solutions containing double-stranded DNA leads to inter-DNA attraction and eventual condensation. Surprisingly, the condensation is suppressed in double-stranded RNA, which carries the same negative charge as DNA, but assumes a different double helical form. Here, we combine experiment and atomistic simulations to propose a mechanism that explains the variations in condensation of short (25 base-pairs) nucleic acid (NA) duplexes, from B-like form of homopolymeric DNA, to mixed sequence DNA, to DNA:RNA hybrid, to A-like RNA. Circular dichroism measurements suggest that duplex helical geometry is not the fundamental property that ultimately determines the observed differences in condensation. Instead, these differences are governed by the spatial variation of cobalt hexammine (CoHex) binding to NA. There are two major NA-CoHex binding modes—internal and external—distinguished by the proximity of bound CoHex to the helical axis. We find a significant difference, up to 5-fold, in the fraction of ions bound to the external surfaces of the different NA constructs studied. NA condensation propensity is determined by the fraction of CoHex ions in the external binding mode.  相似文献   

19.
We characterized senile plaques (SPs) immunohistochemically in cynomolgus monkey brains and also examined age-related biochemical changes of Alzheimer's disease (AD)-associated proteins in these brains from monkeys of various ages. In the neocortex of aged monkeys (>20 years old), we found SPs but no neurofibrillary tangles (NFTs). Antibodies against beta-amyloid precursor protein (APP) or apolipoprotein E (ApoE) stained SPs; however, the pattern of immunostaining was different for the two antigens. APP was present only in swollen neurites, but ApoE was present throughout all parts of SPs. Western blot analysis revealed that the pattern of APP expression changed with age. Although full-length APP695 protein was mainly expressed in brains from young monkeys (4-years-old), the expression of full-length APP751 protein was increased in brains from older monkeys (>20 years old). Biochemical analyses also showed that levels of various AD-associated proteins increased significantly with age in nerve ending fractions. Both SP-associated (APP) and NFT-associated proteins (tau, activated glycogen synthase kinase 3beta, cyclin dependent kinase 5, p35, and p25) accumulated in the nerve ending fraction with increasing age; however, we found no NFTs or paired helical filaments of tau in aged cynomolgus monkey brains. This age-related accumulation of these proteins in the nerve ending fraction was similar to that observed in our laboratory previously for presenilin-1 (PS-1). The accumulation of these SP-associated proteins in this fraction may be a causal event in the spontaneous formation of SPs; thus, SPs may be formed initially in nerve endings. Taken together, these results suggest that intensive investigation of age-related changes in the nerve ending and in axonal transport will contribute to a better understanding of the pathogenesis of neurodegenerative disorders such as AD.  相似文献   

20.
The optical rotation phenomena exhibited by a citrate-extracted fraction of ichthyocol (from carp swim bladder), as well as by the parent gelatin derived therefrom, have been studied. Dispersion data for all cases follow a single-term Drude equation, but the variations with state are adequately expressed by simple reference to changes in [alpha](D) as follows:- 1. The native collagen fraction, dispersed in 0.15 M citrate buffer at pH 3.7 in the cold (11 degrees C.), yields a high negative specific rotation, [alpha](D), near -350 degrees . 2. During equilibration at 40 degrees C., which causes conversion to a monodisperse parent gelatin, the rotation drops to about -110 degrees . 3. Gelation at 2 degrees C. results in a partial regain of rotation to around -290 degrees . This mutarotation is reversible, depending on temperature. 4. In the range 0.02 to 0.28 per cent the native ichthyocol and the warm gelatin solutions show little concentration dependence, but with the cold gelatin solutions the specific rotation increases with concentration. Gelatin films formed by cold evaporation yield high specific rotation (ca. -620 degrees ), but those formed by hot evaporation retain low optical activity. 5. Since this same collagen-gelatin system has been investigated physicochemically, it is possible to relate molecular changes to the observed variations in optical rotation. Conclusions are similar to those of Robinson (1953), who studied other gelatins: high negative rotation is believed related to a native collagen polypeptide configuration, herein specified as helical (from x-ray diffraction considerations) and destroyed by heating. The possible roles of intermolecular interactions and of prevalent pyrrolidine constituents in influencing the helical configuration and optical activity are discussed.  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