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.
The secondary structure of the purified 70-kDa protein Na+/Ca2+ exchanger, functionally reconstituted into asolectin lipid vesicles, was examined by Fourier transform infrared attenuated total reflection spectroscopy. Fourier transform infrared attenuated total reflection spectroscopy provided evidence that the protein is composed of 44% alpha-helices, 25% beta-sheets, 16% beta-turns, and 15% random structures, notably the proportion of alpha-helices is greater than that corresponding to the transmembrane domains predicted by exchanger hydropathy profile. Polarized infrared spectroscopy showed that the orientation of helices is almost perpendicular to the membrane. Tertiary structure modifications, induced by addition of Ca2+, were evaluated by deuterium/hydrogen exchange kinetic measurements for the reconstituted exchanger. This approach was previously proven as a useful tool for detection of tertiary structure modifications induced by an interaction between a protein and its specific ligand. Deuterium/hydrogen exchange kinetic measurements indicated that, in the absence of Ca2+, a large fraction of the protein (40%) is inaccessible to solvent. Addition of Ca2+ increased to 55% the inaccessibility to solvent, representing a major conformational change characterized by the shielding of at least 93 amino acids.  相似文献   

2.
Calcareous otoliths in the inner ears of fishes are necessary for proper hearing and vestibular function. Sagittal otoliths are usually composed of the calcium carbonate polymorph aragonite but may contain the polymorph vaterite, a phenomenon called otolith crystallization. The causes of otolith crystallization are poorly understood. Thyroid hormone (TH) can influence the chemical microenvironment and structure of the inner ear, suggesting that TH may influence otolith crystallization. The present study examined the effect of exogenous TH treatment on sagittal otolith crystallization and growth in larval and juvenile rainbow trout, Oncorhynchus mykiss. In the first experiment, 110?C179?day-old fish raised from TH-treated oocytes had significantly fewer sagittal otoliths containing the crystalline calcium carbonate polymorph vaterite as compared to untreated fish. Vaterite-containing otoliths were significantly longer than those containing the typical polymorph aragonite, although there was no effect of TH treatment on otolith length. In the second experiment, juveniles immersed in an exogenous solution of TH for 6?weeks had slightly longer otoliths (relative to fish length) than age-matched controls, but this effect was not significant. This juvenile population had a very high percentage (88.3?%) of vaterite sagittae overall and this percentage did not change significantly with treatment, suggesting the switch from aragonite to vaterite occurred prior to inclusion of the fish in the study. These results suggest that early manipulation of TH levels may affect calcium carbonate deposition on the otolith but that later TH exposure is unable to restore typical otolith composition.  相似文献   

3.
A new method for finding hydrophobic nuclei and microclusters in protein structure is proposed. The method uses simple and clear-cut criteria based on an analysis of distances between the hydrocarbon groups of all residues. A detailed analysis of the composition and properties of hydrophobic nucleic and microclusters for proteins of different types has been carried out. This approach reveals that a hydrophobic nucleus can be composed not merely of classical hydrophobic amino acids, but also of dicarboxylic acids, their amides, arginine, lysine, histidine and tyrosine. The hydrophobic nucleus defined by this method should be considered as an individual structural unit along with such elements of the secondary structure as alpha-helices, beta-turns and beta-sheets.  相似文献   

4.
Secondary structure and membrane topology of cytochrome P450s   总被引:1,自引:0,他引:1  
The secondary structure prediction of 19 microsomal cytochrome P450s from two different families was made on the basis of their amino acid sequences. It was shown that there is structural similarity between the heme-binding sites in these enzymes and those in the bacterial P450cam. An average predicted secondary structure of cytochrome P450 proteins with 70% accuracy contains about 46% alpha-helices, 12% beta-sheets, 9% beta-turns, and 33% random coils. In the region of residues 35-120 in microsomal P450s two adjacent beta alpha beta-units (the Rossmann domain), were recognized and may be available to interact with the NADPH-cytochrome P450 reductase. Using the procedure for identification of hydrophobic and membrane-associated alpha-helical segments, only one N-terminal transmembrane anchor was predicted. Also the heme-binding site may include the surface-bound helix. A model for vertebrate microsomal P450s having an amphipathic membrane protein located on the cytoplasmic side of the endoplasmic reticulum membrane, with their active center lying outside or on the bilayer border, is proposed.  相似文献   

5.
Yu P 《Biopolymers》2007,85(4):308-317
Barley protein structure affects the barley quality, fermentation, and degradation behavior in both humans and animals among other factors such as protein matrix. Publications show various biological differences among barley varieties such as Valier and Harrington, which have significantly different degradation behaviors. The objectives of this study were to reveal the molecular structure of barley protein, comparing various varieties (Dolly, Valier, Harrington, LP955, AC Metcalfe, and Sisler), and quantify protein structure profiles using Gaussian and Lorentzian methods of multi-component peak modeling by using the ultra-spatially resolved synchrotron light sourced Fourier transform infrared microspectroscopy (SFTIRM). The items of the protein molecular structure revealed included protein structure alpha-helices, beta-sheets, and others such as beta-turns and random coils. The experiment was performed at the National Synchrotron Light Source in Brookhaven National Laboratory (BNL, US Department of Energy, NY). The results showed that with the SFTIRM, the molecular structure of barley protein could be revealed. Barley protein structures exhibited significant differences among the varieties in terms of proportion and ratio of model-fitted alpha-helices, beta-sheets, and others. By using multi-component peaks modeling at protein amide I region of 1710-1576 cm-1, the results show that barley protein consisted of approximately 18-34% of alpha-helices, 14-25% of beta-sheets, and 44-69% others. AC Metcalfe, Sisler, and LP955 consisted of higher (P<0.05) proportions of alpha-helices (30-34%) than Dolly and Valier (alpha-helices 18-23%). Harrington was in between which was 25%. For protein beta-sheets, AC Metcalfe, and LP955 consisted of higher proportions (22-25%) than Dolly and Valier (13-17%). Different barley varieties contained different alpha-helix to beta-sheet ratios, ranging from 1.4 to 2.0, although the difference were insignificant (P>0.05). The ratio of alpha-helices to others (0.3 to 1.0, P<0.05) and that of beta-sheets to others (0.2 to 0.8, P<0.05) were different among the barley varieties. It needs to be pointed out that using a multi-peak modeling for protein structure analysis is only for making relative estimates and not exact determinations and only for the comparison purpose between varieties. The principal component analysis showed that protein amide I Fourier self-deconvolution spectra were different among the barley varieties, indicating that protein internal molecular structure differed. The above results demonstrate the potential of the SFTIRM to localize relatively pure protein areas in barley tissues and reveal protein molecular structure. The results indicated relative differences in protein structures among the barley varieties, which may partly explain the biological differences among the barley varieties. Further study is needed to understand the relationship between barley molecular chemical structure and biological features in terms of nutrient availability and digestive behavior.  相似文献   

6.
We have analysed the hydration of main-chain carbonyl and amide groups in 24 high-resolution well-refined protein structures as a function of the secondary structure in which these polar groups occur. We find that main-chain atoms in beta-sheets are as hydrated as those in alpha-helices, with most interactions involving "free" amide and carbonyl groups that do not participate in secondary structure hydrogen bonds. The distributions of water molecules around these non-bonded carbonyl groups reflect specific steric interactions due to the local secondary structure. Approximately 20% and 4%, respectively of bonded carbonyl and amide groups interact with solvent. These include interactions with carbonyl groups on the exposed faces of alpha-helices that have been correlated previously with bending of the helix. Water molecules interacting with alpha-helices occur mainly at the amino and carbonyl termini of the helices, in which case the solvent sites maintain the hydrogen bonding by bridging between residues i and i-3 or i-4 at the amino terminus and between i and i+3 or i+4 at the carbonyl terminus. We also see a number of solvent-mediated Ncap and Ccap interactions. The water molecules interacting with beta-sheets occur mainly at the edges, in which case they extend the sheet structure, or at the ends of strands, in which case they extend the beta-ladder. In summary, the solvent networks appear to extend the hydrogen-bonding structure of the secondary structures. In beta-turns, which usually occur at the surface of a protein, exposed amide and carbonyl groups are often hydrated, especially close to glycine residues. Occasionally water molecules form a bridge between residues i and i+3 in the turn and this may provide extra stabilization.  相似文献   

7.
Secondary and tertiary structures of human blood alpha(1)-acid glycoprotein, a member of the lipocalin family, have been studied for the first time by infrared and Raman spectroscopies. Vibrational spectroscopy confirmed details of the secondary structure and the structure content predicted by homology modeling of the protein moiety, i.e., 15% alpha-helices, 41% beta-sheets, 12% beta-turns, 8% bands, and 24% unordered structure at pH 7.4. Our model shows that the protein folds as a highly symmetrical all-beta protein dominated by a single eight-stranded antiparallel beta-sheet. Thermal dynamics in the range 20-70 degrees C followed by Raman spectroscopy and analyzed by principle component analysis revealed full reversibility of the protein motion upon heating dominated by decreasing of beta-sheets. Raman difference spectroscopy confirmed the proximity of Trp(122) to progesterone binding.  相似文献   

8.
One of the functions of MutY from Escherchia coli is removal of adenine mispaired with 7,8-dihydro-8-oxoguanine (8-oxoG), a common lesion in oxidatively damaged DNA. MutY is composed of two domains: the larger N-terminal domain (p26) contains the catalytic properties of the enzyme while the C-terminal domain (p13) affects substrate recognition and enzyme turnover. On the basis of sequence analyses, it has been recently suggested that the C-terminal domain is distantly related to MutT, a dNTPase which hydrolyzes 8-oxo-dGTP [Noll et al. (1999) Biochemistry 38, 6374-6379]. We have studied the solution structure of the C-terminal domain of MutY by NMR and find striking similarity with the reported solution structure of MutT. Despite low sequence identity between the two proteins, they have similar secondary structure and topology. The C-terminal domain of MutY is composed of two alpha-helices and five beta-strands. The NOESY data indicate that the protein has two beta-sheets. MutT is also a mixed alpha/beta protein with two helices and two beta-sheets composed of five strands. The secondary structure elements are similarly arranged in the two proteins.  相似文献   

9.
W Z He  W R Newell  P I Haris  D Chapman  J Barber 《Biochemistry》1991,30(18):4552-4559
The secondary structure of the photosystem II (PSII) reaction center isolated from pea chloroplasts has been characterized by Fourier transform infrared (FTIR) spectroscopy. Spectra were recorded in aqueous buffers containing H2O or D2O; the detergent present for most measurements was dodecyl maltoside. The broad amide I and amide II bands were analyzed by using second-derivative and deconvolution procedures. Absorption bands were assigned to the presence of alpha-helices, beta-sheets, turns, or random structure. Quantitative analysis revealed that this complex contained a high proportion of alpha-helices (67%) and some antiparallel beta-sheets (9%) and turns (11%). An irreversible decrease in the intensity of the band associated with the alpha-helices occurs upon exposure of the isolated PSII reaction center to bright illumination. This loss of alpha-helical content gave rise to an increase in other secondary structures, particularly beta-sheets. After similar pretreatment with light, sodium dodecyl sulfate polyacrylamide gel electrophoresis reveals lower mobility and solubility of constituent D1 and D2 polypeptides of the PSII reaction center. Some degradation of these polypeptides also occurs. In contrast, there is no change in the mobility of the two subunits of cytochrome b559. In the absence of illumination, the PSII reaction center exchanged into dodecyl maltoside shows good thermal stability as compared with samples in Triton X-100. Only at a temperature of about 60 degrees C do spectral changes take place that are indicative of denaturation.  相似文献   

10.
We developed an FTIR (Fourier transform infrared) methodology for quantitatively assessing the secondary structure of proteins suspended in nonaqueous media. This methodology was used to measure the percentages of alpha-helices and beta-sheets of subtilisin Carlsberg, prepared under different conditions, placed in various organic solvents. The title question was addressed with respect to some instances of markedly influencing the subtilisin activity in organic solvents reported in the literature. It is concluded that the mechanism of subtilisin activation by KCl and N-Ac-L-Phe-NH(2) present in the aqueous solution of the enzyme prior to lyophilization may be due to their preservation of the secondary structure, otherwise altered by the dehydration. Likewise, subtilisin inactivation in the protein-dissolving solvent DMSO (dimethyl sulfoxide) is likely caused by enzyme denaturation (the loss of both alpha-helices and beta-sheets). On the other hand, some other ligands, as well as protein nondissolving organic solvents, while greatly affecting the subtilisin activity, have little effect on its secondary structure, thus ruling out the causal relationship between the two. (c) 1997 John Wiley & Sons, Inc. Biotechnol Bioeng 53: 351-362, 1997.  相似文献   

11.
The 120-kDa Na+/Ca2+ exchanger was purified and reconstituted into lipid vesicles. The secondary structure composition of the exchanger was 39% alpha-helices, 20% beta-sheets, 25% beta-turns, and 16% random coils, as analyzed by Fourier transform infrared attenuated total reflection spectroscopy. The secondary structure composition of the COOH-terminal portion of the protein was compatible with a topology model containing 4-6 transmembrane segments. Furthermore, the secondary structure of the NH2-terminal portion of the cytoplasmic loop was analyzed and found to be different from that of the COOH-terminal portion. Ca2+ and/or the exchange inhibitory peptide (XIP) failed to affect the secondary structure of the 120-kDa protein. Tertiary structure modifications induced by Ca2+ and XIP were analyzed by monitoring the hydrogen/deuterium exchange rate for the reconstituted exchanger. In the absence of ligand, 51% of the protein was accessible to solvent. Ca2+ decreased accessibility to 40%, implicating the shielding of at least 103 amino acids. When both Ca2+ and XIP were added, accessibility increased to 66%. No modification was obtained when XIP was added alone. Likewise, in the presence of Ca2+, XIP failed to modify the tertiary structure of the 70-kDa protein, suggesting that XIP acts at the level of the COOH-terminal portion of the intracellular loop. The present data describe, for the first time, conformational changes of the Na+/Ca2+ exchanger induced by Ca2+ and XIP, compatible with an interaction model where regulatory Ca2+ and inhibitory XIP bind to distinct sites, and where XIP binding requires the presence of Ca2+.  相似文献   

12.
J Cladera  M Sabés  E Padrós 《Biochemistry》1992,31(49):12363-12368
Fourier transform infrared spectroscopy at a resolution of 1 cm-1 has been used to study the conformation of dark-adapted bacteriorhodopsin in the native purple membrane, in H2O and D2O suspensions. A detailed analysis of the amide I bands was made using derivative and deconvolution techniques. Curve-fitting results of four independent experiments indicate, after estimation of the methodological errors, that native bacteriorhodopsin contains 52-73% alpha-helices, 13-19% reverse turns, 11-16% beta-sheets, and 3-7% unordered segments. Our analysis has enabled the identification of several components corresponding to alpha-helices, beta-sheets, and reverse turns. Besides the alpha I- and alpha II-helices (peaking at 1658 and 1665 cm-1), we propose that two more infrared bands arise from alpha-helical structures: one at 1650 cm-1 from alpha I and another one at 1642 cm-1 in H2O suspension, which could originate from type III beta-turns (i.e., one turn of 3(10)-helix). The relatively high content of reverse turns suggests the presence of one reverse turn per loop, plus another one in the C-terminal segment. On the other hand, several reasons argue that the calculated mean beta-sheet content of around 14% should be decreased somewhat. These beta-sheets could be located in the noncytoplasmatic links of the bacteriorhodopsin molecule.  相似文献   

13.
Huang JT  Cheng JP  Chen H 《Proteins》2007,67(1):12-17
We present a simple method for determining the folding rates of two- and three-state proteins from the number of residues in their secondary structures (secondary structure length). The method is based on the hypothesis that two- and three-state foldings share a common pattern. Three-state proteins first condense into metastable intermediates, subsequent forming of alpha-helices, turns, and beta-sheets at slow rate-limiting step. The folding rate of such proteins anticorrelate with the length of these beta-secondary structures. It is also assumed that in two-state folding, rapidly folded alpha-helices and turns may facilitate formation of fleeting unobservable intermediates and thus show two-state behavior. There is an inverse relationship between the folding rate and the length of beta-sheets and loops. Our study achieves 94.0 and 88.1% correlations with folding rates determined experimentally for 21 three- and 38 two-state proteins, respectively, suggesting that protein-folding rates are determined by the secondary structure length. The kinetic kinds are selected on the basis of a competitive formation of hydrophobic collapse and alpha-structure in early intermediates.  相似文献   

14.
The secondary structure prediction of 19 microsomal cytochrome P-450s from two different families was made based on their amino acid sequences. It was shown that there is a structural similarity between the heme-binding sites of these enzymes and the bacterial P-450cam. An average predicted secondary structure of cytochrome P-450 proteins with 70% accuracy contains about 46% alpha-helices, 12% beta-strands, 9% beta-turns and 33% random coil. In the region of the 35-120 residues in microsomal P-450s two adjacent beta alpha beta-units (the Rossmann domain) were recognized, which may interact with the NADPH-cytochrome P-450 reductase. Using the procedure of identification of hydrophobic and membrane-associated alpha-helical segments of 23 cytochromes, only one N-terminal transmembrane anchor was predicted. Also the heme-binding site perhaps includes surface-bound helix. A model of vertebrate microsomal P-450s is proposed. That is an amphypathic membrane protein located on the cytoplasmic face of the endoplasmic reticulum, their active center lies out/on the bilayer border.  相似文献   

15.
Conformations of polyomavirus (Py) major capsid protein VP1 were analyzed by circular dichroism (CD) and fluorescence spectroscopy in the presence of sodium dodecyl sulfate (SDS). Binding of PyVP1 to SDS induced marked conformational changes of PyVP1, which were reflected on the CD and fluorescence spectra. Abrupt changes in both optical properties occurred within the narrow ranges of SDS concentrations with the transition midpoints closely related to SDS micelle formation. Analysis of circular dichroism spectra showed that the contents of alpha-helices, beta-sheets, beta-turns and random coils in PyVP1 varied upon addition of SDS, demonstrating the exquisite sensitivity of the conformations of the protein to the environment. The interactions of PyVP1 with SDS were shown to be dependent on the ionic strength of the protein solution, suggesting that both hydrophobic and electrostatic forces contribute to the PyVP1-SDS complex formation. The SDS-induced conformational changes of PyVP1 appeared to be a two-stage process.  相似文献   

16.
Taking advantage of recent advances in polarized Raman microspectroscopy, and based on a rational decomposition of the amide I band, the conformation and orientation of proteins have been determined for cocoon silks of the silkworms Bombyx mori and Samia cynthia ricini and dragline silks of the spiders Nephila clavipes and Nephila edulis. This study distinguished between band components due to beta-sheets, beta-turns, 3(1)-helices, and unordered structure for the four fibers. For B. mori, the beta-sheet content is 50%, which matches the proportion of residues that form the GAGAGS fibroin motifs. For the Nephila dragline and S. c. ricini cocoon, the beta-sheet content (36-37% and 45%, respectively) is higher than the proportion of residues that belong to polyalanine blocks (18% and 42%, respectively), showing that adjacent GGA motifs are incorporated into the beta-sheets. Nephila spidroins contain fewer beta-sheets and more flexible secondary structures than silkworm fibroins. The amorphous polypeptide chains are preferentially aligned parallel to the fiber direction, although their level of orientation is much lower than that of beta-sheets. Overall, the results show that the four silks exhibit a common molecular organization, with mixtures of different amounts of beta-sheets and flexible structures, which are organized with specific orientation levels.  相似文献   

17.
To explore the secondary structures of lingual and pancreatic lipases, circular dichroism measurements were performed. Maximum average ellipticities were used to calculate the percentage of alpha-helices, beta-sheets, and random coils. Lingual lipase had an ellipticity of -20235 +/- 140 deg cm2/dmol (mean +/- SE) at 220 nm suggesting 60% alpha-helix, 20% beta-sheet and 20% random coil structure, but the mean ellipticity for pancreatic lipase was -14093 +/- 82 deg cm2/dmol (mean +/- SE) at 210 nm suggesting a 34.8% alpha-helical, 25% beta-sheet and 40% random coil secondary structure. An alpha-helical stretch of residues with a large hydrophobic moment ("globular" alpha-helix by hydrophobic moment plot) from amino acids 382 through 389 at the COOH-terminal end of lingual lipase was noted. This sequence, absent in pancreatic lipase, may account for the avid binding of lingual lipase to fat emulsion particles.  相似文献   

18.
The central domain is an evolutionarily conservative region that is invariant in length in the A and Hc-A families of silkmoth chorion proteins. This domain shows strong sixfold periodicities for various amino acid residues, such as glycine and large non-polar residues. The periodicities and their phase relationships, together with the documented prevalence of beta-sheets and beta-turns in the chorion, strongly support a secondary structure model in which short (4-residue) beta-sheet strands alternate with beta-turns, forming a compact antiparallel, probably twisted beta-sheet. This structure should be important for the establishment of higher order structure in the chorion.  相似文献   

19.
The creation of native-like macromolecules in copying nature's way represents a fascinating challenge in protein chemistry today. In the absence of a detailed knowledge of the complex folding pathway the ultimate goal in protein de novo design, the construction of artificial proteins with predetermined three-dimensional structure and tailor-made functions based on a defined, generally valid set of rules, appears to be still out of reach. With progress in synthesis strategies and biostructural characterization methods, topological templates have become a versatile tool for inducing and stabilizing secondary and tertiary structures, such as protein loops, beta-turns, alpha-helices, beta-sheets and a variety of folding motifs. In this article, we extend the concept of template-assembled synthetic proteins for the construction of protein-like topologies with multiply bridged, oligocyclic chain architectures termed locked-in tertiary folds that exhibit unique physicochemical and folding properties because of the highly confined conformational space. Furthermore, we show that some fundamental questions in protein assembly can be approached applying the template concept. Using covalent template trapping of self-associated peptide assemblies in aqueous solution the structural and physical forces guiding protein folding, supramolecular assembly and molecular recognition processes can be studied on a molecular level.  相似文献   

20.
The secondary structure for two murine recombinant proteins, interleukins 1 alpha and 1 beta (rmIL-1 alpha and -1 beta), has been analyzed by Fourier transform infrared (IR) spectroscopy and then compared to results obtained by X-ray diffraction, circular dichroism (CD), and nuclear magnetic resonance (NMR) spectroscopy. The IR results obtained here for rmIL-1 alpha and -1 beta suggested that their secondary structures consisted predominantly of beta-sheets or strands. However, the analysis also revealed a significant absorption band near 1656 cm-1, which is typically assigned to alpha-helical or random structures. When these same murine polypeptides were analyzed by CD, no evidence of alpha-helical structures was observed. Further, published X-ray diffraction and NMR studies characterizing the human forms of IL-1 alpha and -1 beta indicate the absence of alpha-helices and that the human proteins are composed mainly of beta-strands (i.e., greater than 55%), with approximately 24% of the amino acids involved in large loops connecting the strands. The murine IL-1 proteins, when compared to their respective human counterparts, each show greater than 80% sequence homology. Given this fact, the CD analyses, and the result that this IR band amounted to 21% of the overall integrated area, the absorption peak at 1656 cm-1 was attributed to the presence of large loops rather than to alpha-helical or random structures. Such a structural assignment appears reasonable and is totally consistent with the established existence of large loops in the human forms as well as in other proteins found to fold similarly (viz., human bFGF).(ABSTRACT TRUNCATED AT 250 WORDS)  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