首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
A total of 1,911 proteins with N-terminal methionyl residues were computer screened for potential N-terminal alpha-helices with strong amphipathic character. By the criteria of D. Eisenberg (Annu. Rev. Biochem. 53:595-623, 1984), only 3.5% of nonplastid, nonviral proteins exhibited potential N-terminal alpha-helices, 18 residues in length, with hydrophobic moment values per amino acyl residue ([muH]) in excess of 0.4. By contrast, 10% of viral proteins exhibited corresponding [muH] values in excess of 0.4. Of these viral proteins with known functions, 55% were found to interact functionally with nucleic acids, 30% were membrane-interacting proteins or their precursors, and 15% were structural proteins, primarily concerned with host cell interactions. These observations suggest that N-terminal amphipathic alpha-helices of viral proteins may (i) function in nucleic acid binding, (ii) facilitate membrane insertion, and (iii) promote host cell interactions. Analyses of potential amphipathic N-terminal alpha-helices of cellular proteins are also reported, and their significance to organellar or envelope targeting is discussed.  相似文献   

2.
Statistical analysis of amino acid patterns in approximately 160,000 alpha-helices in experimentally determined structures revealed di-, tri-, and tetrapeptides, whose frequencies deviate most from the statistical model. Importantly, some sequences were never found in alpha- helices. This fact was detected initially with tripeptides, where nearly 1% of the possible sequences were never seen in the helical segments. For tetrapeptides, this effect is very strong and significant; almost 43% of the possible sequences never appear in alpha-helices. It is possible that there are some steric and energetic restrictions that do not allow these tetrameric amino acid sequences to form alpha-helical structure.  相似文献   

3.
UV- and CD-spectra of homogeneous enzymes have been measured. Extinction coefficients estimated from the UV-spectra are 0.97 for restriction endonuclease EcoRII at 279.5 nm and 1.17 for DNA-methylase EcoRII at 279 nm. As it follows from the CD spectra, both enzymes have a well developed tertiary structure and a highly ordered secondary structure, which consists of 22% alpha-helices, 64% beta-structure and 9% bends for REcoRII and of 44% alpha-helices, 48% beta-structure and 4% bends for MEcoRII. Restriction endonuclease denatures at 50 degrees C, while DNA-methylase denatures at 45 degrees C, with partial reversibility upon cooling.  相似文献   

4.
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   

5.
Helix geometry in proteins   总被引:39,自引:0,他引:39  
In this report we describe a general survey of all helices found in 57 of the known protein crystal structures, together with a detailed analysis of 48 alpha-helices found in 16 of the structures that are determined to high resolution. The survey of all helices reveals a total of 291 alpha-helices, 71 3(10)-helices and no examples of pi-helices. The conformations of the observed helices are significantly different from the "ideal" linear structures. The mean phi, psi angles for the alpha- and 3(10)-helices found in proteins are, respectively, (-62 degrees, -41 degrees) and (-71 degrees, -18 degrees). A computer program, HBEND, is used to characterize and to quantify the different types of helix distortion. alpha-Helices are classified as regular or irregular, linear, curved or kinked. Of the 48 alpha-helices analysed, only 15% are considered to be linear; 17% are kinked, and 58% are curved. The curvature of helices is caused by differences in the peptide hydrogen bonding on opposite faces of the helix, reflecting carbonyl-solvent/side-chain interactions for the exposed residues, and packing constraints for residues involved in the hydrophobic core. Kinked helices arise either as a result of included proline residues, or because of conflicting requirements for the optimal packing of the helix side-chains. In alpha-helices where there are kinks caused by proline residues, we show that the angle of kink is relatively constant (approximately 26 degrees), and that there is minimal disruption of the helix hydrogen bonding. The proline residues responsible for the kinks are highly conserved, suggesting that these distortions may be structurally/functionally important.  相似文献   

6.
To elucidate the structure of denatured proteins, we measured the vacuum-ultraviolet circular dichroism (VUVCD) spectra from 260 to 172 nm of three proteins (metmyoglobin, staphylococcal nuclease, and thioredoxin) in the native and the acid-, cold-, and heat-denatured states, using a synchrotron-radiation VUVCD spectrophotometer. The circular dichroism spectra of proteins fully unfolded by guanidine hydrochloride (GdnHCl) were also measured down to 197 nm for comparison. These denatured proteins exhibited characteristic VUVCD spectra that reflected a considerable amount of residual secondary structures. The contents of alpha-helices, beta-strands, turns, poly-L-proline type II (PPII), and unordered structures were estimated for each denatured state of the three proteins using the SELCON3 program with Protein Data Bank data and the VUVCD spectra of 31 reference proteins reported in our previous study. Based on these contents, the characteristics of the four types of denaturation were discussed for each protein. In all types of denaturation, a decrease in alpha-helices was accompanied by increases in beta-strands, PPII, and unordered structures. About 20% beta-strands were present even in the proteins fully unfolded by GdnHCl in which beta-sheets should be broken. From these results, we propose that denatured proteins constitute an ensemble of residual alpha-helices and beta-sheets, partly unfolded (or distorted) alpha-helices and beta-strands, PPII, and unordered structures.  相似文献   

7.
Shestopalov BV 《Biofizika》2005,50(6):998-1001
The methods of X-ray diffraction analysis and nuclear magnetic resonance spectroscopy were compared using the data on the identification of alpha-helices and beta-strands of the same proteins. The goal of the study was to determine whether these identifications can be considered as equivalent in the structural classification of proteins and in the solution of other problems. The identifications obtained by the method of Kabsch and Sander for 56% specially chosen water-soluble proteins were chosen. It was found that the identification of alpha-helices and beta-strands are almost equivalent if used for the structural classification of the proteins. In the analysis of the conformational properties of amino acid residues or their combinations, it is reasonable to use the identifications of alpha-helices and beta-strands, obtained only from the data of X-ray diffraction analysis. An additional outcome of the study is a unique collection of pairs of protein structures obtained by the methods of X-ray diffraction analysis and nuclear magnetic resonance spectroscopy for the same proteins.  相似文献   

8.
The pore-forming domain of colicin A (pfColA) fused to a prokaryotic signal peptide (sp-pfColA) is transported across and inserts into the inner membrane of Escherichia coli from the periplasmic side and forms a functional channel. The soluble structure of pfColA consists of a ten-helix bundle containing a hydrophobic helical hairpin. Here, we generated a series of mutants in which an increasing number of sp-pfColA alpha-helices was deleted. These peptides were tested for their ability to form ion channels in vivo and in vitro. We found that the shortest sp-pfColA mutant protein that killed Escherichia coli was composed of the five last alpha-helices of sp-pfColA, whereas the shortest peptide that formed a channel in planar lipid bilayer membranes similar to that of intact pfColA was the protein composed of the last six alpha-helices. The peptide composed of the last five alpha-helices of pfColA generated a voltage-independent conductance in planar lipid bilayer with properties very different from that of intact pfColA. Thus, helices 1 to 4 are unnecessary for channel formation, while helix 5, or some part of it, is important but not absolutely necessary. Voltage-dependence of colicin is evidently controlled by the first four alpha-helices of pfColA.  相似文献   

9.
Cooperativity between pepsin and crystallization of calcium carbonate in distilled water was studied. The results show that vaterite was formed under the influence of pepsin and the crystalline product was a composite of vaterite and pepsin. The component of this material was similar to that of nacre. At the same time, the crystallization of calcium carbonate had also an important effect on the secondary structure of the pepsin. The secondary structure of the pepsin was characterized through FT-IR technology. The result indicated that the pure pepsin is composed of 24.38% alpha-helices, 29.91% beta-sheets, 39.32% beta-turns and 6.49% random structures and the pepsin in the CaCO(3)-pepsin solution is composed of 2.09% alpha-helices, 93.304% beta-sheets, 4.592% beta-turns and 0.006% random structure. During the crystallization of the calcium carbonate from the pepsin solution, the secondary structure of the pepsin transformed. These results showed that there was cooperativity between the crystallization of vaterite and the pepsin. The cooperative mechanism is discussed.  相似文献   

10.
Occurrence of bifurcated three-center hydrogen bonds in proteins   总被引:2,自引:0,他引:2  
R Preissner  U Egner  W Saenger 《FEBS letters》1991,288(1-2):192-196
Analysis of 13 high-resolution protein X-ray crystal structures shows that 1204 (24%) of all the 4974 hydrogen bonds are of the bifurcated three-center type with the donor X-H opposing two acceptors A1, A2. They occur systematically in alpha-helices where 90% of the hydrogen bonds are of this type; the major component is (n + 4)N-H ... O = C(n) as expected for a 3.6(13) alpha-helix, and the minor component is (n + 4)N-H ... O = C(n + 1), as observed in 3(10) helices; distortions at the C-termini of alpha-helices are stabilized by three-center bonds. In beta-sheets 40% of the hydrogen bonds are three-centered. The frequent occurrence of three-center hydrogen bonds suggests that they should not be neglected in protein structural studies.  相似文献   

11.
Local determinants of 3(10)-helix stabilization have been ascertained from the analysis of the crystal structure data base. We have clustered all 5-length substructures from 51 nonhomologous proteins into classes based on the conformational similarity of their backbone dihedral angles. Several clusters, derived from 3(10)-helices and multiple-turn conformations, had strong amino acid sequence patterns not evident among alpha-helices. Aspartate occurred over twice as frequently in the N-cap position of 3(10)-helices as in the N-cap position of alpha-helices. Unlike alpha-helices, 3(10)-helices had few C-termini ending in a left-handed alpha conformation; most 3(10) C-caps adopted an extended conformation. Differences in the distribution of hydrophobic residues among 3(10)- and alpha-helices were also apparent, producing amphipathic 3(10)-helices. Local interactions that stabilize 3(10)-helices can be inferred both from the strong amino acid preferences found for these short helices, as well as from the existence of substructures in which tertiary interactions replace consensus local interactions. Because the folding and unfolding of alpha-helices have been postulated to proceed through reverse-turn and 3(10)-helix intermediates, sequence differences between 3(10)- and alpha-helices can also lend insight into factors influencing alpha-helix initiation and propagation.  相似文献   

12.
The GXXXG motif is a frequently occurring sequence of residues that is known to favor helix-helix interactions in membrane proteins. Here we show that the GXXXG motif is also prevalent in soluble proteins whose structures have been determined. Some 152 proteins from a non-redundant PDB set contain at least one alpha-helix with the GXXXG motif, 41 +/- 9% more than expected if glycine residues were uniformly distributed in those alpha-helices. More than 50% of the GXXXG-containing alpha-helices participate in helix-helix interactions. In fact, 26 of those helix-helix interactions are structurally similar to the helix-helix interaction of the glycophorin A dimer, where two transmembrane helices associate to form a dimer stabilized by the GXXXG motif. As for the glycophorin A structure, we find backbone-to-backbone atomic contacts of the C alpha-H...O type in each of these 26 helix-helix interactions that display the stereochemical hallmarks of hydrogen bond formation. These glycophorin A-like helix-helix interactions are enriched in the general set of helix-helix interactions containing the GXXXG motif, suggesting that the inferred C alpha-H...O hydrogen bonds stabilize the helix-helix interactions. In addition to the GXXXG motif, some 808 proteins from the non-redundant PDB set contain at least one alpha-helix with the AXXXA motif (30 +/- 3% greater than expected). Both the GXXXG and AXXXA motifs occur frequently in predicted alpha-helices from 24 fully sequenced genomes. Occurrence of the AXXXA motif is enhanced to a greater extent in thermophiles than in mesophiles, suggesting that helical interaction based on the AXXXA motif may be a common mechanism of thermostability in protein structures. We conclude that the GXXXG sequence motif stabilizes helix-helix interactions in proteins, and that the AXXXA sequence motif also stabilizes the folded state of proteins.  相似文献   

13.
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   

14.
Recent Fourier transform infrared spectroscopy (FTIR) with attenuated total reflection technique (ATR) has been applied to alpha-keratin fibers (horse-hair) extended in water both at 21 and 95 degrees C. Infrared absorption bands in the Amide 1 region indicated that at extensions to 40-50% strain in water at 21 degrees C alpha-helices had completely disappeared and parallel beta-sheets were formed [Appl. Spectrosc. 55 (2001) 552]. However, when the hair fibers were extended to the same strain at 95 degrees C in water the result was the formation of anti-parallel beta-sheets. These results suggest that the relatively more stable anti-parallel beta-state [Polymer 10 (1969) 810] is only attained in extended alpha-keratin fibers at elevated temperatures and must result from major molecular rearrangement. It was concluded that the alpha-helices in the intermediate filaments (IFs) of alpha-keratin fibers must be parallel. This is in contrast to the previously accepted orientation of anti-parallel alpha-helices, based primarily on findings of X-ray diffraction studies of the structure of beta-keratin in highly extended fibers [Polymer 10 (1969) 810; Keratins, IL: Thomas Springfield (1972); Nature 316 (1985) 767].  相似文献   

15.
16.
The mitochondrial oxoglutarate carrier belongs to the mitochondrial carrier family and exchanges oxoglutarate for malate and other dicarboxylates across the mitochondrial inner membrane. Here, single-cysteine mutant carriers were engineered for every residue in the amino- and carboxy-terminus, cytoplasmic loops, and matrix alpha-helices and their transport activity was measured in the presence and absence of sulfhydryl reagents. The analysis of the cytoplasmic side of the oxoglutarate carrier showed that the conserved and symmetric residues of the mitochondrial carrier motif [DE]XX[RK] localized at the C-terminal end of the even-numbered transmembrane alpha-helices are important for the function of the carrier, but the non-conserved cytoplasmic loops and termini are not. On the mitochondrial matrix side of the carrier most residues of the three matrix alpha-helices that are in the interface with the transmembrane alpha-helical bundle are important for function. Among these are the residues of the symmetric [ED]G motif present at the C-terminus of the matrix alpha-helices; the tyrosines of the symmetric YK motif at the N-terminus of the matrix alpha-helices; and the hydrophobic residues M147, I171 and I247. The functional role of these residues was assessed in the structural context of the homology model of OGC. Furthermore, in this study no evidence was found for the presence of a specific homo-dimerisation interface on the surface of the carrier consisting of conserved, asymmetric and transport-critical residues.  相似文献   

17.
Protein segments that form amphipathic alpha-helices in their native state have periodic variation in the hydrophobicity values of the residues along the segment, with a 3.6 residue per cycle period characteristic of the alpha-helix. The assignment of hydrophobicity values to amino acids (hydrophobicity scale) affects the display of periodicity. Thirty-eight published hydrophobicity scales are compared for their ability to identify the characteristic period of alpha-helices, and an optimum scale for this purpose is computed using a new eigenvector method. Two of the published scales are also characterized by eigenvectors. We compare the usual method for detecting periodicity based on the discrete Fourier transform with a method based on a least-squares fit of a harmonic sequence to a sequence of hydrophobicity values. The two become equivalent for very long sequences, but, for shorter sequences with lengths commonly found in alpha-helices, the least-squares procedure gives a more reliable estimate of the period. The analog to the usual Fourier transform power spectrum is the "least-squares power spectrum", the sum of squares accounted for in fitting a sinusoid of given frequency to a sequence of hydrophobicity values. The sum of the spectra of the alpha-helices in our data base peaks at 97.5 degrees, and approximately 50% of the helices can account for this peak. Thus, approximately 50% of the alpha-helices appear to be amphipathic, and, of those that are, the dominant frequency at 97.5 degrees rather than 100 degrees indicates that the helix is slightly more open than previously thought, with the number of residues per turn closer to 3.7 than 3.6. The extra openness is examined in crystallographic data, and is shown to be associated with the C terminus of the helix. The alpha amphipathic index, the key quantity in our analysis, measures the fraction of the total spectral area that is under the 97.5 degrees peak, and is a characteristic of hydrophobicity scales that is consistent for different sets of helices. Our optimized scale maximizes the amphipathic index and has a correlation of 0.85 or higher with nine previously published scales. The most surprising feature of the optimized scale is that arginine tends to behave as if it were hydrophobic; i.e. in the crystallographic data base it has a tendency to be on the hydrophobic face of teh amphipathic helix. Although the scale is optimal only for predicting alpha-amphipathicity, it also ranks high in identifying beta-amphipathicity and in distinguishing interior from exterior residues in a protein.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

18.
Investigations are reported on the possibility of the direct transformations from alpha-helices to beta-sheets in poly(D-alanine) by heating and the immersion into formic acid, and were concluded that intrahelical hydrogen bonds in the alpha-helix were stable, while interhelical interactions in the alpha-helix crystal were weakened, and beta-sheets did not transform directly from alpha-helices, but grew from random coils.  相似文献   

19.
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   

20.
The crystal structure of a catalytically active fragment of glucoamylase-I from Aspergillus awamori var. X100 has been determined to a resolution of 2.2 A. Twelve of its 13 alpha-helices are arranged into an "alpha/alpha-barrel." An inner core of six mutually parallel alpha-helices are connected to each other through a peripheral set of six alpha-helices. The peripheral helices are parallel to each other, but approximately antiparallel to the inner core of alpha-helices. The putative active site lies in the packing void of the inner set of helices. The last 30 residues of the enzyme comprise a separate domain containing 10 sites of O-glycosylation. Each instance of O-glycosylation involves a serine or threonine side chain linked to the alpha-anomer of a single mannosyl residue. The O-glycosylated domain is in an extended conformation, wrapping around the "waist" of the alpha/alpha-barrel. Two additional sites of N-glycosylation contribute well ordered glycosyl chains that lie in proximity to the belt of O-glycosylation. The model developed for glucoamylase is a rare and valuable structural example of a glycoprotein and an exo-acting amylolytic enzyme.  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