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 GPI (glycosylphosphatidylinositol) transamidase complex catalyses the attachment of GPI anchors to eukaryotic proteins in the lumen of ER (endoplasmic reticulum). The Saccharomyces cerevisiae GPI transamidase complex consists of the subunits yPIG-K (Gpi8p), yPIG-S (Gpi17p), yPIG-T (Gpi16p), yPIG-U (CDC91/GAB1) and yGPAA1. We present the production of the two recombinant proteins yGPAA170–247 and yGPAA170–339 of the luminal domain of S. cerevisiae GPAA1, covering the amino acids 70–247 and 70–339 respectively. The secondary structural content of the stable and monodisperse yGPAA170–247 has been determined to be 28% α-helix and 27% β-sheet. SAXS (small-angle X-ray scattering) data showed that yGPAA170–247 has an Rg (radius of gyration) of 2.72±0.025 nm and Dmax (maximum dimension) of 9.14 nm. These data enabled the determination of the two domain low-resolution solution structure of yGPAA170–247. The large elliptical shape of yGPAA170–247 is connected via a short stalk to the smaller hook-like domain of 0.8 nm in length and 3.5 nm in width. The topological arrangement of yGPAA170–247 will be discussed together with the recently determined low-resolution structures of yPIG-K24–337 and yPIG-S38–467 from S. cerevisiae in the GPI transamidase complex.  相似文献   

2.
Ohishi K  Inoue N  Kinoshita T 《The EMBO journal》2001,20(15):4088-4098
Many eukaryotic cell surface proteins are anchored to the plasma membrane via glycosylphosphatidylinositol (GPI). The GPI transamidase mediates GPI anchoring in the endoplasmic reticulum, by replacing a protein's C-terminal GPI attachment signal peptide with a pre-assembled GPI. During this transamidation reaction, the GPI transamidase forms a carbonyl intermediate with a substrate protein. It was known that the GPI transamidase is a complex containing GAA1 and GPI8. Here, we report two new components of this enzyme: PIG-S and PIG-T. To determine roles for PIG-S and PIG-T, we disrupted these genes in mouse F9 cells by homologous recombination. PIG-S and PIG-T knockout cells were defective in transfer of GPI to proteins, particularly in formation of the carbonyl intermediates. We also demonstrate that PIG-S and PIG-T form a protein complex with GAA1 and GPI8, and that PIG-T maintains the complex by stabilizing the expression of GAA1 and GPI8. Saccharomyces cerevisiae Gpi16p (YHR188C) and Gpi17p (YDR434W) are orthologues of PIG-T and PIG-S, respectively.  相似文献   

3.
The secondary structure implications of precipitation induced by a chaotropic salt, KSCN, and a structure stabilizing salt, Na2SO4, were studied for twelve different proteins. α-helix and β-sheet content of precipitate and native structures were estimated from the analysis of amide I band Raman spectra. A statistical analysis of the estimated perturbations in the secondary structure contents indicated that the most significant event is the formation of β-sheet structures with a concomitant loss of α-helix on precipitation with KSCN. The conformational changes for each protein were also analyzed with respect to elements of primary, secondary and tertiary structure existing in the native protein; primary structure was quantified by the fractions of hydrophobic and charged amino acids, secondary structure by x-ray estimates of α-helix and β-sheet contents of native proteins and tertiary structure by the dipole moment and solvent-accessible surface area. For the KSCN precipitates, factors affecting β-sheet content included the fraction of charged amino acids in the primary sequence and the surface area. Changes in α-helix content were influenced by the initial helical content and the dipole moment. The enhanced β-sheet contents of precipitates observed in this work parallel protein structural changes occurring in other aggregative phenomena.  相似文献   

4.
Hayward S  Milner-White EJ 《Proteins》2011,79(11):3193-3207
α-sheet has been proposed to be the main constituent of the toxic amyloid intermediate. Molecular dynamics simulations on proteins known to be involved in amyloid diseases have demonstrated that β-sheet can, under certain conditions, spontaneously convert to α-sheet via ββ→α(R)α(L) peptide-plane flipping. Using torsion-angle driving to simulate this flip the transition has been investigated for parallel and antiparallel sheets. Concerted and sequential flipping processes were simulated, the former allowing direct calculation of helical parameters. For antiparallel sheet, the strands tend to splay apart during the transition. This can be understood by consideration of the geometry of repeating dipeptide conformations. At the end of the transition antiparallel α-sheet is slightly twisted, comprising gently curving strands. In parallel sheet, the strands maintain identical conformations and stay hydrogen bonded during the transition as they curl up to suggest a hitherto unseen structure, the multi-helix α-nanotube. Intriguingly, the α-nanotube has some of the characteristics of the parallel β-helix, a single-helix structure also implicated in amyloid. Unlike the β-helix, α-nanotube formation could involve identical strands aligning with each other in register as in most amyloids.  相似文献   

5.
The Chou-Fasman conformational parameters, P, for amino acid residues in proteins are shown to be a linear function of intermolecular force and steric parameters. For α- helix, coil and turn parameters, steric effects are predominant; whereas for β-sheet parameters, intramolecular forces are predominant. Turn and coil parameters show little or no difference in their dependence which is different from that of α-helix and in some ways almost reciprocal. Factors which increase the probability of finding an amino acid residue in an α-helix usually decrease the probability of finding it in coil or turn. Values of P were calculated for several of the less common amino acids.  相似文献   

6.
A multiscale simulation method of protein folding is proposed, using atomic representation of protein and solvent, combing genetic algorithms to determine the key protein structures from a global view, with molecular dynamic simulations to reveal the local folding pathways, thus providing an integrated landscape of protein folding. The method is found to be superior to previously investigated global search algorithms or dynamic simulations alone. For secondary structure formation of a selected peptide, RN24, the structures and dynamics produced by this method agree well with corresponding experimental results. Three most populated conformations are observed, including hairpin, β-sheet and α-helix. The energetic barriers separating these three structures are comparable to the kinetic energy of the atoms of the peptide, implying that the transition between these states can be easily triggered by kinetic perturbations, mainly through electrostatic interactions between charged atoms. Transitions between α-helix and β-sheet should jump over at least two energy barriers and may stay in the energetic trap of hairpin. It is proposed that the structure of proteins should be jointly governed by thermodynamic and dynamic factors; free energy is not the exclusive dominant for stability of proteins.  相似文献   

7.
Despite sequence and structural similarity, TGF-β3 has low solubility among other isoforms of TGF-β. We used nanosecond of molecular dynamic simulations (MD) with explicit solvent, alone and in presence of urea, to investigate the intermediates resulting from the unfolding process of TGF-β3 and TGF-β1. MD simulations of the full-length proteins show a very early loss of α-helix in TGF-β3 compared to the one in the TGF-β1. MD simulation of a small fragment consisting of H3 α-helix of TGF-β3 shows conversion of this segment to β-sheet. Relative instability of H3 α-helix in TGF-β3 and its propensity to form β-sheet may explain the poor solubility of TGF-β3 compared to TGF-β1. The other reasons for poor solubility of TGF-β3 may be the hydrophobic patches on its surface and low charge over the entire range of pH.  相似文献   

8.
Prion diseases are neurodegenerative disorders associated with the accumulation of an abnormal isoform of the mammalian prion protein (PrP). Fourier transform infrared spectroscopy (FTIR) has previously been used to show that the conformation of aggregated, infectious PrP (PrP(Sc) ) varies between prion strains and these unique conformations may determine strain-specific disease phenotypes. However, the relative amounts of α-helix, β-sheet and other secondary structures have not always been consistent between studies, suggesting that other proteins might be confounding the analysis of PrP(Sc) secondary structure. We have used FTIR and LC-MS/MS to analyze enriched PrP(Sc) from mouse and hamster prion strains both before and after the removal of protein contaminants that commonly co-purify with PrP(Sc) . Our data show that non-PrP proteins do contribute to absorbances that have been associated with α-helical, loop, turn and β-sheet structures attributed to PrP(Sc) . The major contaminant, the α-helical protein ferritin, absorbs strongly at 1652 cm(-1) in the FTIR spectrum associated with PrP(Sc) . However, even the removal of more than 99% of the ferritin from PrP(Sc) did not completely abolish absorbance at 1652 cm(-1) . Our results show that contaminating proteins alter the FTIR spectrum attributed to PrP(Sc) and suggest that the α-helical, loop/turn and β-sheet secondary structure that remains following their removal are derived from PrP(Sc) itself.  相似文献   

9.
Many eukaryotic proteins are anchored to the cell surface via glycosylphosphatidylinositol (GPI), which is posttranslationally attached to the carboxyl-terminus by GPI transamidase. The mammalian GPI transamidase is a complex of at least four subunits, GPI8, GAA1, PIG-S, and PIG-T. Here, we report Chinese hamster ovary cells representing a new complementation group of GPI-anchored protein-deficient mutants, class U. The class U cells accumulated mature and immature GPI and did not have in vitro GPI transamidase activity. We cloned the gene responsible, termed PIG-U, that encoded a 435-amino-acid hydrophobic protein. The GPI transamidase complex affinity-purified from cells expressing epitope-tagged-GPI8 contained PIG-U and four other known components. Cells lacking PIG-U formed complexes of the four other components normally but had no ability to cleave the GPI attachment signal peptide. Saccharomyces cerevisiae Cdc91p, with 28% amino acid identity to PIG-U, partially restored GPI-anchored proteins on the surface of class U cells. PIG-U and Cdc91p have a functionally important short region with similarity to a region conserved in long-chain fatty acid elongases. Taken together, PIG-U and the yeast orthologue Cdc91p are the fifth component of GPI transamidase that may be involved in the recognition of either the GPI attachment signal or the lipid portion of GPI.  相似文献   

10.
Chakraborty H  Lentz BR 《Biochemistry》2012,51(5):1005-1008
Circular dichroism (CD) spectroscopy is an important technique in structural biology for examining folding and conformational changes of proteins in solution. However, the use of CD spectroscopy in a membrane medium (and also in a nonhomogeneous medium) is limited by (i) high light scattering and (ii) differential scattering of incident left and right circularly polarized light, especially at shorter wavelengths (<200 nm). We report a novel methodology for estimating the distortion of CD spectra caused by light scattering for membrane-bound peptides and proteins. The method is applied to three proteins with very different secondary structures to illustrate the limits of its capabilities when calibrated with a simple soluble peptide ([Ac]ANLKALEAQKQKEQRQAAEELANAK[OH], standard peptide) with a balanced secondary structure. The method with this calibration standard was quite successful in estimating α-helix but more limited when it comes to proteins with very high β-sheet or β-turn content.  相似文献   

11.
The N-terminal domain of human immunodeficiency virus (HIV)-1 glycoprotein 41,000 (FP; residues 1–23; NH2-AVGIGALFLGFLGAAGSTMGARS-CONH2) is involved in the fusion and cytolytic processes underlying viral-cell infection. Here, we use circular dichroism (CD) spectroscopy, along with electrospray ionization (ESI) mass spectrometry and tandem (MS/MS) mass spectrometry during the course of hydrogen/deuterium exchange, to probe the local conformations of this synthetic peptide in two membrane mimics. Since amino acids that participate in defined secondary structure (i.e., α-helix or β-sheet) exchange amido hydrogens more slowly than residues in random structures, deuterium exchange was combined with CD spectroscopy to map conformations to specific residues. For FP suspended in the highly structure-promoting solvent hexafluoroisopropanol (HFIP), CD spectra indicated high α-helix and disordered structures, whereas ESI and MS/MS mass spectrometry indicated that residues 5–15 were α-helical and 16–23 were disordered. For FP suspended in the less structure-promoting solvent trifluoroethanol (TFE), CD spectra showed lower α-helix, with ESI and MS/MS mass spectrometry indicating that only residues 9–15 participated in the α-helix. These results compare favorably with previous two-dimensional nuclear magnetic resonance studies on the same peptide. Proteins Suppl. 2:38–49, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

12.
In order to elucidate the protein folding problem, we performed molecular dynamics simulations for small- and middle-sized two unfolding and six refolding proteins in an explicit solvent. Histidine-containing phosphocarrier protein and small designed protein were chosen for the simulations. We found that the protein folding process of these proteins was divided into three phases: an α -helix formation phase, a packing phase and a β -sheet formation phase. In the α -helix formation phase, an α -helix was developed from a β -turn structure through a 310-helix state. In the packing phase, proteins became compact, and tertiary structures (α / α or pre- β / β packing) were formed. Formation of a hydrophobic nucleus occurred concomitant with the α -helix formation and packing phase. Finally, in the β -sheet formation phase, a β -sheet was developed owing to the sequential formation of hydrogen bonds between two neighbouring strands, just like a "closing zipper".  相似文献   

13.
The G subunit of the vacuolar ATPase (V-ATPase) is a component of the stalk connecting the V(1) and V(O) sectors of the enzyme and is essential for normal assembly and function. Subunit G (Vma10p) of the yeast V-ATPase was expressed in Escherichia coli as a soluble protein and was purified to homogeneity. The molecular mass of subunit G, determined by Native-polyacrylamide gel electrophoresis, gel filtration analysis and small-angle X-ray scattering, was approximately 28+/-2 kDa, indicating that this protein is dimeric. With a radius of gyration (R(g)) and a maximum size (D(max)) of 2.7+/-0.2 nm and 8.0+/-0.3 nm, respectively, the G-dimer is rather elongated. To understand which region of subunit G is required to mediate dimerization, a G(38-144) form (the carboxyl-terminus) was expressed and purified. G(38-144) is homogeneous, with a molecular mass of approximately 12+/-3 kDa, indicating a monomeric form in solution.  相似文献   

14.
Bacteriocins are proteins secreted by many bacterial cells to kill related bacteria of the same niche. To avoid their own suicide through reuptake of secreted bacteriocins, these bacteria protect themselves by co-expression of immunity proteins in the compartment of colicin destination. In Escherichia coli the colicin M (Cma) is inactivated by the interaction with the Cma immunity protein (Cmi). We have crystallized and solved the structure of Cmi at a resolution of 1.95? by the recently developed ab initio phasing program ARCIMBOLDO. The monomeric structure of the mature 10kDa protein comprises a long N-terminal α-helix and a four-stranded C-terminal β-sheet. Dimerization of this fold is mediated by an extended interface of hydrogen bond interactions between the α-helix and the four-stranded β-sheet of the symmetry related molecule. Two intermolecular disulfide bridges covalently connect this dimer to further lock this complex. The Cmi protein resembles an example of a 3D domain swapping being stalled through physical linkage. The dimer is a highly charged complex with a significant surplus of negative charges presumably responsible for interactions with Cma. Dimerization of Cmi was also demonstrated to occur in vivo. Although the Cmi-Cma complex is unique among bacteria, the general fold of Cmi is representative for a class of YebF-like proteins which are known to be secreted into the external medium by some Gram-negative bacteria.  相似文献   

15.
The 17-amino-acid N-terminal segment (htt(NT)) that leads into the polyglutamine (polyQ) segment in the Huntington's disease protein huntingtin (htt) dramatically increases aggregation rates and changes the aggregation mechanism, compared to a simple polyQ peptide of similar length. With polyQ segments near or above the pathological repeat length threshold of about 37, aggregation of htt N-terminal fragments is so rapid that it is difficult to tease out mechanistic details. We describe here the use of very short polyQ repeat lengths in htt N-terminal fragments to slow this disease-associated aggregation. Although all of these peptides, in addition to htt(NT) itself, form α-helix-rich oligomeric intermediates, only peptides with Q(N) of eight or longer mature into amyloid-like aggregates, doing so by a slow increase in β-structure. Concentration-dependent circular dichroism and analytical ultracentrifugation suggest that the htt(NT) sequence, with or without added glutamine residues, exists in solution as an equilibrium between disordered monomer and α-helical tetramer. Higher order, α-helix rich oligomers appear to be built up via these tetramers. However, only htt(NT)Q(N) peptides with N=8 or more undergo conversion into polyQ β-sheet aggregates. These final amyloid-like aggregates not only feature the expected high β-sheet content but also retain an element of solvent-exposed α-helix. The α-helix-rich oligomeric intermediates appear to be both on- and off-pathway, with some oligomers serving as the pool from within which nuclei emerge, while those that fail to undergo amyloid nucleation serve as a reservoir for release of monomers to support fibril elongation. Based on a regular pattern of multimers observed in analytical ultracentrifugation, and a concentration dependence of α-helix formation in CD spectroscopy, it is likely that these oligomers assemble via a four-helix assembly unit. PolyQ expansion in these peptides appears to enhance the rates of both oligomer formation and nucleation from within the oligomer population, by structural mechanisms that remain unclear.  相似文献   

16.
As part of a continuing investigation of netrins, an emerging class of extracellular matrix proteins that are involved in axon guidance activity, we have used dynamic light scattering (DLS) and small angle X-ray scattering to investigate the solution conformation of a truncated version of netrin-4 (Δnetrin-4) that lacks the C-terminal portion. The protein is characterized by a hydrodynamic (Stokes) radius (r(H)) of 4.60 (±0.20) nm, a radius of gyration (r(G)) of 4.42 (±0.20) nm and a maximum particle dimension (D(max)) of 16nm. More detailed ab initio modeling of the SAXS data indicates an extended rod like conformation for Δnetrin-4 in solution-a concept supported by the excellent agreement observed between experimental parameter estimates and those calculated for the ab initio models for Δnetrin-4 by the HYDROPRO program.  相似文献   

17.
Poly-β-benzyl-L -aspartate (poly[Asp(OBzl)]) forms either a lefthanded α-helix, β-sheet, ω-helix, or random coil under appropriate conditions. In this paper the Raman spectra of the above poly[Asp(OBzl)] conformations are compared. The Raman active amide I line shifts from 1663 cm?1 to 1679 cm?1 upon thermal conversion of poly[Asp(OBzl)] from the α-helical to β-sheet conformation while an intense line appearing at 890 cm?1 in the spectrum of the α-helix decreases in intensity. The 890 cm?1 line also displays weak intensity when the polymer is dissolved in chloroform–dichloroacetic acid solution and therefore is converted to the random coil. This line probably arises from a skeletal vibration and is expected to be conformationally sensitive. Similar behavior in the intensity of skeletal vibrations is discussed for other polypeptides undergoing conformational transitions. The Raman spectra of two cross-β-sheet copolypeptides, poly(Ala-Gly) and poly(Ser-Gly), are examined. These sequential polypeptides are model compounds for the crystalline regions of Bombyx mori silk fibroin which forms an extensive β-sheet structure. The amide I, III, and skeletal vibrations appeared in the Raman spectra of these polypeptides at the frequencies and intensities associated with β-sheet homopolypeptides. Since the sequential copolypeptides are intermediate in complexity between the homopolypeptides and the proteins, these results indicate that Raman structure–frequency correlations obtained from homopolypeptide studies can now be applied to protein spectra with greater confidence. The perturbation scheme developed by Krimm and Abe for explaining the frequency splitting of the amide I vibrations in β-sheet polyglycine is applied to poly(L -valine), poly-(Ala-Gly), poly(Ser-Gly), and poly[Asp(OBzl)]. The value of the “unperturbed” frequency, V0, for poly[Asp(OBzl)] was significantly greater than the corresponding values for the other polypeptides. A structural origin for this difference may be displacement of adjacent hydrogen-bonded chains relative to the standard β-sheet conformation.  相似文献   

18.
The venom of the North African scorpion Androctonus mauretanicus mauretanicus possesses numerous highly active neurotoxins that specifically bind to various ion channels. One of these, P05, has been found to bind specifically to calcium-activated potassium channels and also to compete with apamin, a toxin extracted from bee venom. Besides the highly potent ones, several of these peptides (including that of P01) have been purified and been found to possess only a very weak, although significant, activity in competition with apamin. The amino acid sequence of P01 shows that it is shorter than P05 by two residues. This deletion occurs within an α-helix stretch (residues 5–12). This α-helix has been shown to be involved in the interaction of P05 with its receptor via two arginine residues. These two arginines are absent in the P01 sequence. Furthermore, a proline residue in position 7 of the P01 sequence may act as an α-helix breaker. We have determined the solution structure of P01 by conventional two-dimensional 1H nuclear magnetic resonance and show that 1) the proline residue does not disturb the α-helix running from residues 5 to 12; 2) the two arginines are topologically replaced by two acidic residues, which explains the drop in activity; 3) the residual binding activity may be due to the histidine residue in position 9; and 4) the overall secondary structure is conserved, i.e., an α-helix running from residues 5 to 12, two antiparallel stretches of β-sheet (residues 15–20 and 23–27) connected by a type I′ β-turn, and three disulfide bridges connecting the α-helix to the β-sheet.  相似文献   

19.
We investigated the effect of gradual-elastic compression stockings (GCSs) on running economy (RE), kinematics, and performance in endurance runners. Sixteen endurance trained athletes (age: 34.73 ± 6.27 years; VO2max: 62.83 ± 9.03 ml·kg(-1)·min(-1); 38 minutes in 10 km; 1 hour 24 minutes in half marathon) performed in random order 4 bouts of 6 minutes at a recent half-marathon pace on a treadmill to evaluate RE with or without GCSs. Subsequently, 12 athletes were divided into 2 equal groups matched by their VO2max, and they performed a time limit test (T(lim)) on a treadmill at 105% of a recent 10-km pace with or without GCSs for evaluation of physiological responses and running kinematics. There were no significant differences in the RE test in all of the variables analyzed for the conditions, but a moderate reproducibility for some physiological responses was detected in the condition with GCSs. In the T(lim), the group that wore GCSs reached a lower % of maximum heart rate (HRmax) compared with the control group (96.00 ± 2.94 vs. 99.83 ± 0.40) (p = 0.01). Kinematics did not differ between conditions during the T(lim) (p > 0.05). There were improvement trends for time to fatigue (337 vs. 387 seconds; d = 0.32) and a lower VO2peak (≈53 vs. 62 ml·kg(-1)·min(-1); d = 1.19) that were detected with GCSs during the T(lim). These results indicate that GCSs reduce the % of HRmax reached during a test at competition pace. The lower reproducibility of the condition with GCSs perhaps suggests that athletes may possibly need an accommodation period for systematically experiencing the benefits of this garment, but this hypothesis should be further investigated.  相似文献   

20.
The secondary structure ofCerebratulus lacteus toxin B-IV, a neurotoxic polypeptide containing 55 amino acid residues and four disulfide bonds, was experimentally estimated by computer analyses of toxin circular dichroism (CD) and laser Raman spectra. The CD spectrum of the toxin displayed typical α-helical peaks at 191, 208, and 222 nm. At neutralpH, the α-helix estimates from CD varied between 49 and 55%, when nonrepresentative spectrum analytical methods were used. Analysis of the laser Raman spectrum obtained at a much higher toxin concentration yielded a 78% α-helix estimate. Both CD and Raman spectroscopic methods failed to detect any β-sheet structure. The spectroscopic analyses revealed significantly more α-helix and less β-sheet for toxin B-IV than was predicted from its sequence. To account for the difference between the 49–55% helix estimate from CD spectra and the 78% helix estimate from the Raman spectrum, we postulate that some terminal residues are unfolded at the low toxin concentrations used for CD measurements but form helix at the high toxin concentration used for Raman measurements. Our CD observations showing thatCerebatulus toxin B-IV helix content increases about 15% in trifluoroethanol or at highpH are consistent with this interpretation.  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