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 most conspicuous structural characteristic of the alpha-helical membrane proteins is their long transmembrane alpha-helices. However, other structural elements, as yet largely ignored in statistical studies of membrane protein structure, are found in those parts of the protein that are located in the membrane-water interface region. Here, we show that this region is enriched in irregular structure and in interfacial helices running roughly parallel with the membrane surface, while beta-strands are extremely rare. The average amino acid composition is different between the interfacial helices, the parts of the transmembrane helices located in the interface region, and the irregular structures. In this region, hydrophobic and aromatic residues tend to point toward the membrane and charged/polar residues tend to point away from the membrane. The interface region thus imposes different constraints on protein structure than do the central hydrocarbon core of the membrane and the surrounding aqueous phase.  相似文献   

2.
We address the question of the thermal stability of proteins in thermophiles through comprehensive genome comparison, focussing on the occurrence of salt bridges. We compared a set of 12 genomes (from four thermophilic archaeons, one eukaryote, six mesophilic eubacteria, and one thermophilic eubacteria). Our results showed that thermophiles have a greater content of charged residues than mesophiles, both at the overall genomic level and in alpha helices. Furthermore, we found that in thermophiles the charged residues in helices tend to be preferentially arranged with a 1–4 helical spacing and oriented so that intra-helical charge pairs agree with the helix dipole. Collectively, these results imply that intra-helical salt bridges are more prevalent in thermophiles than mesophiles and thus suggest that they are an important factor stabilizing thermophilic proteins. We also found that the proteins in thermophiles appear to be somewhat shorter than those in mesophiles. However, this later observation may have more to do with evolutionary relationships than with physically stabilizing factors. In all our statistics we were careful to controls for various biases. These could have, for instance, arisen due to repetitive or duplicated sequences. In particular, we repeated our calculation using a variety of random and directed sampling schemes. One of these involved making a "stratified sample," a representative cross-section of the genomes derived from a set of 52 orthologous proteins present roughly once in each genome. For another sample, we focused on the subset of the 52 orthologs that had a known 3D structure. This allowed us to determine the frequency of tertiary as well as main-chain salt bridges. Our statistical controls supported our overall conclusion about the prevalence of salt bridges in thermophiles in comparison to mesophiles. Electronic Publication  相似文献   

3.
It was established previously that helical propensities of different amino acid residues in the middle of α‐helix in peptides and in proteins are very similar. The statistical analysis of the protein helices from the known three‐dimensional structures shows no difference in the frequency of noncharged residues in the middle and at the C terminus. Yet, experimental studies show distinctive differences for the helical propensities of noncharged residues in the middle and in the C terminus in model peptides. Is this a general effect, and is it applicable to protein helices or is it specific to the model alanine‐based peptides? To answer this question, the effects of substitutions at positions 28 (middle residue) and 32 (C2 position at the C terminus) of the α‐helix of ubiquitin on the stability of this protein are measured by using differential scanning calorimetry. The two data sets produce similar values for intrinsic helix propensity, leading to a conclusion that noncharged amino acid residues at the solvent‐exposed positions in the middle and at the C terminus of the α‐helix have the same helical propensity. This conclusion is further supported with an excellent correlation between the helix propensity scale obtained for the two positions in ubiquitin with the experimental helix propensity scale established previously and with the statistical distribution of the residues in protein helices.  相似文献   

4.
The general distribution of 2259 amino acid residues in α-helical and in non-helical regions has been calculated using data from 12 proteins. The irregular distribution of hydrophobic residues in the helical parts of the sample permits classification of the helical regions in hydrophobic-clustered and hydrophobic-depleted zones. In comparison with the composition of α-helices, Ala, Leu, Val, Ile and Tyr predominate in hydrophobic-clustered zones, whereas Cys and Gln tend to accumulate in hydrophobic-depleted zones. The N-terminal part of the helices is rich in Asp and Pro and poor in Leu and His, whereas the C-terminal is rich in Lys and Gin and does not contain Pro.For helical regions, hydrophobic residues, located along the sequence with a relation of 1-2-5 or 1-4-5, are more frequent than those with a relation of 1-2-3, 1-2-4, 1-3-4 or 1-3-5. The opposite is found for non-helical regions. All these deviations are statistically significant. It is concluded that hydrophobic triplets 1-2-5 and 1-4-5 are required for stabilizing the helices.The relevance of our results, and others reported in the bibliography, towards a better understanding of the structural code for proteins is discussed.  相似文献   

5.
Balaji S  Aruna S  Srinivasan N 《Proteins》2003,53(4):783-791
Occurrence and accommodation of charged amino acid residues in proteins that are structurally equivalent to buried non-polar residues in homologues have been investigated. Using a dataset of 1,852 homologous pairs of crystal structures of proteins available at 2A or better resolution, 14,024 examples of apolar residues in the structurally conserved regions replaced by charged residues in homologues have been identified. Out of 2,530 cases of buried apolar residues, 1,677 of the equivalent charged residues in homologues are exposed and the rest of the charged residues are buried. These drastic substitutions are most often observed in homologous protein pairs with low sequence identity (<30%) and in large protein domains (>300 residues). Such buried charged residues in the large proteins are often located in the interface of sub-domains or in the interface of structural repeats, Beyond 7A of residue depth of buried apolar residues, or less than 4% of solvent accessibility, almost all the substituting charged residues are buried. It is also observed that acidic sidechains have higher preference to get buried than the positively charged residues. There is a preference for buried charged residues to get accommodated in the interior by forming hydrogen bonds with another sidechain than the main chain. The sidechains interacting with a buried charged residue are most often located in the structurally conserved regions of the alignment. About 50% of the observations involving hydrogen bond between buried charged sidechain and another sidechain correspond to salt bridges. Among the buried charged residues interacting with the main chain, positively charged sidechains form hydrogen bonds commonly with main chain carbonyls while the negatively charged residues are accommodated by hydrogen bonding with the main chain amides. These carbonyls and amides are usually located in the loops that are structurally variable among homologous proteins.  相似文献   

6.
Equine β-lactoglobulin (ELG) assumes non-native helices during refolding and in partially folded states. Previously, circular dichroism (CD) combined with site-directed mutagenesis identified helical regions in the acid- and cold-denatured states of ELG. It is also known that a fragment of ELG, CHIBL (residues 88-142), has a structure similar to that of the cold-denatured state. For the study reported herein, the structure of a shorter fragment, CHIBLΔF (residues 97-142), was investigated by CD and nuclear magnetic resonance spectroscopy. The secondary chemical shifts clearly showed that non-native α-helices are present in two different regions, residues 98-107 and 114-135, and are connected by a native disulfide bond. The CD spectra of two peptides that correspond to the helical regions are characterized by weak helical signatures, and the sum of their CD spectra is nearly the same as the spectrum of disulfide-reduced CHIBLΔF. Therefore, the non-native helices are stabilized by the disulfide, and non-native helix formation may occur only during the refolding of the disulfide-intact protein. Supporting this conclusion is the observation that tear lipocalin, a homologue of ELG that lacks the disulfide, does not form non-native helices during folding.  相似文献   

7.
Ruta V  Chen J  MacKinnon R 《Cell》2005,123(3):463-475
Voltage-dependent ion channels open and conduct ions in response to changes in cell-membrane voltage. The voltage sensitivity of these channels arises from the motion of charged arginine residues located on the S4 helices of the channel's voltage sensors. In KvAP, a prokaryotic voltage-dependent K+ channel, the S4 helix forms part of a helical hairpin structure, the voltage-sensor paddle. We have measured the membrane depth of residues throughout the KvAP channel using avidin accessibility to different-length tethered biotin reagents. From these measurements, we have calibrated the tether lengths and derived the thickness of the membrane that forms a barrier to avidin penetration, allowing us to determine the magnitude of displacement of the voltage-sensor paddles during channel gating. Here we show that the voltage-sensor paddles are highly mobile compared to other regions of the channel and transfer the gating-charge arginines 15-20 A through the membrane to open the pore.  相似文献   

8.
Adamian L  Liang J 《Proteins》2002,47(2):209-218
Polar and ionizable amino acid residues are frequently found in the transmembrane (TM) regions of membrane proteins. In this study, we show that they help to form extensive hydrogen bond connections between TM helices. We find that almost all TM helices have interhelical hydrogen bonding. In addition, we find that a pair of contacting TM helices is packed tighter when there are interhelical hydrogen bonds between them. We further describe several spatial motifs in the TM regions, including "Polar Clamp" and "Serine Zipper," where conserved Ser residues coincide with tightly packed locations in the TM region. With the examples of halorhodopsin, calcium-transporting ATPase, and bovine cytochrome c oxidase, we discuss the roles of hydrogen bonds in stabilizing helical bundles in polytopic membrane proteins and in protein functions.  相似文献   

9.
Alpha-helical bundles and beta-barrel proteins represent the two basic types of architecture known for integral membrane proteins. Irregular structural motifs have been revealed with the growing number of structures determined. "Discontinuous" helices are present in membrane proteins that actively transport ions. In the Ca(2+)-ATPase, a primary active transporter, and in the secondary transporters NhaA, LeuT(Aa), ClC H(+)/Cl(-) exchanger and Glt(Ph), the helical structure of two membrane segments is interrupted and the interjacent polypeptide chain forms an extended peptide. The discontinuous helices are integrated in the membrane either as transmembrane-spanning or hairpin-type segments. In addition, the secondary transporters have inverted internal duplication domains, which are only weakly correlated with their amino acid sequence. The symmetry comprises either parts of or the complete molecule, but always includes the discontinuous helices. The helix-peptide-helix motif is correlated with the ion translocation function. The extended peptides with their backbone atoms, the helix termini and the polar/charged amino acid residues in close vicinity provide the basis for ion recognition, binding and translocation.  相似文献   

10.
Caputo GA  London E 《Biochemistry》2004,43(27):8794-8806
The behavior of model-membrane-inserted polyLeu-rich peptides containing Asp residues located at various positions in their hydrophobic core was investigated. The topography of the bilayer-inserted alpha helices formed by these peptides was evaluated by measuring the emission lambda(max) and quenching the fluorescence of a Trp at the center of the peptide sequence. When Asp residues were protonated (at low pH), peptides that were incorporated into vesicles composed of dioleoylphosphatidylcholine (DOPC) adopted a topography in which the polyLeu sequence predominantly formed a normal transmembrane (TM) helix. When Asp residues were ionized (at neutral or high pH), topography was altered in a manner that would allow the charged Asp residues to reside near the bilayer surface. In DOPC vesicles, most peptides repositioned so that the longest segment of consecutive hydrophobic residues (12 residue minimum) formed a truncated/shifted TM structure. However, peptides with one or two charged Asp residues close to the center of the hydrophobic sequence and thus lacking even a 12-residue continuous hydrophobic segment, formed a helical non-TM state locating near the bilayer surface. At low pH, incorporation of the peptides into thicker bilayers composed of dierucoylphosphatidylcholine (DEuPC) resulted in the formation of a mixture of the normal TM state and the non-TM helical state located near the bilayer surface. In DEuPC vesicles at high pH, the non-TM state tended to predominate. How Asp-ionization-dependent shifts in helix topography may regulate the function of membrane proteins exposed to environments with differing pH in vivo (e.g., endosomes) is discussed.  相似文献   

11.
We present a method for the computer-based iterative assembly of native-like tertiary structures of helical proteins from alpha-helical fragments. For any pair of helices, our method, called MATCHSTIX, first generates an ensemble of possible relative orientations of the helices with various ways to form hydrophobic contacts between them. Those conformations having steric clashes, or a large radius of gyration of hydrophobic residues, or with helices too far separated to be connected by the intervening linking region, are discarded. Then, we attempt to connect the two helical fragments by using a robotics-based loop-closure algorithm. When loop closure is feasible, the algorithm generates an ensemble of viable interconnecting loops. After energy minimization and clustering, we use a representative set of conformations for further assembly with the remaining helices, adding one helix at a time. To efficiently sample the conformational space, the order of assembly generally proceeds from the pair of helices connected by the shortest loop, followed by joining one of its adjacent helices, always proceeding with the shorter connecting loop. We tested MATCHSTIX on 28 helical proteins each containing up to 5 helices and found it to heavily sample native-like conformations. The average rmsd of the best conformations for the 17 helix-bundle proteins that have 2 or 3 helices is less than 2 A; errors increase somewhat for proteins containing more helices. Native-like states are even more densely sampled when disulfide bonds are known and imposed as restraints. We conclude that, at least for helical proteins, if the secondary structures are known, this rapid rigid-body maximization of hydrophobic interactions can lead to small ensembles of highly native-like structures. It may be useful for protein structure prediction.  相似文献   

12.
13.
NMRsolution structures are reported for two mutants (K16E, K16F) of the soluble amyloid beta peptide Abeta(1-28). The structural effects of these mutations of a positively charged residue to anionic and hydrophobic residues at the alpha-secretase cleavage site (Lys16-Leu17) were examined in the membrane-simulating solvent aqueous SDS micelles. Overall the three-dimensional structures were similar to that for the native Abeta(1-28) sequence in that they contained an unstructured N-terminus and a helical C-terminus. These structural elements are similar to those seen in the corresponding regions of full-length Abeta peptides Abeta(1-40) and Abeta(1-42), showing that the shorter peptides are valid model systems. The K16E mutation, which might be expected to stabilize the macrodipole of the helix, slightly increased the helix length (residues 13-24) relative to the K16F mutation, which shortened the helix to between residues 16 and 24. The observed sequence-dependent control over conformation in this region provides an insight into possible conformational switching roles of mutations in the amyloid precursor protein from which Abeta peptides are derived. In addition, if conformational transitions from helix to random coil to sheet precede aggregation of Abeta peptides in vivo, as they do in vitro, the conformation-inducing effects of mutations at Lys16 may also influence aggregation and fibril formation.  相似文献   

14.
Positions of metal-binding residues with respect to helical terminii in protein structures have been analyzed in order to determine if the location of these ligands is influenced by the helix dipole. Most ligands do not show any preference for the amino- or carboxy-terminus of a helix. For steric reasons, peptide ligands can be located only at the C-terminus. The availability of a second ligand residue closely placed along the sequence may be of more importance, rather than the electrostatic interaction involving helix dipole, in cases where ligands are found near the C-terminus. The location of heme-binding histidine residues at the C-terminus may be due to the steric requirements of the heme group and also the intrahelical hydrogen bond that the residues can form at this position. Considerations based on such geometrical features and not just the helix dipole may help us to understand the observed distribution of charged residues along alpha-helices and the favorable role these amino acids have on folding of isolated helices.  相似文献   

15.
The amphipathic helical theory of Segrest and colleagues (FEBS Lett.:38:247-253, 1974) proposes that the lipid-binding segments of serum apolipoproteins are in an alpha helical conformation. Furthermore the helices have a hydrophobic face and a hydrophilic face with a specific distribution of positively and negatively charged residues. The importance of the pattern of the charged residues in the lipid binding and lecithin:cholesterol acyltransferase (LCAT) activation by the segments is still debated. We designed a 30-residue peptide, GALA, which in the alpha helical conformation has a hydrophilic face composed of glutamic acid residues (Sabbarao et al.: Biochemistry 26:2964-2972, 1987). GALA behaves like the serum apolipoproteins in its interaction with dimyristoylphosphatidylcholine (DMPC) at neutral pH; the amino terminal tryptophan of GALA undergoes a blue shift in its fluorescence emission spectrum, and the circular dichroism (CD) spectrum indicates that GALA acquires alpha helical structure in the presence of DMPC. A DMPC-GALA:19/1 (molar ratio) complex can be isolated by gel-permeation chromatography. This complex has a discoidal structure with the approximate dimensions of 44-A edge thickness and a 170- to 350-A diameter. GALA activates LCAT with DMPC but not with unsaturated phospholipids as the substrate. The apparent partition coefficient of GALA into DMPC vesicles is 100-fold larger than into egg phosphatidylcholine vesicles. The interaction of GALA with unsaturated lipids at neutral pH is so weak that no detectable change in the spectroscopic properties of GALA or the structure of the liposomes can be detected under the conditions used here. The sequence of GALA differs from previously studied model Apo A1 peptides by the absence of positively charged residues on the hydrophilic face. This indicates that positive charges in Apo A1-like peptides are not required in order to form discoidal structures with saturated phospholipids or to activate LCAT with such lipid substrates.  相似文献   

16.
Class A G protein-coupled receptors (GPCRs) are known to form dimers and/or oligomeric arrays in vitro and in vivo. These complexes are thought to play important roles in modulating class A GPCR function. Many studies suggest that residues located on the "outer" (lipid-facing) surface of the transmembrane (TM) receptor core are critically involved in the formation of class A receptor dimers (oligomers). However, no clear consensus has emerged regarding the identity of the TM helices or TM subsegments involved in this process. To shed light on this issue, we have used the M(3) muscarinic acetylcholine receptor (M3R), a prototypic class A GPCR, as a model system. Using a comprehensive and unbiased approach, we subjected all outward-facing residues (70 amino acids total) of the TM helical bundle (TM1-7) of the M3R to systematic alanine substitution mutagenesis. We then characterized the resulting mutant receptors in radioligand binding and functional studies and determined their ability to form dimers (oligomers) in bioluminescence resonance energy transfer saturation assays. We found that M3R/M3R interactions are not dependent on the presence of one specific structural motif but involve the outer surfaces of multiple TM subsegments (TM1-5 and -7) located within the central and endofacial portions of the TM receptor core. Moreover, we demonstrated that the outward-facing surfaces of most TM helices play critical roles in proper receptor folding and/or function. Guided by the bioluminescence resonance energy transfer data, molecular modeling studies suggested the existence of multiple dimeric/oligomeric M3R arrangements, which may exist in a dynamic equilibrium. Given the high structural homology found among all class A GPCRs, our results should be of considerable general relevance.  相似文献   

17.
Recent studies with model peptides and statistical analyses of the crystal structures of membrane proteins have shown that buried polar interactions contribute significantly to the stabilization of the three-dimensional structures of membrane proteins. Here, we probe how the location of these polar groups along the transmembrane helices affect their free energies of interaction. Asn residues were placed singly and in pairs at three positions within a model transmembrane helix, which had previously been shown to support the formation of trimers in micelles. The model helix was designed to form a transmembrane coiled coil, with Val side chains at the "a" positions of the heptad repeat. Variants of this peptide were prepared in which an Asn residue was introduced at one or more of the "a" positions, and their free energies of association were determined by analytical ultracentrifugation. When placed near the middle of the transmembrane helix, the formation of trimers was stabilized by at least -2.0 kcal/mol per Asn side chain. When the Asn was placed at the interface between the hydrophobic and polar regions of the peptide, the substitution was neither stabilizing nor destabilizing (0.0 +/- 0.5 kcal/mol of monomer). Finally, it has previously been shown that a Val-for-Asn mutation in a water-soluble coiled coil destabilizes the structure by approximately 1.5 kcal/mol of monomer [Acharya, A., et al. (2002) Biochemistry 41, 14122-14131]. Thus, the headgroup region of a micelle appears to have a conformational impact intermediate between that of bulk water and the apolar region of micelle. A similarly large dependence on the location of the polar residues was found in a statistical survey of helical transmembrane proteins. The tendency of different types of residues to be buried in the interiors versus being exposed to lipids was analyzed. Asn and Gln show a very strong tendency to be buried when they are located near the middle of a transmembrane helix. However, when placed near the ends of transmembrane helices, they show little preference for the surface versus the interior of the protein. These data show that Asn side chains within the apolar region of the transmembrane helix provide a significantly larger driving force for association than Asn residues near the apolar/polar interface. Thus, although polar interactions are able to strongly stabilize the folding of membrane proteins, the energetics of association depend on their location within the hydrophobic region of a transmembrane helix.  相似文献   

18.
Wang G 《FEBS letters》2002,529(2-3):157-161
The X-ray structure of the N-terminal truncated human apoA-I [Borhani et al., Proc. Natl. Acad. Sci. USA 94 (1997) 12291] and the NMR structure of intact human apoA-I [Okon et al., FEBS Lett. 517 (2002) 139] found similar repeating helices. The crystal structure is a twisted circular four-helix bundle, consisting of four molecules of apoA-I(44-243), where four copies of the lecithin:cholesterol acyltransferase (LCAT)-activating domains are located outside the ring structure, while the aromatic-rich strong lipid-binding domains are inside. This architecture suggests a lipid-binding mechanism that lipids directly enter the hole of the crystal structure. Indeed, four copies of Trp50 and Trp72 are exposed and oriented toward the center of the ring, initiating lipid binding. This is followed by the inside-out rotations of the terminal helices to make a belt with all the hydrophobic faces of the helices facing inward. Such lipid-binding induced rotations have an impact on the conformation of the lipid-free form. Indeed, the structure of residues 78-81 changes from helical (free) to disordered (bound) while the structure of residues 221-227 changes from extended to helical.  相似文献   

19.
Small ankyrin 1 (sAnk1), an integral protein of the sarcoplasmic reticulum encoded by the ANK1 gene, binds with nanomolar affinity to the C terminus of obscurin, a giant protein surrounding the contractile apparatus in striated muscle. We used site-directed mutagenesis to characterize the binding site on sAnk1, specifically addressing the role of two putative amphipathic, positively charged helices. We measured binding qualitatively by blot overlay assays and quantitatively by surface plasmon resonance and showed that both positively charged sequences are required for activity. We showed further that substitution of a lysine or arginine with an alanine or glutamate located at the same position along either of the two putative helices has similar inhibitory or stimulatory effects on binding and that the effects of a particular mutation depended on the position of the mutated amino acid in each helix. We modeled the structure of the binding region of sAnk1 by homology with ankyrin repeats of human Notch1, which have a similar pattern of charged and hydrophobic residues. Our modeling suggested that each of the two positively charged sequences forms pairs of amphipathic, anti-parallel alpha-helices flanked by beta-hairpin-like turns. Most of the residues in homologous positions along each helical unit have similar, though not identical, orientations. CD spectroscopy confirmed the alpha-helical content of sAnk1, approximately 33%, predicted by the model. Thus, structural and mutational studies of the binding region on sAnk1 for obscurin suggest that it consists of two ankyrin repeats with very similar structures.  相似文献   

20.
ATP-Binding Cassette transporters are ubiquitous membrane proteins that convert the energy from ATP-binding and hydrolysis into conformational changes of the transmembrane region to allow the translocation of substrates against their concentration gradient. Despite the large amount of structural and biochemical data available for this family, it is still not clear how the energy obtained from ATP hydrolysis in the ATPase domains is "transmitted" to the transmembrane domains. In this work, we focus our attention on the consequences of hydrolysis and inorganic phosphate exit in the maltose uptake system (MalFGK(2)E) from Escherichia coli. The prime goal is to identify and map the structural changes occurring during an ATP-hydrolytic cycle. For that, we use extensive molecular dynamics simulations to study three potential intermediate states (with 10 replicates each): an ATP-bound, an ADP plus inorganic phosphate-bound and an ADP-bound state. Our results show that the residues presenting major rearrangements are located in the A-loop, in the helical sub-domain, and in the "EAA motif" (especially in the "coupling helices" region). Additionally, in one of the simulations with ADP we were able to observe the opening of the NBD dimer accompanied by the dissociation of ADP from the ABC signature motif, but not from its corresponding P-loop motif. This work, together with several other MD studies, suggests a common communication mechanism both for importers and exporters, in which ATP-hydrolysis induces conformational changes in the helical sub-domain region, in turn transferred to the transmembrane domains via the "coupling helices".  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