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.
Integral membrane proteins often contain proline residues in their alpha-helical transmembrane (TM) fragments, which may strongly influence their folding and association. Pro-scanning mutagenesis of the helical domain of glycophorin A (GpA) showed that replacement of the residues located at the center abrogates helix packing while substitution of the residues forming the ending helical turns allows dimer formation. Synthetic TM peptides revealed that a point mutation of one of the residues of the dimerization motif (L75P) located at the N-terminal helical turn of the GpA TM fragment, adopts a secondary structure and oligomeric state similar to the wild-type sequence in detergents. In addition, both glycosylation mapping in biological membranes and molecular dynamics showed that the presence of a proline residue at the lipid/water interface has as an effect the extension of the helical end. Thus, helix packing can be an important factor that determines appearance of proline in TM helices. Membrane proteins might accumulate proline residues at the two ends of their TM segments in order to modulate the exposition of key amino acid residues at the interface for molecular recognition events while allowing stable association and native folding.  相似文献   

2.
Chen Z  Xu Y 《Proteins》2006,62(2):539-552
The energetics and stability of the packing of transmembrane helices were investigated by Monte Carlo simulations with the replica-exchange method. The helices were modeled with a united atom representation, and the CHARMM19 force field was employed. Based on known experimental structures of membrane proteins, an implicit knowledge-based potential was developed to describe the helix-membrane interactions at the residue level, whose validity was tested through prediction of the orientations when single helices were inserted into a membrane. Two systems were studied in this article, namely the glycophorin A dimer, and helices A and B of Bacteriorhodopsin. For the glycophorin A dimer, the most stable structure (0.5 A away from the experimental structure) is mainly stabilized by the favorable helix-helix interactions, and has the most population regardless of the helix-membrane interaction. However, for helices A and B of Bacteriorhodopsin, it was found that the packing determined by helix-helix interactions is nonspecific, and a native-like structure (0.2 A from the experimental one) can be identified from several structural analogs as the most stable one only after applying the membrane potential. Our results suggest that the contribution from the helix-membrane interaction could be critical in the correct packing of transmembrane helices in the membrane.  相似文献   

3.
Despite some promising progress in the understanding of membrane protein folding and assembly, there is little experimental information regarding the thermodynamic stability of transmembrane helix interactions and even less on the stability of transmembrane helix-helix interactions in a biological membrane. Here we describe an approach that allows quantitative measurement of transmembrane helix interactions in a biological membrane, and calculation of changes in the interaction free energy resulting from substitution of single amino acids. Dimerization of several variants of the glycophorin A transmembrane domain are characterized and compared to the wild-type (wt) glycophorin A transmembrane helix dimerization. The calculated DeltaDeltaG(app) values are further compared with values found in the literature. In addition, we compare interactions between the wt glycophorin A transmembrane domain and helices in which critical glycine residues are replaced by alanine or serine, respectively. The data demonstrate that replacement of the glycine residues by serine is less destabilizing than replacement by alanine with a DeltaDeltaG(app) value of about 0.4 kcal/mol. Our study comprises the first measurement of a transmembrane helix interaction in a biological membrane, and we are optimistic that it can be further developed and applied.  相似文献   

4.
Immunoglobulin (Ig) domains are the most prevalent protein domain structure and share a highly conserved folding pattern; however, this structural family of proteins is also the most diverse in terms of biological roles and tissue expression. Ig domains vary significantly in amino acid sequence but share a highly conserved tryptophan in the hydrophobic core of this beta‐stranded protein. Palladin is an actin binding and bundling protein that has five Ig domains and plays an important role in normal cell adhesion and motility. Mutation of the core tryptophan in one Ig domain of palladin has been identified in a pancreatic cancer cell line, suggesting a crucial role for this sole tryptophan in palladin Ig domain structure, stability, and function. We found that actin binding and bundling was not completely abolished with removal of this tryptophan despite a partially unfolded structure and significantly reduced stability of the mutant Ig domain as shown by circular dichroism investigations. In addition, this mutant palladin domain displays a tryptophan‐like fluorescence attributed to an anomalous tyrosine emission at 341 nm. Our results indicate that this emission originates from a tyrosinate that may be formed in the excited ground state by proton transfer to a nearby aspartic acid residue. Furthermore, this study emphasizes the importance of tryptophan in protein structural stability and illustrates how tyrosinate emission contributions may be overlooked during the interpretation of the fluorescence properties of proteins.  相似文献   

5.
BNIP3 is a mitochondrial 19-kDa proapoptotic protein, a member of the Bcl-2 family. It has a single COOH-terminal transmembrane (TM) alpha-helical domain, which is required for membrane targeting, proapoptotic activity, hetero- and homo-dimerization in membrane. The role and the molecular details of association of TM helices of BNIP3 are yet to be established. Here, we present a molecular modeling study of helix interactions in its membrane domain. The approach combines Monte Carlo conformational search in an implicit hydrophobic slab followed by molecular dynamics simulations in a hydrated full-atom lipid bilayer. The former technique was used for exhaustive sampling of the peptides' conformational space and for generation of putative "native-like" structures of the dimer. The latter ones were taken as realistic starting points to assess stability and dynamic behavior of the complex in explicit lipid-water surrounding. As a result, several groups of tightly packed right-handed structures of the dimer were proposed. They have almost similar helix-helix interface, which includes the motif A(176)xxxG(180)xxxG(184) and agrees well with previous mutagenesis data and preliminary NMR analysis. Molecular dynamics simulations of these structures reveal perfect adaptation of most of them to heterogeneous membrane environment. A remarkable feature of the predicted dimeric structures is the occurrence of a cluster of H-bonded histidine 173 and serines 168 and 172 on the helix interface, near the N-terminus. Because of specific polar interactions between the monomers, this part of the dimer has no such dense packing as the C-terminal one, thus allowing penetration of water from the extramembrane side into the membrane interior. We propose that the ionization state of His(173) can mediate structural and dynamic properties of the dimer. This, in turn, may be related to pH-dependent proapoptotic activity of BNIP3, which is triggering on by acidosis appearing under hypoxic conditions.  相似文献   

6.
7.
Visible and infrared spectra of bacteriorhodopsin films under different humidities at room and low temperatures are investigated. On dehydration of purple membranes at room temperatures an additional chromophore state with the absorption band at 506 nm is revealed. The photocycle of purple membranes in the dry state is devoid of the 550 nm intermediate and involves the long-lived intermediate at 412 nm. As water is removed, the 550 nm intermediate becomes undetectable. The analysis of the infrared spectra shows that dehydration does not affect the ordering of the main network of the interpeptide hydrogen bonds which stabilizes the -helical conformation (slightly distorted in the initial humid dark- and light-adapted state); light adaptation (cis-trans isomerization) of bacteriorhodopsin results in an increase of sorbed water in purple membranes. Dehydration of purple membranes decreases the reaction rate of cis-trans isomerization.  相似文献   

8.
T he T yrosine and tryptophan hydroxylase enzymes have been proposed as rate-limiting steps in the biosynthesis of catecholamines and 5-hydroxytryptamine (5-HT), respectively; thus under normal physiological conditions the rate of amine synthesis appears to be controlled by the activity of these hydroxylase enzymes (see U denfriend , 1966; L ovenberg , J equier and S joerdsma , 1968). Subtle changes in neuronal activity may result not in changes in the levels of the amine neurotransmitters, but rather in alteration in their production and metabolism without measurable change in their levels. Previous studies of the effect of hypoxia on monoamines have dealt with amine levels, but there have been no studies of the effect of lowered oxygen on the synthesis of these substances.  相似文献   

9.
We used frequency-domain measurements of fluorescence resonance energy transfer to measure the distribution of distances between Trp-19 of melittin and a 1-dimethylamino-5-sulfonylnaphthalene (dansyl) residue on the N-terminal-alpha-amino group. Distance distributions were obtained for melittin free in solution and when complexed with calmodulin (CaM), troponin C (TnC), or palmitoyloleoyl-L-alpha-phosphatidylcholine (POPC) vesicles. A wide range of donor (Trp-19)-to-acceptor (dansyl) distances was found for free melittin, which is consistent with that expected for the random coil state, characterized by a Gaussian width (full width at half maxima) of 28.2 A. In contrast, narrow distance distributions were found for melittin complexed with CaM, 8.2 A, or with POPC vesicles, 4.9 A. A somewhat wider distribution was found for the melittin complex with TnC, 12.8 A, suggesting the presence of heterogeneity in the mode of binding between melittin and TnC. For all the complexes the mean Trp-19 to dansyl distance was near 20 A. This value is somewhat smaller than expected for the free alpha-helical state of melittin, suggesting that binding with CaM or TnC results in a modest decrease in the length of the melittin molecule.  相似文献   

10.
11.
Cellulases are increasingly being used for industrial purposes, particularly in washing powders, yet little is known of the factors governing the stability of proteins in detergent solutions. We present a comparative analysis of the behavior of the cellulase Cel45 from Humicola insolens in the presence of the denaturant guanidinium chloride and the anionic detergent C12-LAS. Although Cel45 unfolds in GdmCl according to a simple two-state model under equilibrium conditions, it accumulates a transient intermediate during refolding. The four disulfide bonds do not contribute detectably to the stability of the native state. Cel45 is unfolded by very low concentrations of C12-LAS (1-4 mM). An analysis of 16 mutants of Cel45 shows a very weak correlation between unfolding rates in denaturant and detergent; mutants that have the same unfolding rate in GdmCl (within a factor of 1.5) vary 1,000-fold in their unfolding rates in C12-LAS. The data support a simple model for unfolding by detergent, in which the introduction of positive charges or removal of negative charges greatly increases detergent sensitivity, while interactions with the hydrophobic detergent tail contribute to a smaller extent. This implies that different detergent-mediated unfolding pathways exist, whose accessibilities depend on individual residues. Double-mutant cycles reveal that mutations in two proximal residues lead to repulsion and a destabilization greater than the sum of the individual mutations as measured by GdmCl denaturation, but they also reduce the affinity for LAS and therefore actually stabilize the protein relative to wild-type. Ligands that interact strongly with the denatured state may therefore alter the unfolding process.  相似文献   

12.
Despite the large number of publications on three‐helix protein folding, there is no study devoted to the influence of handedness on the rate of three‐helix protein folding. From the experimental studies, we make a conclusion that the left‐handed three‐helix proteins fold faster than the right‐handed ones. What may explain this difference? An important question arising in this paper is whether the modeling of protein folding can catch the difference between the protein folding rates of proteins with similar structures but with different folding mechanisms. To answer this question, the folding of eight three‐helix proteins (four right‐handed and four left‐handed), which are similar in size, was modeled using the Monte Carlo and dynamic programming methods. The studies allowed us to determine the orders of folding of the secondary‐structure elements in these domains and amino acid residues which are important for the folding. The obtained data are in good correlation with each other and with the experimental data. Structural analysis of these proteins demonstrated that the left‐handed domains have a lesser number of contacts per residue and a smaller radius of cross section than the right‐handed domains. This may be one of the explanations of the observed fact. The same tendency is observed for the large dataset consisting of 332 three‐helix proteins (238 right‐ and 94 left‐handed). From our analysis, we found that the left‐handed three‐helix proteins have some less‐dense packing that should result in faster folding for some proteins as compared to the case of right‐handed proteins.Proteins 2013; © 2013 Wiley Periodicals, Inc.  相似文献   

13.
Wild-type flavocytochrome b2 (L-lactate dehydrogenase) from Saccharomyces cerevisiae, as well as a number of its point mutants, can be expressed to a reasonable level as recombinant proteins in Escherichia coli (20-25 mg per liter culture) with a full complement of prosthetic groups. At the same expression level, active-site mutants Y254L and D282N, on the other hand, were obtained with an FMN/heme ratio significantly less than unity, which could not be raised by addition of free FMN. Evidence is provided that the flavin deficit is due to incomplete prosthetic group incorporation during biosynthesis. Flavin-free and holo-forms for both mutants could be separated on a Blue-Trisacryl M column. The far-UV CD spectra of the two forms of each mutant protein were very similar to one another and to that of the wild-type enzyme, suggesting the existence of only local conformational differences between the active holo-enzymes and the nonreconstitutable flavin-free forms. Selective proteolysis with chymotrypsin attacked the same bond for the two mutant holo-enzymes as in the wild-type one, in the protease-sensitive loop. In contrast, for the flavin-free forms of both mutants, cleavage occurred at more than a single bond. Identification of the cleaved bonds suggested that the structural differences between the mutant flavin-free and holo-forms are located mostly at the C-terminal end of the barrel, which carries the prosthetic group and the active site. Altogether, these findings suggest that the two mutations induce an alteration of the protein-folding process during biosynthesis in E. coli; as a result, the synchrony between folding and flavin insertion is lost. Finally, a preliminary kinetic characterization of the mutant holo-forms showed the Km value for lactate to be little affected; kcat values fell by a factor of about 70 for the D282N mutant and of more than 500 for the Y254L mutant, compared to the wild-type enzyme.  相似文献   

14.
The kinetic folding mechanism for Escherichia coli dihydrofolate reductase postulates two distinct types of transient intermediates. The first forms within 5 ms and has substantial secondary structure but little stability. The second is a set of four species that appear over the course of several hundred milliseconds and have secondary structure, specific tertiary structure, and significant stability (Jennings PA, Finn BE, Jones BE, Matthews CR, 1993, Biochemistry 32:3783-3789). Pulse labeling hydrogen exchange experiments were performed to determine the specific amide hydrogens in alpha-helices and beta-strands that become protected from exchange through the formation of stable hydrogen bonds during this time period. A significant degree of protection was observed for two subsets of the amide hydrogens within the dead time of this experiment (6 ms). The side chains of one subset form a continuous nonpolar strip linking six of the eight strands in the beta-sheet. The other subset corresponds to a nonpolar cluster on the opposite face of the sheet and links three of the strands and two alpha-helices. Taken together, these data demonstrate that the complex strand topology of this eight-stranded sheet can be formed correctly within 6 ms. Measurement of the protection factors at three different folding times (13 ms, 141 ms, and 500 ms) indicates that, of the 13 amide hydrogens displaying significant protection within 6 ms, 8 exhibit an increase in their protection factors from approximately 5 to approximately 50 over this time range; the remaining five exhibit protection factors > 100 at 13 ms. Only approximately half of the population of molecules form this set of stable hydrogen bonds. Thirteen additional hydrogens in the beta-sheet become protected from exchange as the set of native conformers appear, suggesting that the stabilization of this network reflects the global cooperativity of the folding reaction.  相似文献   

15.
Takeuchi H 《Biopolymers》2003,72(5):305-317
The Raman spectrum of a protein contains a wealth of information on the structure and interaction of the protein. To extract the structural information from the Raman spectrum, it is necessary to identify and interpret the marker bands that reflect the structure and interaction in the protein. Recently, new Raman structural markers have been proposed for the tryptophan and histidine side chains by examining the spectra-structure correlations of model compounds. Raman structural markers are now available for the conformation, hydrogen bonding, hydrophobic interaction, and cation-pi interaction of the indole ring of Trp. For His, protonation, tautomerism, and metal coordination of the imidazole ring can be studied by using Raman markers. The high-resolution X-ray crystal structures of proteins provide the basis for testing and modifying the Raman structural markers of Trp and His. The structures derived from Raman spectra are generally consistent with the X-ray crystal structures, giving support for the applicability of most Raman structural makers. Possible modifications and limitations to some marker bands are also discussed.  相似文献   

16.
Folding of outer membrane proteins (OMPs) has been studied extensively in vitro. However, most of these studies have been conducted in dilute buffer solution, which is different from the crowded environment in the cell periplasm, where the folding and membrane insertion of OMPs actually occur. Using OmpA and OmpT as model proteins and Ficoll 70 as the crowding agent, here we investigated the effect of the macromolecular crowding condition on OMP membrane insertion. We found that the presence of Ficoll 70 significantly slowed down the rate of membrane insertion of OmpA while had little effect on those of OmpT. To investigate if the soluble domain of OmpA slowed down membrane insertion in the presence of the crowding agent, we created a truncated OmpA construct that contains only the transmembrane domain (OmpA171). In the absence of crowding agent, OmpA171 refolded at a similar rate as OmpA, although with decreased efficiency. However, under the crowding condition, OmpA171 refolded significantly faster than OmpA. Our results suggest that the periplasmic domain slows down the rate, while improves the efficiency, of OmpA folding and membrane insertion under the crowding condition. Such an effect was not obvious when refolding was studied in buffer solution in the absence of crowding.  相似文献   

17.
Ashish Shelar  Manju Bansal 《Proteins》2014,82(12):3420-3436
α‐helices are amongst the most common secondary structural elements seen in membrane proteins and are packed in the form of helix bundles. These α‐helices encounter varying external environments (hydrophobic, hydrophilic) that may influence the sequence preferences at their N and C‐termini. The role of the external environment in stabilization of the helix termini in membrane proteins is still unknown. Here we analyze α‐helices in a high‐resolution dataset of integral α‐helical membrane proteins and establish that their sequence and conformational preferences differ from those in globular proteins. We specifically examine these preferences at the N and C‐termini in helices initiating/terminating inside the membrane core as well as in linkers connecting these transmembrane helices. We find that the sequence preferences and structural motifs at capping (Ncap and Ccap) and near‐helical (N' and C') positions are influenced by a combination of features including the membrane environment and the innate helix initiation and termination property of residues forming structural motifs. We also find that a large number of helix termini which do not form any particular capping motif are stabilized by formation of hydrogen bonds and hydrophobic interactions contributed from the neighboring helices in the membrane protein. We further validate the sequence preferences obtained from our analysis with data from an ultradeep sequencing study that identifies evolutionarily conserved amino acids in the rat neurotensin receptor. The results from our analysis provide insights for the secondary structure prediction, modeling and design of membrane proteins. Proteins 2014; 82:3420–3436. © 2014 Wiley Periodicals, Inc.  相似文献   

18.
Fluorescence quenching by a series of spin-labelled fatty acids is used to map the transverse disposition of tryptophan residues in bacteriorhodopsin (the sole protein in the purple membranes of Halobacterium halobium). A new method of data analysis is employed which takes into account differences in the uptake of the quenchers into the membrane. Energy transfer from tryptophan to a set of n-(9-anthroyloxy) fatty acids is used as a second technique to confirm the transverse map of tryptophan residues revealed by the quenching experiments. The relative efficiencies of quenching and energy transfer obtained experimentally are compared with those predicted on the basis of current models of bacteriorhodopsin structure. Most of the tryptophan fluorescence is located near the surface of the purple membrane. When the retinal chromophore of bacteriorhodopsin is removed, tryptophan residues deep in the membrane become fluorescent. These results indicate that the deeper residues transfer their energy to retinal in the native membrane. The retinal moiety is therefore located deep within the membrane rather than at the membrane surface.  相似文献   

19.
The amino terminal dimerization/docking domain and the two-tandem, carboxy-terminal cAMP-binding domains (A and B) of cAMP-dependent protein kinase regulatory (R) subunits are connected by a variable linker region. In addition to providing a docking site for the catalytic subunit, the linker region is a major source of sequence diversity between the R-subunit isoforms. The RIIbeta isoform uniquely contains two endogenous tryptophan residues, one at position 58 in the linker region and the other at position 243 in cAMP-binding domain A, which can act as intrinsic reporter groups of their dynamics and microenvironment. Two single-point mutations, W58F and W243F, allowed the local environment of each Trp to be probed using steady-state and time-resolved fluorescence techniques. We report that: (a) the tryptophan fluorescence of the wild-type protein largely reflects Trp243 emission; (2) cAMP selectively quenches Trp243 and thus acts as a cAMP sensor; (3) Trp58 resides in a highly solvated, unstructured, and mobile region of the protein; and (4) Trp243 resides in a stable, folded domain and is relatively buried and rigid within the domain. The use of endogenous Trp residues presents a non-perturbing method for studying R-subunit subdomain characteristics in addition to providing the first biophysical data on the RIIbeta linker region.  相似文献   

20.
The tumor suppresser protein p53 has been called the “guardian of the genome.” DNA damage induces p53 to either halt the cell cycle, allowing for repair, or initiate apoptosis. P53 is mutated in over 50% of human tumors and it has been proposed that many tumorigenic mutations are deleterious to p53 because they induce local unfolding. To explore this hypothesis, peptide models have been developed to study tumorigenic mutations in the H2 helix of the p53 core domain. This helix is rich with charged residues and is a key component of the DNA binding region. A 16‐residue peptide corresponding to the H2 wild‐type sequence extended with an Ala‐rich C‐terminus was synthesized and studied by 1H‐nmr (500 MHz) and CD. The nmr studies demonstrate that this peptide adopts helical structure in solution. Six additional peptides corresponding to subtle tumorigenic mutations were synthesized and CD was used to assess the relative stability of these “mutant analogues.” All six mutations studied are destabilizing relative to the wild type, with ΔΔG values in the range of 0.26 to 1.35 kcal mol−1. Surprisingly, substitution of Asp 281 with Ala resulted in a peptide with the greatest destabilization even though Ala possesses the largest helix propensity of the common 20 amino acids. Because this helix appears to be stabilized mainly by local electrostatics, we conclude that its structure is susceptible to even the most conservative mutations. These results provide support for the hypothesis that tumorigenic mutations induce local unfolding of p53. © 1999 John Wiley & Sons, Inc. Biopoly 49: 215–224, 1999  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