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.
The superfamily of small, monomeric GTP-binding proteins, in Arabidopsis thaliana comprising 93 members, is classified into four families: Arf/Sar, Rab, Rop/Rac, and Ran families. All monomeric G proteins function as molecular switches that are activated by GTP and inactivated by the hydrolysis of GTP to GDP. GTP/GDP cycling is controlled by three classes of regulatory protein: guanine-nucleotide-exchange factors (GEFs), GTPase-activating proteins (GAPs), and guanine-nucleotide-dissociation inhibitors (GDIs). Proteins of Arf family are primarily involved in regulation of membrane traffic and organization of the cytoskeleton. Arf1/Sar1 proteins regulate the formation of vesicle coat at different steps in the exocytic and endocytic pathways. Rab GTPases are regulators of vesicular transport. They are involved in vesicle formation, recruitment of cytoskeletal motor proteins, and in vesicle tethering and fusion. Rop proteins serve as key regulators of cytoskeletal reorganization in response to extracellular signals. Several data have also shown that Rop proteins play additional roles in membrane trafficking and regulation of enzymes activity. Ran proteins are involved in nucleocytoplasmic transport.  相似文献   

2.
Interior and surface of monomeric proteins   总被引:47,自引:0,他引:47  
The solvent-accessible surface area (As) of 46 monomeric proteins is calculated using atomic co-ordinates from high-resolution and well-refined crystal structures. The As of these proteins can be determined to within 1 to 2% and that of their individual residues to within 10 to 20%. The As values of proteins are correlated with their molecular weight (Mr) in the range 4000 to 35,000: the power law As = 6.3 M0.73 predicts protein As values to within 4% on average. The average water-accessible surface is found to be 57% non-polar, 24% polar and 19% charged, with 5% root-mean-square variations. The molecular surface buried inside the protein is 58% non-polar, 39% polar and 4% charged. The buried surface contains more uncharged polar groups (mostly peptides) than the surface that remains accessible, but many fewer charged groups. On average, 15% of residues in small proteins and 32% in larger ones may be classed as "buried residues", having less than 5% of their surface accessible to the solvent. The accessibilities of most other residues are evenly distributed in the range 5 to 50%. Although the fraction of buried residues increases with molecular weight, the amino acid compositions of the protein interior and surface show no systematic variation with molecular weight, except for small proteins that are often very rich in buried cysteines. From amino acid compositions of protein surfaces and interiors we calculate an effective coefficient of partition for each type of residue, and derive an implied set of transfer free energy values. This is compared with other sets of partition coefficients derived directly from experimental data. The extent to which groups of residues (charged, polar and non-polar) are buried within proteins correlates well with their hydrophobicity derived from amino acid transfer experiments. Within these three groups, the correlation is low.  相似文献   

3.
The largest known monomeric globular proteins   总被引:5,自引:0,他引:5  
  相似文献   

4.
We have analyzed the buried water molecules and internal cavities in a set of 75 high-resolution, nonhomologous, monomeric protein structures. The number of hydrogen bonds formed between each water molecule and the protein varies from 0 to 4, with 3 being most common. Nearly half of the water molecules are found in pairs or larger clusters. Approximately 90% are shown to be associated with large cavities within the protein, as determined by a novel program, PRO_ACT. The total volume of a protein's large cavities is proportional to its molecular weight and is not dependent on structural class. The largest cavities in proteins are generally elongated rather than globular. There are many more empty cavities than hydrated cavities. The likelihood of a cavity being occupied by a water molecule increases with cavity size and the number of available hydrogen bond partners, with each additional partner typically stabilizing the occupied state by 0.6 kcal/mol.  相似文献   

5.
Adipocyte lipid-binding protein (ALBP or aP2) is an intracellular fatty acid-binding protein that is found in adipocytes and macrophages and binds a large variety of intracellular lipids with high affinity. Although intracellular lipids are frequently charged, biochemical studies of lipid-binding proteins and their interactions often focus most heavily on the hydrophobic aspects of these proteins and their interactions. In this study, we have characterized the effects of KCl on the stability and lipid binding properties of ALBP. We find that added salt dramatically stabilizes ALBP, increasing its Delta G of unfolding by 3-5 kcal/mol. At 37 degrees C salt can more than double the stability of the protein. At the same time, salt inhibits the binding of the fluorescent lipid 1-anilinonaphthalene-8-sulfonate (ANS) to the protein and induces direct displacement of the lipid from the protein. Thermodynamic linkage analysis of the salt inhibition of ANS binding shows a nearly 1:1 reciprocal linkage: i.e. one ion is released from ALBP when ANS binds, and vice versa. Kinetic experiments show that salt reduces the rate of association between ANS and ALBP while simultaneously increasing the dissociation rate of ANS from the protein. We depict and discuss the thermodynamic linkages among stability, lipid binding, and salt effects for ALBP, including the use of these linkages to calculate the affinity of ANS for the denatured state of ALBP and its dependence on salt concentration. We also discuss the potential molecular origins and potential intracellular consequences of the demonstrated salt linkages to stability and lipid binding in ALBP.  相似文献   

6.
A A Ismail  H H Mantsch 《Biopolymers》1992,32(9):1181-1186
The carboxylate-containing homopolypeptides poly(L-glutamate) [poly(Glu)] and poly(L-aspartate) [poly(Asp)] were found to form different types of ordered structures in the presence of poly(L-lysine) [poly(Lys)]. Mixing poly(Glu) with poly(Lys) in aqueous solution at neutral pH results in the instantaneous formation of a gel-like precipitate. The secondary structure of the gel precipitate can be best described as intermolecular antiparallel beta-strands, involving the backbone amide groups, as evidenced by the presence of characteristic amide I bands in the ir spectrum at 1684 and 1612 cm-1. Mixing poly(Asp) with poly(Lys) under identical conditions results in the formation of a fine precipitate with a different morphology. Examination of the ir spectrum of the precipitate revealed that unlike poly(Glu), poly(Asp) did not yield any discrete secondary structure upon precipitation with poly(Lys). Addition of solutions containing Ca2+ or Mg2+ to the poly(Glu)/poly(Lys) aggregates resulted in complete dissolution of the gel, with the disappearance of the ir bands characteristic of the intermolecular hydrogen-bonded network. The results demonstrate the importance of salt bridges in establishing strong hydrogen bonds between the backbone amide groups. Reaggregation occurred upon heating the poly(Glu)/poly(Lys) mixture in the presence of Ca2+, but not in the presence of Mg2+ ions. In the presence of Ca2+ ions, aggregation and formation of an extended hydrogen-bonded network occurred upon heating. The aggregates formed upon heating poly(Glu)/poly(Lys) in the presence of Ca2+ were attributed solely to complexation of Ca2+ to the carboxylate groups of poly(Glu) with poly(Lys) remaining free in solution. Dissolution of the aggregate could be accomplished through addition of Mg2+ at room temperature.  相似文献   

7.
8.
A general phenomenological model is proposed for the estimation of the influence of the formation of complexes with ligands on thermal stability of proteins. In this model the reversible processes of unfolding-refolding and of association-dissociation of protein-ligand complexes and of the irreversible chemical degradation of the unfolded protein were analyzed jointly. By using certain approximations, the analytical expressions for both the thermodynamic and kinetic stabilization are obtained. Two thermodynamic and four kinetic regimes of stabilization and destabilization can exist in such system. Each thermodynamic regime appears to be compatible with three different kinetic regimes. The effect of the formation of complexes on thermodynamic and kinetic stability of the protein is determined by the degrees of binding of the ligand to the folded and unfolded protein species and by the rates of irreversible degradation of free protein and protein in complex.  相似文献   

9.
Scores calculated from intermolecular contacts of proteins in the crystalline state are used to differentiate monomeric and homodimeric proteins, by classification into two categories separated by a cut-off score value. The generalized classification error is estimated by using bootstrap re-sampling on a nonredundant set of 172 water-soluble proteins whose prevalent quaternary state in solution is known to be either monomeric or homodimeric. A statistical potential, based on atom-pair frequencies across interfaces observed with homodimers, is found to yield an error rate of 12.5%. This indicates a small but significant improvement over the measure of solvent accessible surface area buried in the contact interface, which achieves an error rate of 15.4%. A further modification of the latter parameter relating the two most extensive contacts of the crystal results in an even lower error rate of 11.1%.  相似文献   

10.
The solvent-exposed residue Ala32 in the second alpha-helix of barnase was replaced by all other naturally occurring amino acids and the concomitant effects on the protein stability were determined. The results are assumed to reflect both the distinct conformational preferences of the different amino acids and also possible intrahelical interactions. The conformational preferences may be fully rationalized by invoking only a few physical principles. The results agree well with recently experimentally determined rank-order of helix-forming tendencies determined on a model peptide. There is very weak correlation between the results and the experimental host-guest values. There is a weak correlation between our results and the statistical helix propensities and a slightly better correlation with the positional-dependent statistical parameters of J. S. Richardson, and D. C. Richardson.  相似文献   

11.
Adrenodoxin and the mutants at the positions T54, H56, D76, Y82, and C95, as well as the deletion mutants 4-114 and 4-108, were studied by high-sensitivity scanning microcalorimetry, limited proteolysis, and absorption spectroscopy. The mutants show thermal transition temperatures ranging from 46 to 56 degrees C, enthalpy changes from 250 to 370 kJ/mol, and heat capacity change delta Cp = 7.28 +/- 0.67 kJ/mol/K, except H56R. The amino acid replacement H56R produces substantial local changes in the region around positions 56 and Y82, as indicated by reduced heat capacity change (delta Cp = 4.29 +/- 0.37 kJ/mol/K) and enhanced fluorescence. Deletion mutant 4-108 is apparently more stable than the wild type, as judged by higher specific denaturation enthalpy and resistance toward proteolytic degradation. No simple correlation between conformational stability and functional properties could be found.  相似文献   

12.
The structure and stability of the fluorescent protein monomeric Kusabira Orange (mKO), a GFP-like protein, was studied under different pressure levels and in different chemical environments. At different pH values (between pH 7.4 and pH 4.0) and under a pressure up to 600 MPa (at 25 °C), mKO did not show significant fluorescence spectral changes, indicating a structural stability of the protein. In more extreme chemical conditions (at pH 4.0 in the presence of 0.8 M guanidine hydrochloride), a marked reduction of mKO fluorescence intensity emission was observed at pressures above 300 MPa. This fluorescence emission quenching may be due to the loss of the intermolecular bonds and, consequently, to the destructuration of the mKO chromophore structure. Since the electrostatic and hydrophobic interactions as well as the salt bridges present in proteins are usually perturbed under high pressure, the reduction of mKO fluorescence intensity emission is associated to the perturbation of the protein salt bridges network.  相似文献   

13.
Salt stress proteins induced in Listeria monocytogenes   总被引:3,自引:0,他引:3  
The ability of Listeria monocytogenes to tolerate salt stress is of particular importance, as this pathogen is often exposed to such environments during both food processing and food preservation. In order to understand the survival mechanisms of L. monocytogenes, an initial approach using two-dimensional polyacrylamide gel electrophoresis was performed to analyze the pattern of protein synthesis in response to salt stress. Of 400 to 500 visible proteins, the synthesis of 40 proteins (P < 0.05) was repressed or induced at a higher rate during salt stress. Some of the proteins were identified on the basis of mass spectrometry or N-terminal sequence analysis and database searching. Twelve proteins showing high induction after salt stress were similar to general stress proteins (Ctc and DnaK), transporters (GbuA and mannose-specific phosphotransferase system enzyme IIAB), and general metabolism proteins (alanine dehydrogenase, CcpA, CysK, EF-Tu, Gap, GuaB, PdhA, and PdhD).  相似文献   

14.
Salt bridges are frequently observed in protein structures. Because the energetic contribution of salt bridges is strongly dependent on the environmental context, salt bridges are believed to contribute to the structural specificity rather than the stability. To test the role of salt bridges in enhancing structural specificity, we investigated the contribution of a salt bridge to the energetics of native‐state partial unfolding in a cysteine‐free version of Escherichia coli ribonuclease H (RNase H*). Thermolysin cleaves a protruding loop of RNase H* through transient partial unfolding under native conditions. Lys86 and Asp108 in RNase H* form a partially buried salt bridge that tethers the protruding loop. Investigation of the global stability of K86Q/D108N RNase H* showed that the salt bridge does not significantly contribute to the global stability. However, K86Q/D108N RNase H* is greatly more susceptible to proteolysis by thermolysin than wild‐type RNase H* is. The free energy for partial unfolding determined by native‐state proteolysis indicates that the salt bridge significantly increases the energy for partial unfolding by destabilizing the partially unfolded form. Double mutant cycles with single and double mutations of the salt bridge suggest that the partially unfolded form is destabilized due to a significant decrease in the interaction energy between Lys86 and Asp108 upon partial unfolding. This study demonstrates that, even in the case that a salt bridge does not contribute to the global stability, the salt bridge may function as a gatekeeper against partial unfolding that disturbs the optimal geometry of the salt bridge.  相似文献   

15.
The control of plant growth, differentiation, and development is considered in relation to the involvement of monomeric GTP-binding proteins (mG-proteins) in the extra-and intracellular signal transduction. The principal attention is paid to Rop mG-proteins, unique small GTPases of eukaryotic cells functioning during various developmental stages of plants, from pollen tube and root hair growth to plant responses to biotic and abiotic stresses.  相似文献   

16.
The axial structure of the bacterial flagellum is composed of many different proteins, such as hook protein and flagellin, and each protein forms a short or long axial segment one after another in a well-defined order along the axis. Under physiological conditions, most of these proteins are stable in the monomeric state in solution, and spontaneous polymerization appears to be suppressed, as demonstrated clearly for flagellin, probably to avoid undesirable self-assembly in the cytoplasmic space. However, no systematic studies of the possible associations between monomeric axial proteins in solution have been carried out. We therefore studied self and cross-association between hook protein, flagellin and three hook-associated proteins, HAP1, HAP2 and HAP3, in all possible pairs, by gel-filtration and analytical centrifugation, and found interactions in the following two cases only. Flagellin facilitated HAP3 aggregation into beta-amyloid-like filaments, but without stable binding between the two. Addition of HAP3 to HAP2 resulted in disassembly of preformed HAP2 decamers and formation of stable HAP2-HAP3 heterodimers. HAP2 missing either of its disordered terminal regions did not form the heterodimer, whereas HAP3 missing either of its disordered terminal regions showed stable heterodimer formation. This polarity in the heterodimer interactions suggests that the interactions between HAP2 and HAP3 in solution are basically the same as those in the flagellar axial structure. We discuss these results in relation to the assembly mechanism of the flagellum.  相似文献   

17.
C F Burant  G I Bell 《Biochemistry》1992,31(42):10414-10420
Four facilitative glucose transporters isoforms, GLUT1/erythrocyte, GLUT2/liver, GLUT3/brain, and GLUT4/muscle-fat, as well as chimeric transporter proteins were expressed in Xenopus oocytes, and their properties were studied. The relative Km's of the transporters for 2-deoxyglucose were GLUT3 (Km = 1.8 mM) > GLUT4 (Km = 4.6 mM) > GLUT1 (Km = 6.9 mM) > GLUT2 (Km = 17.1 mM). In a similar fashion, the uptake of 2-deoxyglucose by GLUT1-, GLUT2-, and GLUT3-expressing oocytes was inhibited by a series of unlabeled hexoses and pentoses and by cytochalasin B in a similar hierarchical order. To determine if the functional unit of the glucose transporter was a monomer or higher-order multimer, the high-affinity transporter GLUT3 was coexpressed with either the low-affinity GLUT2 or a GLUT3 mutant which contained a transport inactivating Trp410-->Leu substitution. In oocytes expressing both GLUT2 and GLUT3, the transport activity associated with each transporter isoform could be distinguished kinetically. Similarly, there was no alteration in the kinetic parameters of GLUT3, or the ability of glucose or cytochalasin B to inhibit 2-deoxyglucose uptake, when coexpressed with up to a 3-fold greater amount of functionally inactive mutant of GLUT3. These studies suggest that the family of glucose transporters have similar binding sites which may be in the form of a functional monomeric unit when expressed in Xenopus oocytes.  相似文献   

18.
A theoretical study on the stability of the salt bridges in the gas phase, in solution, and in the interior of proteins is presented. The study is mainly focused on the interaction between acetate and methylguanidinium ions, which were used as model compounds for the salt bridge between Asp (Glu) and Arg. Two different solvents (water and chloroform) were used to analyze the effect of varying the dielectric constant of the surrounding media on the salt bridge interaction. Calculations in protein environments were performed by using a set of selected protein crystal structures. In all cases attention was paid to the difference in stability between the ion pair and neutral hydrogen-bonded forms. Comparison of the results determined in the gas phase and in solution allows us to stress the large influence of the environment on the binding process, as well as on the relative stability between the ionic and neutral complexes. The high anisotropy of proteins and the local microenvironment in the interior of proteins make a decisive contribution in modulating the energetics of the salt bridge. In general, the formation of salt bridges in proteins is not particularly favored, with the ion pair structure being preferred over the interaction between neutral species. Proteins 32:67–79, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

19.
Due to the low absorbance in the far-red (FR) and near-infrared (NIR) “optical window”, NIR fluorescent proteins (FPs) are powerful tools for deep imaging. Here, we report three new, highly bright NIR FPs termed BDFP1.8, BDFP1.8:1.8 (tandem BDFP1.8) and BDFP1.9, which evolved from a previously reported FR FP, BDFP1.6: a derivative of ApcF2 from Chroococcidiopsis thermalis sp. PCC7203. ApcF2 binds phycocyanobilin (PCB) non-covalently, while BDFPs, the derivatives of ApcF2, can bind biliverdin (BV) covalently. We identified that dimeric BDFP1.8 and monomeric BDFP1.8:1.8 have a 2.4-and 4.4-fold higher effective brightness, respectively, than iRFP720, which has the highest effective brightness among the reported NIR FPs. Monomeric DBFP1.9 (17 kDa) has one of the smallest masses among highly bright FPs in the FR and NIR regions. Enhancing the affinity between the apo-proteins and the BV chromophore is an effective method to improve the effective brightness of biliprotein FPs. Moreover, BDFP1.8 and 1.9 exhibit higher stability to temperature, pH and light than iRFP720. Finally, the highly bright NIR BDFP1.8 together with FR BDFP1.6 could effectively biolabel cells in dual colors.  相似文献   

20.
Zhang M  Chang H  Zhang Y  Yu J  Wu L  Ji W  Chen J  Liu B  Lu J  Liu Y  Zhang J  Xu P  Xu T 《Nature methods》2012,9(7):727-729
Monomeric (m)Eos2 is an engineered photoactivatable fluorescent protein widely used for super-resolution microscopy. We show that mEos2 forms oligomers at high concentrations and forms aggregates when labeling membrane proteins, limiting its application as a fusion partner. We solved the crystal structure of tetrameric mEos2 and rationally designed improved versions, mEos3.1 and mEos3.2, that are truly monomeric, are brighter, mature faster and exhibit higher photon budget and label density.  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