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.
Ichimaru T  Kikuchi T 《Proteins》2003,51(4):515-530
It is a general notion that proteins with very similar three-dimensional structures would show very similar folding kinetics. However, recent studies reveal that the folding kinetic properties of some proteins contradict this thought (i.e., the members in a same protein family fold through different pathways). For example, it has been reported that some beta-proteins in the intracellular lipid-binding protein family fold through quite different pathways (Burns et al., Proteins 1998;33:107-118). Similar differences in folding kinetics are also observed in the members of the globin family (Nishimura et al., Nat Struct Biol 2000;7:679-686). In our study, we examine the possibility of predicting qualitative differences in folding kinetics of the intracellular lipid-binding proteins and two globin proteins (i.e., myoglobin and leghemoglobin). The problem is tackled by means of a contact map based on the average distance statistics between residues, the Average Distance Map (ADM), as constructed from sequence. The ADMs for the three proteins show overall similarity, but some local differences among maps are also observed. Our results demonstrate that some properties of the protein folding kinetics are consistent with local differences in the ADMs. We also discuss the general possibility of predicting folding kinetics from sequence information.  相似文献   

2.
Concrete structural features of integrin alpha(IIb)beta(3) on the surface of platelets (at rest and after activation) have been obtained from epitope maps based on cross-competition among monoclonal antibodies directed against the alpha(IIb) subunit calf-2 domain and the beta(3) subunit betaA domain of alpha(IIb)beta(3). At rest, the observed intersubunit interface is formed by the sequence stretches beta(3)-(150-216), alpha(IIb) light chain-(1-92), and alpha(IIb) heavy chain-(826-856); and the alpha(IIb) interchain interface is formed by the two latter sequence stretches, disulfide-bonded between alpha(IIb) heavy chain Cys(826) and alpha(IIb) light chain Cys(9). These structural features agree with those observed in the alpha(IIb)beta(3) rudimentary connectivity map in solution and with the alpha(v)beta(3) V-shaped crystal structure (Xiong, J.-P., Zhang, R., Dunker, R., Scott, D. L., Joachimiak, A., Goodman, S. L., and Arnaout, M. A. (2001) Science 294, 339-345), but they disagree with the domain disposition suggested by the actual ultrastructural model. The epitope maps in platelets activated by ADP, thrombin receptor activation peptide, and arachidonic acid differ not only from those in platelets at rest, but also among themselves. The structural rearrangements observed confirm the presence in activated platelets of the crystallographically observed knee and argue against the switchblade mechanism proposed for activation (Beglova, N., Blacklow, S. C., Takagi, J., and Springer, T. A. (2002) Nat. Struct. Biol. 9, 282-287), demonstrate the existence of alpha(IIb)beta(3) agonist-specific activation states, explain the specificity for ligand binding and functional inhibition for some agonists, and predict the existence of agonist-specific final effectors and receptor activation mechanisms. The distinct non-reciprocal competition patterns observed at rest and after activation support the agonist-specific activation states and the existence of intrasubunit and intersubunit allosteric effects, previously proposed as the mechanism for alpha(IIb)beta(3) transmembrane activation.  相似文献   

3.
Wilkens S  Borchardt D  Weber J  Senior AE 《Biochemistry》2005,44(35):11786-11794
A critical point of interaction between F(1) and F(0) in the bacterial F(1)F(0)-ATP synthase is formed by the alpha and delta subunits. Previous work has shown that the N-terminal domain (residues 3-105) of the delta subunit forms a 6 alpha-helix bundle [Wilkens, S., Dunn, S. D., Chandler, J., Dahlquist, F. W., and Capaldi, R. A. (1997) Nat. Struct. Biol. 4, 198-201] and that the majority of the binding energy between delta and F(1) is provided by the interaction between the N-terminal 22 residues of the alpha- and N-terminal domain of the delta subunit [Weber, J., Muharemagic, A., Wilke-Mounts, S., and Senior, A. E. (2003) J. Biol. Chem. 278, 13623-13626]. We have now analyzed a 1:1 complex of the delta-subunit N-terminal domain and a peptide comprising the N-terminal 22 residues of the alpha subunit by heteronuclear protein NMR spectroscopy. A comparison of the chemical-shift values of delta-subunit residues with and without alpha N-terminal peptide bound indicates that the binding interface on the N-terminal domain of the delta subunit is formed by alpha helices I and V. NOE cross-peak patterns in 2D (12)C/(12)C-filtered NOESY spectra of the (13)C-labeled delta-subunit N-terminal domain in complex with unlabeled peptide verify that residues 8-18 in the alpha-subunit N-terminal peptide are folded as an alpha helix when bound to delta N-terminal domain. On the basis of intermolecular contacts observed in (12)C/(13)C-filtered NOESY experiments, we describe structural details of the interaction of the delta-subunit N-terminal domain with the alpha-subunit N-terminal alpha helix.  相似文献   

4.
Sharp KH  Moody PC  Brown KA  Raven EL 《Biochemistry》2004,43(27):8644-8651
  相似文献   

5.
Autism encompasses a wide spectrum of disorders arising during brain development. Recent studies reported that sequence polymorphisms in neuroligin-3 (NLGN3) and neuroligin-4 (NLGN4) genes have been linked to autism spectrum disorders indicating neuroligin genes as candidate targets in brain disorders. We have characterized a single mutation found in two affected brothers that substituted Arg451 to Cys in NL3. Our data show that the exposed Cys causes retention of the protein in the endoplasmic reticulum (ER) when expressed in HEK-293 cells. To examine whether the introduction of a Cys in the C-terminal region of other alpha/beta-hydrolase fold proteins could promote the same cellular phenotype, we made homologous mutations in acetylcholinesterase (AChE) and butyrylcholinesterase (BChE) and found a similar processing deficiency and intracellular retention (De Jaco et al., J Biol Chem. 2006, 281:9667-76). NL3, AChE and BChE mutant proteins are recognized as misfolded in the ER, and degraded via the proteasome pathway. A 2D electrophoresis coupled with mass spectrometry based approach was used to analyze proteins co-immunoprecipitating with NL3 and show differential expression of factors interacting with wild type and mutant NL3. We identified several proteins belonging to distinct ER resident chaperones families, including calnexin, responsible for playing a role in the folding steps of the AChE and NLs.  相似文献   

6.
The (beta alpha)(8)-barrel is a versatile single-domain protein fold that is adopted by a large number of enzymes. The (beta alpha)(8)-barrel fold has been used as a model to elucidate the structural basis of protein thermostability and in studies to interconvert catalytic activities or substrate specificities by rational design or directed evolution. Recently, the (beta alpha)(4)-half-barrel was identified as a possible structural subdomain.  相似文献   

7.
Sun L  Warncke K 《Proteins》2006,64(2):308-319
The structure of the EutB protein from Salmonella typhimurium, which contains the active site of the coenzyme B12 (adenosylcobalamin)-dependent enzyme, ethanolamine ammonia-lyase, has been predicted by using structural proteomics techniques of comparative modelling. The 453-residue EutB protein displays no significant sequence identity with proteins of known structure. Therefore, secondary structure prediction and fold recognition algorithms were used to identify templates. Multiple three-dimensional template matching (threading) servers identified predominantly beta8alpha8, TIM-barrel proteins, and in particular, the large subunits of diol dehydratase (PDB: 1eex:A, 1dio:A) and glycerol dehydratase (PDB: 1mmf:A), as templates. Consistent with this identification, the dehydratases are, like ethanolamine ammonia-lyase, Class II coenzyme B12-dependent enzymes. Model building was performed by using MODELLER. Models were evaluated by using different programs, including PROCHECK and VERIFY3D. The results identify a beta8alpha8, TIM-barrel fold for EutB. The beta8alpha8, TIM-barrel fold is consistent with a central role of the alpha/beta-barrel structures in radical catalysis conducted by the coenzyme B12- and S-adenosylmethionine-dependent (radical SAM) enzyme superfamilies. The EutB model and multiple sequence alignment among ethanolamine ammonia-lyase, diol dehydratase, and glycerol dehydratase from different species reveal the following protein structural features: (1) a "cap" loop segment that closes the N-terminal region of the barrel, (2) a common cobalamin cofactor binding topography at the C-terminal region of the barrel, and (3) a beta-barrel-internal guanidinium group from EutB R160 that overlaps the position of the active-site potassium ion found in the dehydratases. R160 is proposed to have a role in substrate binding and radical catalysis.  相似文献   

8.
Sialic acid is an essential sugar in biology that plays key roles in numerous cellular processes and interactions. The biosynthesis of sialylated glycoconjugates is catalyzed by five distinct families of sialyltransferases. In the last 25 years, there has been much research on the enzymes themselves, their genes, and their reaction products, but we still do not know the precise molecular mechanism of action for this class of glycosyltransferase. We previously reported the first detailed structural and kinetic characterization of Cst-II, a bifunctional sialyltransferase (CAZy GT-42) from the bacterium Campylobacter jejuni [Chiu et al. (2004) Nat. Struct. Mol. Biol. 11, 163-170]. This enzyme can use both Gal-beta-1,3/4-R and Neu5Ac-alpha-2,3-Gal-beta-1,3/4-R as acceptor sugars. A second sialyltransferase from this bacterium, Cst-I, has been shown to utilize solely Gal-beta-1,3/4-R as the acceptor sugar in its transferase reaction. We report here the structural and kinetic characterization of this monofunctional enzyme, which belongs to the same sialyltransferase family as Cst-II, in both apo and substrate bound form. Our structural data show that Cst-I adopts a similar GTA-type glycosyltransferase fold to that of the bifunctional Cst-II, with conservation of several key noncharged catalytic residues. Significant differences are found, however, between the two enzymes in the lid domain region, which is critical to the creation of the acceptor sugar binding site. Furthermore, molecular modeling of various acceptor sugars within the active sites of these enzymes provides significant new insights into the structural basis for substrate specificities within this biologically important enzyme class.  相似文献   

9.
R67 dihydrofolate reductase (DHFR) shares no sequence or structural homology with chromosomal DHFRs. This enzyme arose recently in response to the clinical use of the antibacterial drug trimethoprim. R67 DHFR is a homotetramer possessing a single active site pore. A high-resolution crystal structure shows the homotetramer possesses exact 222 symmetry [Narayana, N., et al. (1995) Nat. Struct. Biol. 2, 1018-1025]. This symmetry dictates four symmetry-related binding sites must exist for each substrate as well as each cofactor. Isothermal titration calorimetry studies, however, indicate only two molecules bind: either two dihydrofolate molecules, two NADPH molecules, or one substrate and one cofactor [Bradrick, T. D., et al. (1996) Biochemistry 35, 11414-11424]. The latter is the productive ternary complex. To evaluate the role of S65, Q67, I68, and Y69 residues, located near the center of the active site pore, site-directed mutagenesis was performed. One mutation in the gene creates four mutations per active site pore which typically result in large cumulative effects. Steady state kinetic data indicate the mutants have altered K(m) values for both cofactor and substrate. For example, the Y69F R67 DHFR displays an 8-fold increase in the K(m) for dihydrofolate and a 20-fold increase in the K(m) for NADPH. Residues involved in ligand binding in R67 DHFR display very little, if any, specificity, consistent with their possessing dual roles in binding. These results support a model where R67 DHFR utilizes an unusual "hot spot" binding surface capable of binding both ligands and indicate this enzyme has adopted a novel yet simple approach to catalysis.  相似文献   

10.
The microtubule organizing center of the animal cell (S. D. Fulleret al.,1992,Curr. Opin. Struct. Biol.2,264–274; D. M. Gloveret al.,1993,Sci. Am.268,62–68; E. B. Wilson, 1925), (The Cell in Development and Heredity) comprises two centrioles and the pericentriolar material. We have completed several three-dimensional reconstructions of individual centrioles from tilt series of cryoelectron micrographs. The reconstruction procedure uses minimization of the common lines residual to define the orientation of the centriolar ninefold symmetry axis and then uses this symmetry to generate a structure by weighted backprojection to 28-nm resolution. Many of the features of these reconstructions agree with previous, conventional transmission electron microscopy studies (M. Paintrandet al.,1992,J. Struct. Biol.108,107–128). The microtubule barrel of the centriole is roughly 500 nm long and 300 nm in diameter and the microtubule bundles appear to taper toward the distal end. In addition, we see a handedness to the pericentriolar material at the base (distal end) of the centriole which is opposite to the skew of the microtubule triplets. The region at which the microtubule barrel joins this base is intriguingly complex and includes an internal cylindrical feature which is a site of γ tubulin localization.  相似文献   

11.
The structure of the calcium-saturated regulatory domain of skeletal troponin C (sNTnC) complexed with the switch peptide comprising residues 115-131 of troponin I (TnI), and with a bifunctional rhodamine fluorescent label attached to residues 56 (E56C) and 63 (E63C) on the C helix of sNTnC, has been determined using nuclear magnetic resonance (NMR) spectroscopy. The structure shows that the integrity of the C helix is not altered by the E(56,63)C mutations or by the presence of the bifunctional rhodamine and that the label does not interact with the hydrophobic cleft of sNTnC. Moreover, the overall fold of the protein and the position of the TnI peptide are similar to those observed previously with related cardiac NTnC complexes with residues 147-163 of cardiac TnI [Li et al. (1999) Biochemistry 38, 8289-8298] and including the drug bepridil [Wang et al. (2002) J. Biol. Chem. 277, 31124-31133]. The degree of opening of the structure is reduced as compared to that of calcium-saturated sNTnC in the absence of the switch peptide [Gagné et al. (1995) Nat. Struct. Biol. 2, 784-789]. The switch peptide is bound in a shallow and complementary hydrophobic surface cleft largely defined by helices A and B and also has key ionic interactions with sNTnC. These results show that bifunctional rhodamine probes can be attached to surface helices via suitable pairs of solvent-accessible residues that have been mutated to cysteines, without altering the conformation of the labeled domain. A set of such probes can be used to determine the orientation and motion of the target domain in the cellular environment [Corrie et al. (1999) Nature 400, 425-430; Ferguson et al. (2003) Mol. Cell 11(4), in press].  相似文献   

12.
Recently 3D structural models of the photosystem II (PSII) core dimer complexes of higher plants (spinach) and cyanobacteria (Synechococcus elongatus) have been derived by electron [Rhee et al. (1998) Nature 396, 283-286; Hankamer et al. (2001) J. Struct. Biol., in press] and X-ray [Zouni et al. (2001) Nature 409, 739-743] crystallography respectively. The intermediate resolutions of these structures do not allow direct identification of side chains and therefore many of the individual subunits within the structure are unassigned. Here we review the structure of the higher plant PSII core dimer and provide evidence for the tentative assignment of the low molecular weight subunits. In so doing we highlight the similarities and differences between the higher plant and cyanobacterial structures.  相似文献   

13.
N Lee  H Suga 《Biochemistry》2001,40(45):13633-13643
Numerous studies on naturally occurring ribozymes have shown that the functional roles of metal ions in promoting RNA catalysis are diverse. Earlier studies performed on the in vitro selected aminoacyl-transferase ribozyme (ATRib) have revealed that a fully hydrated Mg2+ ion plays an essential role in catalysis [Suga, H., Cowan, J. A., and Szostak, J. W. (1998) Biochemistry 28, 10118-10125]. More recently, we have evolved this ATRib into a bifunctional ribozyme, called AD02 [Lee, N., et al. (2000) Nat. Struct. Biol. 7, 28-33]. This new ribozyme consists of two catalytic domains, the original ATRib domain and a new glutamine-recognition (QR) domain, and exhibits a function of charging glutamine to tRNA. Here we elucidate crucial roles of metal ions involved in the QR domain, that are distinct from those in the ATRib domain. The metal ions in the QR domain require innersphere coordinations, and both Mg2+ and Ca2+ can support catalysis. Extensive Tb3+-Mg2+ and Tb3+-Co(NH3)6(3+) competition cleavage experiments have shown that the QR domain has high and low affinity metal binding sites, which are involved in the Mg2+-dependent structural alteration to form the glutamine binding site [Lee, N., and Suga, H. (2001) RNA 7, 1043-1051]. Kinetic studies in the presence of divalent and monovalent ions have suggested that the essential role of the metal ions in the QR domain is most likely structural.  相似文献   

14.
The (beta/alpha)(8)-barrel is the most common fold in structurally characterized enzymes. Whether the functionally diverse enzymes that share this fold are the products of either divergent or convergent evolution (or both) is an unresolved question that will probably be answered as the sequence databases continue to expand. Recent work has examined natural, designed, and directed evolution of function in several superfamilies of (beta/alpha)(8)-barrel containing enzymes.  相似文献   

15.
Previously we demonstrated by random saturation mutagenesis a set of mutations in the extracellular (EC) loops that constitutively activate the C5a receptor (C5aR) (Klco et al., Nat Struct Mol Biol 2005;12:320-326; Klco et al., J Biol Chem 2006;281:12010-12019). In this study, molecular modeling revealed possible conformations for the extracellular loops of the C5a receptors with mutations in the EC2 loop or in the EC3 loop. Comparison of low-energy conformations of the EC loops defined two distinct clusters of conformations typical either for strongly constitutively active mutants of C5aR (the CAM cluster) or for nonconstitutively active mutants (the non-CAM cluster). In the CAM cluster, the EC3 loop was turned towards the transmembrane (TM) helical bundle and more closely interacted with EC2 than in the non-CAM cluster. This suggested a structural mechanism of constitutive activity where EC3 contacts EC2 leading to EC2 interactions with helix TM3, thus triggering movement of TM7 towards TM2 and TM3. The movement initiates rearrangement of the system of hydrogen bonds between TM2, TM3 and TM7 including formation of the hydrogen bond between the side chains of D82(2.50) in TM2 and N296(7.49) in TM7, which is crucial for formation of the activated states of the C5a receptors (Nikiforovich et al., Proteins: Struct Funct Gene 2011;79:787-802). Since the relative large length of EC3 in C5aR (13 residues) is comparable with those in many other members of rhodopsin family of GPCRs (13-19 residues), our findings might reflect general mechanisms of receptor constitutive activation. The very recent X-ray structure of the agonist-induced constitutively active mutant of rhodopsin (Standfuss et al., Nature 2011;471:656-660) is discussed in view of our modeling results.  相似文献   

16.
Human cathepsin K, matrix metalloproteinase 9, and alpha(V)beta(3) integrin are the key regulators in osteoclast-mediated bone resorption. In this paper, we found natural inhibitors 1-10 for them by enzyme inhibition assays. Inhibitors 1-7, 8-9, and 10 are novel inhibitors of human cathepsin K, matrix metalloproteinase 9, and alpha(V)beta(3), respectively.  相似文献   

17.
Mutations in the notch ligand delta-like 3 have been identified in both the pudgy mouse (Dll3(pu); Kusumi et al.: Nat Genet 19:274-278, 1998) and the human disorder spondylocostal dysostosis (SCD; Bulman et al.: Nat Genet 24:438-441, 2000), and a targeted mutation has been generated (Dll3(neo); Dunwoodie et al.: Development 129:1795-1806, 2002). Vertebral and rib malformations deriving from defects in somitic patterning are key features of these disorders. In the mouse, notch pathway genes such as Lfng, Hes1, Hes7, and Hey2 display dynamic patterns of expression in paraxial mesoderm, cycling in synchrony with somite formation (Aulehla and Johnson: Dev Biol 207:49-61, 1999; Forsberg et al.: Curr Biol 8:1027-1030, 1998; Jouve et al.: Development 127:1421-1429, 2000; McGrew et al.: Curr Biol 8:979-982, 1998; Nakagawa et al.: Dev Biol 216:72-84, 1999). We report here that the Dll3(pu) mutation has different effects on the expression of cycling (Lfng and Hes7) and stage-specific genes (Hey3 and Mesp2). This suggests a more complex situation than a single oscillatory mechanism in somitogenesis and provides an explanation for the unique radiological features of the human DLL3-type of SCD.  相似文献   

18.
Gag, the major structural protein of retroviruses such as HIV-1, comprises a series of domains connected by flexible linkers. These domains drive viral assembly by mediating multiple interactions between adjacent Gag molecules and by binding to viral genomic RNA and host cell membranes. Upon viral budding, Gag is processed by the viral protease to liberate distinct domains as separate proteins. The first two regions of Gag are MA, a membrane-binding module, and CA, which is a two-domain protein that makes important Gag-Gag interactions, forms the cone-shaped outer shell of the core (the capsid) in the mature HIV-1 particle, and makes an important interaction with the cellular protein cyclophilin A (CypA). Here, we report crystal structures of the mature CA N-terminal domain (CA(N)(133-278)) and a MA-CA(N) fusion (Gag(1-278)) at resolutions/R(free) values of 1.9 A/25.7% and 2.2 A/25.8%, respectively. Consistent with earlier studies, a comparison of these structures indicates that processing at the MA-CA junction causes CA to adopt an N-terminal beta-hairpin conformation that seems to be required for capsid morphology and viral infectivity. In contrast with an NMR study (Tang, C., et al. (2002) Nat. Struct. Biol. 9, 537-543), structural overlap reveals only small relative displacements for helix 6, which is located between the beta-hairpin and the CypA-binding loop. These observations argue against the proposal that CypA binding is coupled with beta-hairpin formation and support an earlier surface plasmon resonance study (Yoo, S., et al. (1997) J. Mol. Biol. 269, 780-795), which concluded that beta-hairpin formation and CypA-binding are energetically independent events.  相似文献   

19.
The question of parallel (alpha/beta)8-barrel fold evolution remains unclear, owing mainly to the lack of sequence homology throughout the amino acid sequences of (alpha/beta)8-barrel enzymes. The "classical" approaches used in the search for homologies among (alpha/beta)8-barrels (e.g., production of structurally based alignments) have yielded alignments perfect from the structural point of view, but the approaches have been unable to reveal the homologies. These are proposed to be "hidden" in (alpha/beta)8-barrel enzymes. The term "hidden homology" means that the alignment of sequence stretches proposed to be homologous need not be structurally fully satisfactory. This is due to the very long evolutionary history of all (alpha/beta)8-barrels. This work identifies so-called hidden homology around the strand beta 2 that is flanked by loops containing invariant glycines and prolines in 17 different (alpha/beta)8-barrel enzymes, i.e., roughly in half of all currently known (alpha/beta)8-barrel proteins. The search was based on the idea that a conserved sequence region of an (alpha/beta)8-barrel enzyme should be more or less conserved also in the equivalent part of the structure of the other enzymes with this folding motif, given their mutual evolutionary relatedness. For this purpose, the sequence region around the well-conserved second beta-strand of alpha-amylase flanked by the invariant glycine and proline (56_GFTAIWITP, Aspergillus oryzae alpha-amylase numbering), was used as the sequence-structural template. The proposal that the second beta-strand of (alpha/beta)8-barrel fold is important from the evolutionary point of view is strongly supported by the increasing trend of the observed beta 2-strand structural similarity for the pairs of (alpha/beta)8-barrel enzymes: alpha-amylase and the alpha-subunit of tryptophan synthase, alpha-amylase and mandelate racemase, and alpha-amylase and cyclodextrin glycosyltransferase. This trend is also in agreement with the existing evolutionary division of the entire family of (alpha/beta)8-barrel proteins.  相似文献   

20.
Enzymes with the (beta/alpha)(8)-barrel fold are involved in the catalysis of a wide variety of biochemical reactions. The active sites of these enzymes are located on the C-terminal face of the central beta-barrel. Conserved amino acid sequence, as well as secondary, tertiary and quaternary structure patterns are providing a rich body of data to support the premise of a common ancestry of many members of the (beta/alpha)(8)-barrel fold family of enzymes. Recent data indicate that there is at least one example of a bienzyme that functions as an ammonia channel, adding a new level of functional diversity to the (beta/alpha)(8)-barrel fold. These proteins have become ideal tools that can be used in conjunction with directed evolution techniques to engineer novel catalytic activities.  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