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 values of four conformational properties, unperturbed dimensions 0, dipole moment , mean squared optical anisotropy and molar Kerr constant have been calculated for polyglycine chains of x = 100 repeat units with varying composition of alpha-helix, beta-sheet and random coil conformations. The influence of the conformational sequence on these properties has been investigated by calculating the four above-mentioned properties together with the end-to-end vector for several polyglycine oligomers.  相似文献   

2.
Gray DM 《Biopolymers》1997,42(7):795-810
Using nearest-neighbor models consisting of independent short sequence combinations of nearest neighbors (ISS models), values of thermodynamic parameters for sets of independent sequences are derived from published oligomer data for DNA.RNA hybrids [N. Sugimoto, S. Nakano, M. Katoh, A. Matsumura, H. Nakamuta, T. Ohmichi, M. Yoneyama, and M. Sasaki (1995) Biochemistry, Vol. 34, pp. 11211-11216] and dsDNA duplexes [J. SantaLucia, Jr., H. T. Allawi, and P. A. Seneviratne (1996) Biochemistry, Vol. 35, pp. 3555-3562]. The results are compared with those from models that assign values of thermodynamic parameters to individual nearest neighbors (INN models). Differences in the use of ISS and INN models are also illustrated in an appendix, which shows examples of analyses for values of a fictitious nearest-neighbor property. INN models that include an initiation parameter contain an implicit assumption that combinations of end neighbors have the same value of a property. It is found that combinations of end neighbors (e.g., base pairs neighboring solvent) in oligomers can have significant and different apparent values of thermodynamic properties, so that the assumption inherent in INN models is not always correct. Even though ISS models do not allow the assignment of values to individual nearest neighbors, except for the like neighbors [such as d(AA)/r(UU), etc., for hybrids and d(AA)/d(TT) and d(GG)/d(CC) for DNA duplexes], they do provide physically meaningful values for the like neighbors, for sequence combinations, and for specified combinations of end neighbors.  相似文献   

3.
Coagulation factor V (FV) circulates in two forms, FV1 and FV2, having slightly different molecular masses and phospholipid-binding properties. The aim was to determine whether this heterogeneity is due to the degree of glycosylation of Asn(2181). FVa1 and FVa2 were isolated and digested with endoglycosidase PNGase F. As judged by Western blotting, the FVa2 light chain contained two N-linked carbohydrates, whereas FVa1 contained three. Wild-type FV and three mutants, Asn(2181)Gln, Ser(2183)Thr, and Ser(2183)Ala, were expressed in COS1 cells, activated by thrombin, and analyzed by Western blotting. Wild-type FVa contained the 71 kDa-74 kDa doublet, whereas the Asn(2181)Gln and Ser(2183)Ala mutants contained only the 71 kDa light chain. In contrast, the Ser(2183)Thr mutant gave a 74 kDa light chain. This demonstrated that the third position in the Asn-X-Ser/Thr consensus affects glycosylation efficiency, Thr being associated with a higher degree of glycosylation than Ser. The Ser(2183)Thr mutant FVa was functionally indistinguishable from plasma-purified FVa1, whereas Asn(2181)Gln and Ser(2183)Ala mutants behaved like FVa2. Thus, the carbohydrate at Asn(2181) impaired the interaction between FVa and the phospholipid membrane, an interpretation consistent with a structural analysis of a three-dimensional model of the C2 domain and the position of a proposed phospholipid-binding site. In conclusion, we show that the FV1-FV2 heterogeneity is caused by differential glycosylation of Asn(2181) related to the presence of a Ser rather than a Thr at the third position in the consensus sequence of glycosylation.  相似文献   

4.
In fish retina, four kinds of photoreceptor cells (or cones) are two-dimensionally arranged in a very regular manner, forming cone mosaics. Mosaic pattern differs between species--two typical patterns are "row mosaic" and "square mosaic", exemplified by the cone mosaics in zebrafish and in medaka, respectively. In this paper, we study a cell-rearrangement model. Cells with pre-fixed fate exchange their locations between nearest neighbors and form regular mosaic patterns spontaneously, if the adhesive force between nearest neighbors and between next-nearest neighbors depend on their cell types in an appropriate manner. The same model can produce both row and square mosaic patterns. However, if the cell-cell interaction is restricted to nearest neighbors only, the square mosaic (medaka pattern) cannot be generated, showing the importance of interaction between next-nearest neighbors. In determining whether row mosaic (zebrafish pattern) or square mosaic (medaka pattern) is to be formed, two shape factors are very important, which control the way adhesions in different geometric relations are combined. We also developed theoretical analysis of the parameter ranges for the row mosaic and the square mosaic to have higher total adhesion than alternative spatial patterns.  相似文献   

5.
Gray DM 《Biopolymers》1997,42(7):783-793
The constraints on combinations of nearest neighbors in nucleic acid sequences and the numbers of independent sequences needed to describe nearest-neighbor properties of oligomers and polymers are derived and summarized. It has been pointed out in previous work [D. M. Gray and I. Tinoco, Jr. (1970) Biopolymers, Vol. 9, pp. 223-244; R. F. Goldstein and A. S. Benight (1992) Biopolymers, Vol. 32, pp. 1679-1693] that these constraints restrict the information available from measurements of properties of sequence combinations. The emphasis in this paper is on the properties of oligomer sequences that vary in length, where each nucleotide or base pair at the end of the sequence makes a significant contribution to the measured property by interacting with its boundary of fixed sequence or solvent. In such cases it is not be possible to determine values of properties of individual nearest neighbors, except for the like neighbors [e.g., d(A-A), d(G-G), d(T-T), and d(C-C) nucleotide neighbors in single-stranded DNA or d(A-A)/d(T-T) and d(G-G)/d(C-C) base pair neighbors in double-stranded DNA], solely from measurements of properties of different sequences. Even values for properties of the like neighbors cannot be determined from such oligomeric sequences if the sequences are all of the same length. Nearest-neighbor properties of oligomer sequences that vary in length can be summarized in terms of the values for independent sets of sequences that are nearest neighbors and monomers all with boundaries of the fixed sequence or solvent. Straightforward combinations of the values for the independent sequences will give the values of the property for any dependent sequence, without explicit knowledge of the individual nearest-neighbor values. These considerations have important consequences for the derivation of widely used thermodynamic parameters, as discussed in the following paper.  相似文献   

6.
Morton S  Davis RJ  McLaren A  Cohen P 《The EMBO journal》2003,22(15):3876-3886
We have used phospho-specific antibodies to re-examine the multisite phosphorylation of c-Jun in murine RAW macrophages and embryonic fibroblasts. Our results indicate that JNK isoforms are required and sufficient for the phosphorylation of Thr91 and Thr93, as well as the phosphorylation of Ser63 and Ser73, in response to LPS or anisomycin in macrophages and TNFalpha or anisomycin in fibroblasts. However, the phorbol ester (TPA) and EGF-induced phosphorylation of Ser63 and Ser73 is mediated by ERK1/ERK2, as well as JNK1/JNK2, in fibroblasts from wild-type mice and by ERK1/ERK2 alone in fibroblasts from JNK-deficient mice. The phosphorylation of Thr239 is catalysed by GSK3 and the phosphorylation of Ser243 by an as yet unidentified protein kinase. The inhibition of GSK3 is not required for the dephosphorylation of Thr239 in response to LPS, and nor is the phosphorylation of Thr91 and Thr93 required for the TPA- or EGF-induced dephosphorylation of Thr239 in fibroblasts. The agonist-induced dephosphorylation of Thr239 may involve a conformational change that exposes Thr239 to dephosphorylation and/or the activation of a Thr239 phosphatase.  相似文献   

7.
Phosphorylation of phospholamban (PLB) at Ser16 and/ or Thr17 is believed to release its inhibitory effect on sarcoplasmic reticulum calcium ATPase. Ser16 phosphorylation of PLB has been suggested to cause a conformational change that alters the interaction between the enzyme and protein. Using computer simulations, the conformational sampling of Ser16 phosphorylated PLB in implicit membrane environment is compared here with the unphosphorylated PLB system to investigate these conformational changes. The results suggest that conformational changes in the cytoplasmic domain of PLB upon phosphorylation at Ser16 increase the likelihood of unfavorable interactions with SERCA in the E2 state prompting a conformational switch of SERCA from E2 to E1. Phosphorylation of PLB at Thr17 on the other hand does not appear to affect interactions with SERCA significantly suggesting that the mechanism of releasing the inhibitory effect is different between Thr17 phosphorylated and Ser16 phosphorylated PLB.  相似文献   

8.
Phosphorylation sites of protamines by protein kinase C and cAMP-dependent protein kinase (protein kinase A) were studied. Using clupeine Y1 as a substrate, protein kinase C phosphorylates both Ser and Thr residues, whereas protein kinase A phosphorylates only Ser residue(s). Protein kinase C phosphorylates all Ser and Thr residues of clupeine Y2 and Z, however protein kinase A phosphorylates mainly Ser9 and slightly Thr5 in clupeine Y2 and Ser6 and Ser10 in clupeine Z. These results suggest that protein kinase C recognizes more sites than those of protein kinase A and may participate in protamine phosphorylation in vivo.  相似文献   

9.
The principal signal to activate smooth muscle contraction is phosphorylation of the regulatory light chains of myosin (LC(20)) at Ser(19) by Ca(2+)/calmodulin-dependent myosin light chain kinase. Inhibition of myosin light chain phosphatase leads to Ca(2+)-independent phosphorylation at both Ser(19) and Thr(18) by integrin-linked kinase and/or zipper-interacting protein kinase. The functional effects of phosphorylation at Thr(18) on steady-state isometric force and relaxation rate were investigated in Triton-skinned rat caudal arterial smooth muscle strips. Sequential phosphorylation at Ser(19) and Thr(18) was achieved by treatment with adenosine 5'-O-(3-thiotriphosphate) in the presence of Ca(2+), which induced stoichiometric thiophosphorylation at Ser(19), followed by microcystin (phosphatase inhibitor) in the absence of Ca(2+), which induced phosphorylation at Thr(18). Phosphorylation at Thr(18) had no effect on steady-state force induced by Ser(19) thiophosphorylation. However, phosphorylation of Ser(19) or both Ser(19) and Thr(18) to comparable stoichiometries (0.5 mol of P(i)/mol of LC(20)) and similar levels of isometric force revealed differences in the rates of dephosphorylation and relaxation following removal of the stimulus: t(½) values for dephosphorylation were 83.3 and 560 s, and for relaxation were 560 and 1293 s, for monophosphorylated (Ser(19)) and diphosphorylated LC(20), respectively. We conclude that phosphorylation at Thr(18) decreases the rates of LC(20) dephosphorylation and smooth muscle relaxation compared with LC(20) phosphorylated exclusively at Ser(19). These effects of LC(20) diphosphorylation, combined with increased Ser(19) phosphorylation (Ca(2+)-independent), may underlie the hypercontractility that is observed in response to certain physiological contractile stimuli, and under pathological conditions such as cerebral and coronary arterial vasospasm, intimal hyperplasia, and hypertension.  相似文献   

10.
MUC1 is a membrane glycoprotein, which in adenocarninomas is overexpressed and exhibits truncated O‐glycosylation. Overexpression and altered glycosylation make MUC1 into a candidate for immunotherapy. Monoclonal antibodies directed against MUC1 frequently bind an immunodominant epitope that contains a single site for O‐glycosylation. Glycosylation with tumor carbohydrate antigens such as the Tn‐antigen (GalNAc‐O‐Ser/Thr) results in antibodies binding with higher affinity. One proposed model to explain the enhanced affinity of antibodies for the glycosylated antigen is that the addition of a carbohydrate alters the conformational properties, favoring a binding‐competent state. The conformational effects associated with Tn glycosylation of the MUC1 antigen was investigated using solution‐state NMR and molecular dynamics. NMR experiments revealed distinct substructures of the glycosylated MUC1 peptides compared with the unglycosylated peptide. Molecular dynamics simulations of the MUC1 glycopeptide and peptide revealed distinguishing differences in their conformational preferences. Furthermore, the glycopeptide displayed a smaller conformational sampling compared with the peptide, suggesting that the glycopeptide sampled a narrower conformational space and is less dynamic. A comparison of the computed ensemble of conformations assuming random distribution, NMR models, and molecular dynamics simulations indicated that the MUC1 glycopeptide and aglycosylated peptide sampled structurally distinctly ensembles and that these ensembles were different from that of the random coil. Together, these data support the hypothesis that that conformational pre‐selection could be an essential feature of these peptides that dictates the binding affinities to MUC1 specific antibodies.  相似文献   

11.
The influence of peptide sequence and environment on the initiation and elongation of mucin O-glycosylation is not well understood. The in vivo glycosylation pattern of the porcine submaxillary gland mucin (PSM) tandem repeat containing 31 O-glycosylation sites (Gerken, T. A., Gilmore, M., and Zhang, J. (2002) J. Biol. Chem. 277, 7736-7751) reveals a weak inverse correlation with hydroxyamino acid density (and by inference the density of glycosylation) with the extent of GalNAc glycosylation and core-1 substitution. We now report the time course of the in vitro glycosylation of the apoPSM tandem repeat by recombinant UDP-GalNAc:polypeptide alpha-GalNAc transferases (ppGalNAc transferase) T1 and T2 that confirm these findings. A wide range of glycosylation rates are found, with several residues showing apparent plateaus in glycosylation. An adjustable kinetic model that reduces the first-order rate constants proportional to neighboring glycosylation status, plus or minus three residues of the site of glycosylation, was found to reasonably reproduce the experimental rate data for both transferases, including apparent plateaus in glycosylation. The unique, transferase-specific, positional weighting constants reveal information on the peptide/glycopeptide recognition site for each transferase. Both transferases displayed high sensitivities to neighboring Ser/Thr glycosylation, whereas ppGalNAc T2 displayed additional high sensitivities to the presence of nonglycosylated Ser/Thr residues. This is the first demonstration of the ability to model mucin O-glycosylation kinetics, confirming that under the appropriate conditions neighboring glycosylation status can be a significant factor modulating the first step of mucin O-glycan biosynthesis.  相似文献   

12.
Three 2-acetamido-2-deoxy-alpha-D-galactopyranoses attached to Ser2, Thr3 and Thr4 of the amino-terminal portion of glycophorin AM are responsible for the so-called TN blood group specificity. The corresponding glycopeptide H2N-Ser-Ser*-Thr*-Gly-OH obtained by a stepwise peptide coupling strategy was submitted to a detailed high-field nuclear magnetic resonance (n.m.r.) analysis. 13C-n.m.r. spectrum confirms the validity of previous assignments made on M sialo and asialoglycopeptides obtained by specific degradation of human glycophorin AM. In addition, the 400 MHz 1H-n.m.r. spectrum allowed most of the proton resonances to be assigned. A careful examination of the chemical shifts and coupling constants revealed some interesting features of the conformational properties of the GalNAc-Ser and GalNAc-Thr linkage as well as of the rotational isomerism of Thr and Ser side-chains. The data give conclusive evidence that high-field n.m.r. spectroscopy can be successfully used to gain structural and dynamic information on rather sophisticated glycopeptides.  相似文献   

13.
Here we report the use of automated Edman degradation of covalently linked glycopeptides to identify positively the sites of O- and N-glycosylation. The O-glycosidic linkage of carbohydrate to the hydroxy amino acids Ser and Thr is a major form of post-translational modification. However, unlike Asn-linked glycosylation, which is identified by the consensus sequence Asn-Xaa-Thr/Ser, no simple motif conferring O-linkage to Thr and Ser has been described. After sequencing glycopeptides derived from two cell surface glycoproteins, a Thr-O-glycosylation motif of Xaa-Pro-Xaa-Xaa, where at least one Xaa = Thr(Sac), has been defined. This motif predicts the site(s) of Pro- associated Thr-O-glycosylation in O-glycosylated proteins, although it is clear that there are also other forms of Thr-O-glycosylation not associated with Pro.  相似文献   

14.
MAPKAP kinase 2 (MK2) is required for tumor necrosis factor synthesis. Tristetraprolin (TTP) binds to the 3'-untranslated region of tumor necrosis factor mRNA and regulates its fate. We identified in vitro and in vivo phosphorylation sites in TTP using nanoflow high pressure liquid chromatography microelectrospray ionization tandem mass spectrometry and novel methods for direct digestion of TTP bound to affinity matrices (GSH-beads or anti-Myc linked to magnetic beads). MK2Delta3B, activated in Escherichia coli by p38alpha, phosphorylates TTP in vitro at major sites Ser(52) and Ser(178) (>10-fold in abundance) as well as at several minor sites that were detected after enriching for phosphopeptides with immobilized metal affinity chromatography. MK2 phosphorylation of TTP creates a functional 14-3-3 binding site. In cells, TTP was phosphorylated at Ser(52), Ser(178), Thr(250), and Ser(316) and at SP sites in a cluster (Ser(80)/Ser(82)/Ser(85)). Anisomycin treatment of NIH 3T3 cells increased phosphorylation of Ser(52) and Ser(178). Overexpression of MK2 sufficed to increase phosphorylation of Ser(52) and Ser(178) but not Ser(80)/Ser(82)/Ser(85) or Thr(250). Thus, Ser(52) and Ser(178) are putative MK2 sites in vivo. Identified phosphosite(s) may be biologic switches controlling mRNA stability and translation.  相似文献   

15.
The relationship between the Ser, Thr, and Cys side-chain conformation (chi(1) = g(-), t, g(+)) and the main-chain conformation (phi and psi angles) has been studied in a selection of protein structures that contain alpha-helices. The statistical results show that the g(-) conformation of both Ser and Thr residues decreases their phi angles and increases their psi angles relative to Ala, used as a control. The additional hydrogen bond formed between the O(gamma) atom of Ser and Thr and the i-3 or i-4 peptide carbonyl oxygen induces or stabilizes a bending angle in the helix 3-4 degrees larger than for Ala. This is of particular significance for membrane proteins. Incorporation of this small bending angle in the transmembrane alpha-helix at one side of the cell membrane results in a significant displacement of the residues located at the other side of the membrane. We hypothesize that local alterations of the rotamer configurations of these Ser and Thr residues may result in significant conformational changes across transmembrane helices, and thus participate in the molecular mechanisms underlying transmembrane signaling. This finding has provided the structural basis to understand the experimentally observed influence of Ser residues on the conformational equilibrium between inactive and active states of the receptor, in the neurotransmitter subfamily of G protein-coupled receptors.  相似文献   

16.
The β-elimination/Michael addition reaction has been employed for the modification of O-acylated and phosphorylated Ser and Thr residues in a variety of derivatives. The modified Ser and Thr can be analyzed by amino acid composition analysis, N-terminal Edman degradation sequence analysis, and tandem mass spectrometric sequencing which generally allows the identification and localization of the phosphorylation or glycosylation sites. However, the reactivity of the free hydroxyl group on serine and threonine by sodium hydroxide-induced β-elimination has not been critically examined. In this study, two analogous phosphopeptides, KMpSTLSYR and KMSpTLSYR, were subjected to β-elimination under the widely used conditions previously reported, followed by sulfite or ethanethiol addition. After treatment of the phosphopeptides in 0.1 N NaOH/0.6 M Na2SO3 at 37 °C for 24 h, matrix-assisted laser desorption ionization-time of flight mass spectrometric analyses of the products revealed an appreciable mass peak with an additional observed mass of 64 compared to the expected mass from the conversion of phosphate to sulfite. Similarly, treatment of the phosphopeptides in 0.52 N NaOH/1.36 M ethanethiol at 50 °C for 18 h or for even as short as 1 h also yielded additional 44 mass of ethylthiogroup in excess of the expected mass for the modified phosphopeptide. Electrospray ionization tandem mass spectrometric analysis confirms that the modification occurred on the hydroxyl group of Ser and Thr in addition to P-Ser and P-Thr. On the other hand, modification on the free hydroxyl group of Ser or Thr was not detected under the mild condition of 0.1 N NaOH/0.6 M Na2SO3 at 25 °C for 24 h as previously reported (Li et al., Anal. Chem. 74:5701-5710, 2002). This finding suggests that temperatures above 25 °C and excessive alkalinity should be avoided to prevent the β-elimination of the hydroxyl group of Ser and Thr in peptides. This is of particular concern when employing highly sensitive tandem mass spectrometric methods for the identification and localization of Ser and Thr as modification sites by the β-elimination/Michael addition reaction. The additional modification site(s) may complicate the interpretation of data and lead to an erroneous conclusion.  相似文献   

17.
BackgroundGlycosylation of proteins is the most common, multifaceted co- and post-translational modification responsible for many biological processes and cellular functions. Significant alterations and aberrations of these processes are related to various pathological conditions, and often turn out to be disease biomarkers. Conventional N-glycosylation occurs through the recognition of the consensus sequon, asparagine (Asn)-X-serine (Ser)/threonine (Thr), where X is any amino acid except for proline, with N-acetylglucosamine (GlcNAc) as the first glycosidic linkage. Usually, O-glycosylation adds a glycan to the hydroxyl group of Ser or Thr beginning with N-acetylgalactosamine (GalNAc).Scope of reviewProtein glycosylation is further governed by additional diversifications in sequon and structure, which are yet to be fully explored. This review mainly focuses on the occurrence of N-glycosylation in non-consensus motifs, where Ser/Thr at the + 2 position is substituted by other amino acids. Additionally, N-glycosylation is also observed in other amide/amine group-containing amino acids. Similarly, O-glycosylation occurs at hydroxyl group-containing amino acids other than serine/threonine. The neighbouring amino acids and local structural features around the potential glycosylation site also play a significant role in determining the extent of glycosylation. All of these phenomena that yield glycosylation at the atypical sites are reported in a variety of biological systems, including different pathological conditions.Conclusion and SignificanceTherefore, the discovery of more novel sequence patterns for N- and O-glycosylation may help in understanding the functions of complex biological processes and cellular functions. Taken together, all these information provided in this review would be helpful for the biological readers.  相似文献   

18.
Carbon-13 NMR spectroscopic studies of native and sequentially deglycosylated ovine submaxillary mucin (OSM) have been performed to examine the effects of glycosylation on the conformation and dynamics of the peptide core of O-linked glycoproteins. OSM is a large nonglobular glycoprotein in which nearly one-third of the amino acid residues are Ser and Thr which are glycosylated by the alpha-Neu-NAc(2-6)alpha-GalNAc- disaccharide. The beta-carbon resonances of glycosylated Ser and Thr residues in intact and asialo mucin display considerable chemical shift heterogeneity which, upon the complete removal of carbohydrate, coalesces to single sharp resonances. This chemical shift heterogeneity is due to peptide sequence variability and is proposed to reflect the presence of sequence-dependent conformations of the peptide core. These different conformations are thought to be determined by steric interactions of the GalNAc residue with adjacent peptide residues. The absence of chemical shift heterogeneity in apo mucin is taken to indicate a loss in the peptide-carbohydrate steric interactions, consistent with a more relaxed random coiled structure. On the basis of the 13C relaxation behavior (T1 and NOE) the dynamics of the alpha-carbons appear to be unique to each amino acid type and glycosylation state, with alpha-carbon mobilities decreasing in the order Gly greater than Ala = Ser greater than Thr much greater than monoglycosylated Ser/Thr approximately greater than disaccharide linked Ser/Thr.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
Proline-directed protein phosphorylation was shown to depend on the capacity of the targeted Ser(Thr)-Pro bond to exhibit conformational polymorphism. The cis/trans isomer specificity underlying ERK2-catalyzed phosphate transfer leads to a complete discrimination of the cis Ser(Thr)-Pro conformer of oligopeptide substrates. We investigated in vitro the ERK2-catalyzed phosphorylation of Aspergillus oryzae RNase T1 containing two Ser-Pro bonds both of which share high stabilization energy in their respective native state conformation, the cis Ser54-Pro and the trans Ser72-Pro moiety. Despite trans isomer specificity of ERK2, a doubly phosphorylated RNase T1 was found as the final reaction product. Similarly, the RNase T1 S54G/P55N and RNase T1 P73V variants, which retain the prolyl bond conformations of the RNase T1-wt, were both monophosphorylated with a catalytic efficiency kcat/KM of 425 M(-1) s(-1) and 1228 M(-1) s(-1), respectively. However, initial phosphorylation rates did not depend linearly on the ERK2 concentration. The phosphorylation rate of the resulting plateau region at high ERK2 concentrations can be increased up to threefold for the RNase T1 P73V variant in the presence of the peptidyl-prolyl cis/trans isomerase Cyclophilin 18, indicating a conformational interconversion as the rate limiting step in the catalyzed phosphate group transfer. Using peptidyl-prolyl cis/trans isomerases with different substrate specificity, we identified a native state conformational equilibrium of the Ser54-Pro bond with the minor trans Ser54-Pro bond as the phosphorylation-sensitive moiety. This technique can therefore be used for a determination of the ratio and the interconversion rates of prolyl bond isomers in the native state of proteins.  相似文献   

20.
A three-dimensional model of the leucine zipper GCN4 built from its amino acid sequence had been reported previously by us. When the two alternative x-ray structures of the GCN4 dimer became available, the root mean square (r.m.s.) shifts between our model and the structures were determined as approximately 2.7 A on all atoms. These values are similar to the r.m.s. shift of 2.8 A between the two GCN4 structures in the different crystal forms (C2 and P2(1)2(1)2(1)). CONGEN conformational searches were run to better understand the conditions that may determine the preference of different conformers in different environments and to test the sensitivity of our current modeling techniques. With a judicious choice of CONGEN search parameters, the backbone r.m.s. deviation improved to 0.8 A and 2.5 A on all atoms. The side-chain conformations of Val and Leu at the helical interface were well reproduced (1.2 A r.m.s.), and the large side-chain misplacements occurred with only a small number of charged amino acids and a tyrosine. Inclusion of the crystal environment (C2 symmetry), as a passive background, into the side-chain conformational search further improved the accuracy of the model to an r.m.s. deviation of 2.1 A. Conformational searches carried out in the two different crystal environments and employing the AMBER protein/DNA forcefield, as implemented in CONGEN, gave the r.m.s. values of 2.2 A (for the C2 symmetry) and 2.5 A (for the P2(1)2(1)2(1) symmetry). In the C2 symmetry crystal, as much as 40% of the surface of each dimer was involved in crystal contacts with other dimers, and the charged residues on the surface often interacted with immobilized water molecules. Thus, occasional large r.m.s. deviations between the model and the x-ray side chains were due to specific conditions that did not occur in solution.  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