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.
Values of four conformational properties, namely unperturbed dimension [r2]0, dipole moment [mu 2], mean squared optical anisotropy [gamma 2], and molar Kerr constant [mK], have been calculated for polyglycine chains allowing several combinations of the secondary structure with the aim of studying the dependence of these magnitudes on the secondary structure of the chain. Two different approaches to the secondary structure have been used. In the first, chains with all their units in a given conformation (random coil, alpha-helix or beta-sheet) are interrupted at several positions by one unit in a different conformation. In the second, chains with varying composition of two conformations alpha-helix/beta-sheet and beta-sheet/random coil were allowed and the results obtained compared with previous work for alpha-helix/random coil chains.  相似文献   

2.
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   

3.
Spleen cells from mice primed with the thymus-dependent antigen trinitrophenyl keyhold limpet hemocyanin several months earlier can be cultured in vitro to give vigorous IgG antihapten PFC responses to thymus-dependent (TD) and thymus-independent (TI) forms of the same hapten. Here we show that the IgG memory precursors that respond to these two forms of the hapten constitute functionally distinct subpopulations. We have designated these subpopulations as B1gamma and B2gamma to represent secondary precursor cells responding to TI and TD antigens, respectively. Three types of evidence for these subpopulations are presented: 1) In vitro secondary IgG responses to TD and TI forms of the TNP hapten are additive when both forms are added to the same culture. 2) The precursor frequencies for the TD and TI antigens are additive, but addition is not observed between two TD or two TI antigens. 3) Each population can be selectively eliminated by BUdR and light treatment without affecting the other population. The ontogenetic relationships between these subpopulations are discussed in relation to all presently proposed subpopulations B1mu, B2mu, B1gamma, and B2gamma.  相似文献   

4.
The conformational stabilities of two homodimeric class mu glutathione transferases (GSTM1-1 and GSTM2-2) were studied by urea- and guanidinium chloride-induced denaturation. Unfolding is reversible and structural changes were followed with far-ultraviolet circular dichroism, tryptophan fluorescence, enzyme activity, chemical cross-linking, and size-exclusion chromatography. Disruption of secondary structure occurs as a monophasic transition and is independent of protein concentration. Changes in tertiary structure occur as two transitions; the first is protein concentration dependent, while the second is weakly dependent (GSTM1-1) or independent (GSTM2-2). The second transition corresponds with the secondary structure transition. Loss in catalytic activity occurs as two transitions for GSTM1-1 and as one transition for GSTM2-2. These transitions are dependent upon protein concentration. The first deactivation transition coincides with the first tertiary structure transition. Dimer dissociation occurs prior to disruption of secondary structure. The data suggest that the equilibrium unfolding/refolding of the class mu glutathione transferases M1-1 and M2-2 proceed via a three-state process: N(2) <--> 2I <--> 2U. Although GSTM1-1 and GSTM2-2 are homologous (78% identity/94% homology), their N(2) tertiary structures are not identical. Dissociation of the GSTM1-1 dimer to structured monomers (I) occurs at lower denaturant concentrations than for GSTM2-2. The monomeric intermediate for GSTM1-1 is, however, more stable than the intermediate for GSTM2-2. The intermediates are catalytically inactive and display nativelike secondary structure. Guanidinium chloride-induced denaturation yields monomeric intermediates, which have a more loosely packed tertiary structure displaying enhanced solvent exposure of its tryptophans and enhanced ANS binding. The three-state model for the class mu enzymes is in contrast to the equilibrium two-state models previously proposed for representatives of classes alpha/pi/Sj26 GSTs. Class mu subunits appear to be intrinsically more stable than those of the other GST classes.  相似文献   

5.
We have determined the complete nucleotides sequence (2168 bases) of the immunoglobulin mu gene cloned from newborn mouse DNA. The cloned 13kb fragment contained the entire constant region gene sequence that is interrupted by three intervening sequences at the junction of domains as previously shown in the gamma 1, gamma 2 b and alpha genes. The amino acid sequence predicted by the nucleotide sequence agrees with that of the mu chain secreted by a myeloma MOPC104E except for 8 residues out of 448 residues. The homologous domains of the mu, gamma 1 and gamma 2b genes are more similar to each other than the different domains of the mu genes are. The result implicates that the class of the immunoglobulin heavy chain genes diverged after the heavy chain genes established the multi-domain structure. The short intervening sequences of the mu and gamma genes are more conserved than the coding sequences except for the COOH-terminal domains. The results implicate that the nucleotide sequence of the intervening sequence is under selective pressure, possibly to maintain a secondary structure of the nuclear RNA to be spliced.  相似文献   

6.
M A Tsai  R E Waugh    P C Keng 《Biophysical journal》1996,70(4):2023-2029
In this study, the role of cytoskeleton in HL-60 deformability during the cell cycle was investigated. G1, S, and G2/M cell fractions were separated by centrifugal elutriation. Cell deformability was evaluated by pipette aspiration. Tested at the same aspiration pressures, S cells were found to be less deformable than G1 cells. Moreover, HL-60 cells exhibited power-law fluid behavior: mu = mu c(gamma m/ gamma c)-b, where mu is cytoplasmic viscosity, gamma m is mean shear rate, mu c is the characteristic viscosity at the characteristic shear rate gamma c, and b is a material constant. At a given shear rate, S cells (mu c = 276 +/- 14 Pa.s, b = 0.51 +/- 0.03) were more viscous than G1 cells (mu c = 197 +/- 25, b = 0.53 +/- 0.02). To evaluate the relative importance of different cytoskeletal components in these cell cycle-dependent properties, HL-60 cells were treated with 30 microM dihydrocytochalasin B (DHB) to disrupt F-actin or 100 microM colchicine to collapse microtubules. DHB dramatically softened both G1 and S cells, which reduced the material constants mu c by approximately 65% and b by 20-30%. Colchicine had a limited effect on G1 cells but significantly reduced mu c of S cells (approximately 25%). Thus, F-actin plays the predominate role in determining cell mechanical properties, but disruption of microtubules may also influence the behavior of proliferating cells in a cell cycle-dependent fashion.  相似文献   

7.
8.
In order to provide a more detailed understanding of human plasma fibronectin (PFn) solution structure, we examined the effects of pH and ionic strength (mu) variation on the sedimentation velocities (s20,w), fluorescence polarization-derived mean harmonic rotational relaxation times (rho H), far-ultraviolet (UV) circular dichroism (CD), and intrinsic tryptophan fluorescence of dimeric PFn and the monomeric 190/170-kDa PFn fragment. By comparing the biophysical properties of PFn with those of the 190/170-kDa PFn fragment, we could assess the relative importance of intrasubunit and intersubunit electrostatic forces in the stabilization of PFn structure. The rho H derived from isothermal polarization measurements on 1-pyrenebutyrate conjugated PFn decreased markedly (4.5----1.05-1.23 microseconds) when mu was increased from 0.2 to 1.2 or when the pH was adjusted from 7.4 to 2.0 or 11.0. We also noted a significant decrease in the PFn s20,w (13----8.5-9.6S) under these same solvent conditions. In contrast, the rho H and s20,w of the monomeric 190/170-kDa PFn fragment were relatively insensitive to changes in mu or pH. Computer simulations of the observed pH-dependent changes in the far-UV CD of PFn and the 190/170-kDa PFn fragment revealed only minor differences in protein secondary structure. We also observed only small bathochromic shifts (1-3 nm) in the emission maxima of PFn and 190/170-kDa PFn fragment tryptophan fluorescence under acidic or high mu conditions. These results suggest that minimal changes in PFn tertiary (i.e., intrasubunit) structure occur at pH 2, 11, or at mu = 1.2.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
We have established B cell culture systems in which transfected and stably integrated Ig constructs spontaneously undergo high rates of variable (V) region mutation. Mutation rates were determined using reversion analysis of an Ig V region nonsense codon (Vn). A construct (Vn/gamma2a) in which a Vn was associated with the gamma2a constant region and its intervening and immediate flanking sequences mutated at a high rate of 2.2 x 10(-4)/bp/generation in the NSO myeloma cell line. This same Vn, when associated with the mu constant region (Vn/mu), mutated at a 1000-fold lower rate in NSO. The Vn/gamma2a construct also mutated at high rates in the 18.81 pre-B and the S107 myeloma cell lines and at a low rate in the J558 myeloma cell line. In NSO, the presence of the gamma2a construct raised the mutation rate of the mu construct and the mu decreased the mutation rate of gamma2a. These results suggest that there is both positive and negative regulation of V region mutation and that different cell lines express different combinations and/or amounts of trans-acting factors that are involved in the mutational process.  相似文献   

10.
A highly conserved clonotype has been identified within the repertoire of B cells specific for the negatively charged hapten phthalate. The prototype of this phthalate-specific clonotype is a primary-response hybridoma (2E9) that produces a mu,kappa anti-phthalate antibody. The 2E9 monoclonal antibody was found to share idiotypic determinants with several other independently-derived mu,kappa and gamma 1,kappa anti-phthalate monoclonal antibodies and with a significant proportion of conventional anti-phthalate antibodies derived from all of the BALB/c mice immunized with phthalate-keyhole limpet hemocyanin. Competitive RIA analysis of the 2E9 idiotypic relatedness between primary and secondary response antibodies was consistent with the hypothesis that the primary response mu,kappa antibodies represent a conserved germ-line product, whereas the secondary response to gamma 1,kappa antibodies reflect somatic variants of the 2E9 clonotype. Further analysis with a site-specific anti-idiotype reagent suggests that the idiotypic differences between mu,kappa and gamma 1,kappa monoclonal antibodies occur at positions outside of the combining site. Fine specificity analysis of the monoclonal antibodies expressing the 2E9 cross-reactive idiotype (CRI) also supports this hypothesis. Seven to 35% of the anti-phthalate antibodies after a single immunization with phthalate-KLH and 1 to 10% of the antibodies after a second immunization express the 2E9 CRI. The 2E9 CRI was also found in several other strains of mice, and its expression was associated exclusively with anti-phthalate antibodies.  相似文献   

11.
Prediction of rates of inbreeding in selected populations   总被引:2,自引:0,他引:2  
A method is presented for the prediction of rate of inbreeding for populations with discrete generations. The matrix of Wright's numerator relationships is partitioned into 'contribution' matrices which describe the contribution of the Mendelian sampling of genes of ancestors in a given generation to the relationship between individuals in later generations. These contributions stabilize with time and the value to which they stabilize is shown to be related to the asymptotic rate of inbreeding and therefore also the effective population size, Ne approximately 2N/(mu 2r + sigma 2r), where N is the number of individuals per generation and mu r and sigma 2r are the mean and variance of long-term relationships or long-term contributions. These stabilized values are then predicted using a recursive equation via the concept of selective advantage for populations with hierarchical mating structures undergoing mass selection. Account is taken of the change in genetic parameters as a consequence of selection and also the increasing 'competitiveness' of contemporaries as selection proceeds. Examples are given and predicted rates of inbreeding are compared to those calculated in simulations. For populations of 20 males and 20, 40, 100 or 200 females the rate of inbreeding was found to increase by as much as 75% over the rate of inbreeding in an unselected population depending on mating ratio, selection intensity and heritability of the selected trait. The prediction presented here estimated the rate of inbreeding usually within 5% of that calculated from simulation.  相似文献   

12.
During B lymphocyte differentiation, immunoglobulin heavy chain constant region (CH) genes undergo a unique series of DNA recombination events culminating in the CH class switch. CH switch (S) regions are located 2 kb 5' of each CH gene except delta (i.e. mu, gamma 3, gamma 1, gamma 2b, gamma 2a, epsilon and alpha). We describe the structural features of the gamma 3 switch region. Hybridization experiments show that S gamma 3 has remarkable homology to both S mu and other S gamma regions while S mu possesses limited homology to the other S gamma sequences. However, S mu possesses extensive sequence homology with S epsilon and S alpha. The nucleotide sequence of S gamma 3 reveals higher densities of S mu repetitive sequences (GAGCT and GGGGT) and another S region common sequence (YAGGTTG) than observed for S gamma 1, S gamma 2b or S gamma 2a. In addition, the conservation of S mu like repetitive sequences in S gamma regions is correlated with the 5' leads to 3' gamma gene order (i.e. S gamma 3 greater than S gamma 1 greater than S gamma 2b greater than S gamma 2a). A model is presented which suggests that the unique features of S gamma 3 may allow for successive switches from C mu to any C gamma gene.  相似文献   

13.
To determine the significance of the gamma2 calcium-binding site in fibrin polymerization, we synthesized the fibrinogen variant, gammaD298,301A. We expected these two alanine substitutions to prevent calcium binding in the gamma2 site. We examined the influence of calcium on the polymerization of gammaD298,301A fibrinogen, evaluated its plasmin susceptibility, and solved 2.7 and 2.4 A crystal structures of the variant with the peptide ligands Gly-Pro-Arg-Pro-amide (GPRP) and Gly-His-Arg-Pro-amide (GHRP), respectively. We found that thrombin-catalyzed polymerization of gammaD298,301A fibrinogen was modestly impaired, whereas batroxobin-catalyzed polymerization was significantly impaired relative to normal fibrinogen. Notably, the influence of calcium on polymerization was the same for the variant and for normal fibrinogen. Fibrinogen gammaD298,301A was more susceptible to plasmin proteolysis in the presence of GPRP. This finding suggests structural changes in the near-by "a" polymerization site. Comparisons of the structures revealed minor conformational changes in the gamma294-301 loop that are likely responsible for the weakened "a" site. When considered altogether, the data suggest that the gamma2 calcium-binding site does not significantly modulate polymerization. We cannot, however, rule out the possibility that the weakened "a" polymerization site masks an important role for the gamma2 calcium-binding site in normal polymerization. Somewhat unexpectedly, the structure data showed that GPRP bound to the "b" site and induced the same local conformational changes as GHRP to this site. This structure shows that "A:b" interactions can occur and suggests that these may participate in normal polymerization.  相似文献   

14.
Packing energy calculations are made for two possible sheet structures of polyglycine I, i.e. the antiparallel pleated and rippled sheets. They indicate that the rippled sheet is the more stable structure and that its calculated lattice parameters are close to those experimentally determined. Furthermore, the results on the packing of pleated sheets of polyglycine improve understanding of the well-known model of silk fibroin structure of Marsh et al. (1955). They also suggest that the sheet structures of l-polypeptides with short side-chains should pack in monoclinic unit cells rather than the orthorhombic ones which are observed. A possible origin of this discrepancy is discussed.  相似文献   

15.
A large number of models concerning cultures of genetically engineered bacteria have been described. Among them, some are specifically adapted to continuous cultures and lead to the determination of two variables: (i) the difference in the specific growth rates between plasmid-carrying cell and plasmid-free cells (deltamu) and (ii) the frequency of plasmid loss by plasmid-containing cells (p(r)mu(+)). Until now, studies have been performed on the global expression p(r)mu(+) and deltamu, whose value during continuous assays have been supposed approximately constant (mean value) and not on separate values of both terms p(r) and mu(+), respectively, probability of plasmid loss and specific growth rate of the plasmid-carrying cells. So far these studies do not allow examination of the relationship between these two last parameters. Experimental results were obtained with Escherichia coli C600 galk (GAPDH), a genetically engineered strain that synthetizes an elevated quantity of glyceraldehyde-3-phosphate dehydrogenase (GAPDH). From data obtained during continuous cultures, it is shown that during an assay, deltamu, and p(r)mu(+) do not remain constant. An appropriate mathematical analysis of the expression of mu(-) (specific growth rate of the plasmid-free cells) and mu(+) has been built up. This allows the evaluation of the values of mu(+) and mu(-) during the continuous cultures carried out at different dilution rates. Values of p(r) have been calculated from these data. Indeed our results show that p(r) increases with mu(+). A modeling approach which allows correct simulation of this variation is also proposed. This model is derived from the Hill equation regarding cooperative binding of enzymic type reaction. (c) 1993 John Wiley & Sons, Inc.  相似文献   

16.
A. M. Dwivedi  S. Krimm 《Biopolymers》1982,21(12):2377-2397
A force field has been refined for the 31-helix structure of polyglycine II, using the polyglycine I force field plus previous CαHα…?O force constants as a starting point. Besides force constants associated with the hydrogen bonds, which must change since the hydrogen-bond characteristics are different in the two structures, we have had to modify only 10 force constants from the polyglycine I force field to make it suitable for reproducing the polyglycine II frequencies. Most involve the NCα bond, which is the torsion angle that changes from the I to the II structure. Calculations were done for parallel chain and antiparallel chain crystal structures of polyglycine II, the observed spectra being found to agree best with the latter structure. Since this provides strong evidence for the loss of strict threefold symmetry in the chain, our analysis strengthens the support for the existence of CαHα…?O hydrogen bonds in the structure of polyglycine II.  相似文献   

17.
Streptomyces subtilisin inhibitor (SSI) is known to exist in at least two distinct denatured states, cold-denatured (D') and heat-denatured (D) under acidic conditions. In the present work, we investigated the manner how increasing urea concentration from 0 to 8 M changes the polypeptide chain conformation of SSI that exists initially in the D' and D states as well as in the native state (N), in terms of the secondary structure, the tertiary structure, and the chain form, based on the results of the experiments using circular dichroism (CD), small-angle X-ray scattering (SAXS) and 1H-NMR spectroscopy. Our results indicate that the urea-induced conformational transitions of SSI under typical conditions of D' (pH 1.8, 3 degrees C) occur at least in two steps. In the urea concentration range of 0-2 M (step 1), a cooperative destruction of the tertiary structure occurs, resulting in a mildly denatured state (DU), which may still contain a little amount of secondary structures. In the concentration range of 2-4 M urea (step 2), the DU state gradually loses its residual secondary structure, and increases the radius of gyration nearly to a maximum value. At 4 M urea, the polypeptide chain is highly disordered with highly mobile side chains. Increasing the urea concentration up to 8 M probably results in the more highly denatured or alternatively the stiffer chain conformations. The conformational transition starting from the N state proceeds essentially the same way as in the above scheme in which D' is replaced with N. The conformational transition starting from the D state lacks step 1 because the D state contains no tertiary structures and is similar to the DU state. The fact that similar conformations are reached at urea concentrations above 2 M from different conformations of D', D, and N indicates that the effect of urea dominates in determining the polypeptide conformation of SSI in the denatured states rather than the pH and temperature.  相似文献   

18.
X Li  Y Han  X M Pan 《FEBS letters》2001,507(2):169-173
Adenylate kinase (AK) ages in solution in the presence of DL-dithiothreitol (DTT) with a gradual activity decrease. Upon dilution with 4 M guanidine hydrochloride denatured native and aged AK, both recover to the same activity as the fresh enzyme. Mass spectroscopy and 7-chloro-4-nitrobenz-2-oxa-1,3-diazole chloride modification kinetics studies identify that the residue cysteine-25 of the enzyme reacts with DTT to form an adduct. The formation of the unusual bridging DTT adduct of AK appears to be the result of a stable DTT-protein complex. The K(M) for AMP, ADP and MgATP of the DTT-modified enzyme does not differ significantly from that of the intact enzyme, whereas the secondary and tertiary structures of the enzyme change obviously. These results indicate that cysteine-25 may not be involved directly in substrate binding, but may play an important role in maintaining secondary and tertiary structures of native AK, as well as the conformation interconversion in the catalytic cycle.  相似文献   

19.
We describe here an energy based computer software suite for narrowing down the search space of tertiary structures of small globular proteins. The protocol comprises eight different computational modules that form an automated pipeline. It combines physics based potentials with biophysical filters to arrive at 10 plausible candidate structures starting from sequence and secondary structure information. The methodology has been validated here on 50 small globular proteins consisting of 2-3 helices and strands with known tertiary structures. For each of these proteins, a structure within 3-6 A RMSD (root mean square deviation) of the native has been obtained in the 10 lowest energy structures. The protocol has been web enabled and is accessible at http://www.scfbio-iitd.res.in/bhageerath.  相似文献   

20.
We synthesized and solved the crystalline structure of the oligopeptide acetyl-(glycyl-beta-alanyl)2-NH propyl. The crystal is formed by layers of helical molecules with the same chirality; however, right-handed layers alternate with left-handed ones. Inside every layer, the packing of helices is pseudohexagonal with hydrogen bonds between neighbor molecules. The structure found affords direct support for the model proposed by Crick and Rich for polyglycine II and also provides an interpretation for the structure of a newly found family of polyamides that do not form sheets as observed in most nylon structures.  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