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.
Bis(3,5-dibromosalicyl) fumarate was used to crosslink hemoglobin both in the oxy and deoxy states. This double headed diaspirin was known to crosslink oxy Hb A selectively between Lys 82 beta 1 and Lys 82 beta 2 (Walder, J. A., et al. (1979) Biochemistry 18, 4265) and deoxy Hb A between Lys 99 alpha 1 and Lys 99 alpha 2 (Chatterjee R. Y., et al. (1986) J. Biol. Chem. 261, 9929). The autoxidation at 37 degrees C of oxy alpha 99 crosslinked hemoglobin was found to be 1.8 times as fast as that of Hb A while that of the oxy beta 82 crosslinked hemoglobin was only 1.2 times as fast. After 5 hours the formation of methemoglobin in the alpha crosslinked Hb A is 21.3% compared to 10.8% in beta crosslinked Hb A and 6.4% in Hb A. These results may effect the proposed use of alpha 99 crosslinked hemoglobin as a blood substitute by demonstrating the need for protection from autoxidation during storage.  相似文献   

2.
The double-headed aspirin, bis(3,5-dibromosalicyl) fumarate, has been used to crosslink hemoglobin A between Lys 82 beta 1 and Lys 82 beta 2 (J. A. Walder et al. (1979) Biochemistry 18,4265). Denaturation experiments were used to compare the stability of this crosslinked protein to that of hemoglobin A. Thermal denaturations, done in 0.01 M 4-morpholine-propanesulfonic acid, pH 7, containing 0.9 M guanidine to prevent precipitation at high temperatures, were monitored by changes in absorbance between 190 and 650 nm using a diode array spectrophotometer. The sample was heated from 25 to 70 degrees C at 0.3 degrees C/min. The data were analyzed by using both a two-state model and a novel first derivative method. As expected, methemoglobin A had a single, broad transition with a midpoint of 40.7 degrees C. The crosslinked methemoglobin showed a transition at 57.1 degrees C. Two minor transitions, one of which was apparently due to residual unmodified hemoglobin, were also observed in the crosslinked sample. Thus, a single crosslink between only two of the four subunits can lead to a significantly more stable molecule. These results can be explained by Le Chatelier's principle, since crosslinking prevents dissociation of the beta-subunits and, thereby, holds the entire tetramer together.  相似文献   

3.
Hemoglobins A and S were crosslinked between Lys 82 beta 1 and Lys 82 beta 2 using bis (3,5-dibromosalicyl) fumarate (J. A. Walder et al. (1979) Biochemistry 18, 4265). Thermal denaturation experiments were used to compare the stabilities of the met, cyanomet, and carbonmonoxy forms of these crosslinked hemoglobins to the corresponding uncrosslinked proteins. Uncrosslinked carbonmonoxy- and cyanomethemoglobins had transition temperatures about 11 degrees C higher than the corresponding met samples. The increase in denaturation temperature (Tm) due to crosslinking was 15 degrees C for the methemoglobins, 10 degrees C for the cyanomethemoglobins, and 4 degrees C for the carbonmonoxy ones. There was no significant difference in stability between the met and carbonmonoxy crosslinked proteins. In order of increasing stability the samples were: met Hb S less than met Hb A less than CO Hb S less than CO Hb A = CN-met Hb A less than met XL-Hb S = CO XL-Hb S less than met XL-Hb A = CO XL-Hb A less than CN-met XL-Hb A. The slight decrease in the stability of Hb S (beta 6 Glu----Val) compared to Hb A can be explained by the replacement of an external ionic group by a hydrophobic residue in Hb S. In mixtures of crosslinked and normal Hb A, the Tm of the uncrosslinked material was slightly increased by the presence of the more stable crosslinked hemoglobin. The effects of both crosslinking and cyanide or carbon monoxide binding can be explained by Le Chatelier's principle since both would favor the native form of the protein.  相似文献   

4.
The N-terminal RNA binding domain (RBD) of the human U1A snRNP protein binds tightly and specifically to an RNA hairpin that contains a 10-nucleotide loop. The protein is one of a class of RNA binding proteins that adopts a beta alpha beta beta alpha beta global fold, which in turn forms a four-stranded antiparallel beta-sheet. This sheet forms the primary binding surface for the RNA, as shown by the crosslinking results described here, and in more detail by a recently described co-crystal of this RBD with an RNA hairpin (Oubridge C, et al., 1994, Nature 372:432-438). The RNA hairpin sequence used in the crosslinking experiments, containing 5-iodo-uridine, is a variant of the normal U1 snRNA sequence which is able to form a crosslink with the protein, in contrast to the wild-type sequence, which does not. This single uridine substitution in the 10-nucleotide loop is the site of cross-linking to one tyrosine (Tyr 13) in the beta 1 strand of the U1A N-terminal RBD. This same uridine is also crosslinked to a mutant Tyr 13 Phe RBD, at this Phe 13 substitution.  相似文献   

5.
We use the H-Pex (Thomas et al., this issue) to analyze the main chain interactions in 131 proteins. In antiparallel beta-sheets, the geometry of the N...O bond is: median N...O distances, 2.9 SA, C==O...N angles at 154 degrees and the C alpha--C==O...H angles are dispersed around 3 degrees. In some instances, the other side of the C==O axis is occupied by a HC alpha. As recently supported by Vargas et al. (J Am Chem Soc 2000;122:4750-4755) C alpha H...O and NH...O could cooperate to sheet stability. In alpha-helices, the main chain C==O interact with the NH of their n + 4 neighbor on one side, and with a C beta H or C gamma H on the other side. The median O...N distance (3.0 A) and C==N angle (147 degrees) suggest a canonical H-bond, but the C alpha--C==O...H dihedral angle invalidates this option, since the hydrogen attacks the oxygen at 122 degrees, i.e., between the sp(2) and pi orbitals. This supports that the H-bond is noncanonical. In many instances, the C gamma H or the C beta H of the n + 4 residue stands opposite to the NH with respect to the oxygen. Therefore, we propose that, in alpha-helices, the C gamma H or C beta H and the NH of the n + 4 residue hold the oxygen like an electrostatic pincher. Proteins 2001;43:37-44.  相似文献   

6.
The ability of 5,5'-dithiobis-2-nitrobenzoate (Nbs2) to form a disulfide crosslink between the Cys-190s of the alpha alpha and alpha beta molecular components of rabbit skeletal tropomyosin (Tm) and the Cys-36s and Cys-190s of purified beta beta was studied in separate experiments, as a function of urea concentration in 0.5 M NaCl, 20 mM 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid, pH 7.4, 15 degrees C. In the absence of urea, complete reaction of the Cys-190s of Tm with Nbs2 as well as with 2- and 4-pyridine disulfide quantitatively produced two crosslinked species, alpha-alpha and alpha-beta, in a 60/40 ratio, respectively, visualized as bands on sodium dodecyl sulfate-polyacrylamide gels; similar reactions of beta beta produced both doubly (at Cys-36 and Cys-190) and singly crosslinked species (at Cys-190 as identified by amino acid analysis of separated tryptic peptides). In the presence of 4 M urea where the chains were unfolded and separated, only Nbs-blocked uncrosslinked species were obtained after complete reaction with Nbs2. The loss of Nbs2-crosslinking at increasing [urea] showed that the relative stability of the Cys-containing regions of the three species of Tm, alpha alpha, alpha beta, and beta beta increases in the order Cys-36 of beta beta, Cys-190 of alpha beta, Cys-190 of alpha alpha.  相似文献   

7.
Gonadal function is controlled by lutropins and follitropins, heterodimeric cystine knot proteins that have nearly identical alpha-subunits. These heterodimeric proteins are stabilized by a portion of the hormone-specific beta-subunit termed the "seatbelt" that is wrapped around alpha-subunit loop 2 (alpha 2). Here we show that replacing human chorionic gonadotropin (hCG) alpha 2 residue Lys51 with cysteine or alanine nearly abolished its lutropin activity, an observation that implies that alpha Lys51 has a key role in hormone activity. The activity of the heterodimer containing alpha K51C, but not that containing alpha K51A, was increased substantially when beta-subunit seatbelt residue beta Asp99 was converted to cysteine. As had been reported by others, heterodimers containing alpha K51C and beta D99C were crosslinked by a disulfide. The finding that an intersubunit disulfide restored some of the activity lost by replacing alpha Lys51 suggests that this residue is not crucial for receptor binding or signaling and also that hCG and related hormones may be particularly sensitive to mutations that alter interactions between their subunits. We propose the unique structures of hCG and related family members may permit some subunit movement in the heterodimer, making it difficult to deduce key residues involved in receptor contacts simply by correlating the activities of hormone analogs with their amino acid sequences.  相似文献   

8.
The change in the structural stability of Escherichia coli ribonuclease HI (RNase HI) due to single amino acid substitutions has been estimated computationally by the stability profile of mutant protein (SPMP) [Ota, M., Kanaya, S. Nishikawa, K., 1995. Desk-top analysis of the structural stability of various point mutations introduced into ribonuclease H. J. Mol. Biol. 248, 733-738]. As well, an effective strategy using random mutagenesis and genetic selection has been developed to obtain E. coli RNase HI mutants with enhanced thermostability [Haruki, M., Noguchi, E., Akasako, A., Oobatake, M., Itaya, M., Kanaya, S., 1994. A novel strategy for stabilization of Escherichia coli ribonuclease HI involving a screen for an intragenic suppressor of carboxyl-terminal deletions. J. Biol. Chem. 269, 26904-26911]. In this study, both methods were combined: random mutations were individually introduced to Lys99-Val101 on the N-terminus of the alpha-helix IV and the preceding beta-turn, where substitutions of other amino acid residues were expected to significantly increase the stability from SPMP, and then followed by genetic selection. Val101 to Ala, Gln, and Arg mutations were selected by genetic selection. The Val101-->Ala mutation increased the thermal stability of E. coli RNase HI by 2.0 degrees C in Tm at pH 5.5, whereas the Val101-->Gln and Val101-->Arg mutations decreased the thermostability. Separately, the Lys99-->Pro and Asn100-->Gly mutations were also introduced directly. The Lys99-->Pro mutation increased the thermostability of E. coli RNase HI by 1.8 degrees C in Tm at pH 5.5, whereas the Asn100-->Gly mutation decreased the thermostability by 17 degrees C. In addition, the Lys99-->Pro mutation altered the dependence of the enzymatic activity on divalent metal ions.  相似文献   

9.
High-resolution proton nuclear magnetic resonance studies of hemoglobins Providence-Asn (beta82EF6 Lys replaced by Asn) and Providence-Asp (beta82EF6 Lys replaced by Asp) show that different amino acid substitutions at the same position in the hemoglobin molecule have different effects on the structure of the protein molecule. Hemoglobin Providence-Asp appears to be in a low-affinity tertiary structure in both the deoxy and carbonmonoxy forms. Deoxyhemoglobin Providence-Asn has its beta heme resonance shifted downfield slightly from its position in normal adult hemoglobin; however, the tertiary structures of the heme pocket of hemoglobins A and Providence-Asn are very similar when both proteins are in the carbonmonoxy form. These results are consistent with the oxygen equilibrium measurements of Bonaventura, J., et al. [(1976) J. Biol. Chem. 251, 7563] which show that both Hb Providence-Asn and Hb Providence-Asp have oxygen affinities lower than normal adult hemoglobin, with Hb Providence-Asp having the lowest. Our studies of the effects of sodium chloride on the hyperfine shifted proton resonances of deoxyhemoglobins A, Providence-Asn, and Providence-Asp indicate that the beta82EF6 lysine is probably one, but not the only binding site for chloride ions.  相似文献   

10.
Dissociation of alpha beta and alpha gamma dimers of human hemoglobins (Hb) A and F into monomers was studied by alpha chain exchange (Shaeffer, J. R., McDonald, M. J., Turci, S. M., Dinda, D. M., and Bunn, H. F. (1984) J. Biol. Chem. 259, 14544-14547). Unlabeled carbonmonoxy-Hb A was incubated with trace amounts of preparatively purified, native, 3H-alpha subunits in 10 mM sodium phosphate, pH 7.0, at 25 degrees C. At appropriate times, free alpha monomers were separated from Hb A tetramers by anion exchange high performance liquid chromatography. Transfer of radioactivity from the alpha chain pool into Hb A was measured, yielding a first order dimer dissociation rate constant, k2 = (3.2 +/- 0.3) X 10(-3) h-1. The Arrhenius plot of k2 was linear between 7 and 37 degrees C, yielding an enthalpy of activation of 23 kcal/alpha beta dimer. As the chloride concentration was raised from 0 to 0.2 M, the dissociation rate increased 3-fold; with higher salt concentrations, however, the rate gradually returned to baseline. This rate was not altered by raising the pH from 6.5 to 7.2, but as pH was further raised to 8.4, kappa 2 increased about 3-fold. Hb F, which has an increased stability at alkaline pH, dissociated into alpha and gamma monomers 3 times more slowly than Hb A. Moreover, the dimer-monomer dissociation of Hb F was characterized by a significantly reduced pH dependence. These results demonstrate that both alpha beta and alpha gamma dimers of Hb A and Hb F dissociate reversibly into monomers under physiologic conditions. The differential pH dependence for dimer dissociation between Hb A and Hb F suggests that specific amino acid replacement at the alpha 1 gamma 1 interface confers increased resistance to alkaline denaturation.  相似文献   

11.
Structure of Saccharomyces cerevisiae alg3, sec18 mutant oligosaccharides   总被引:3,自引:0,他引:3  
Asparagine-linked oligosaccharides are synthesized by transfer of Glc3Man9GlcNAc2 from dolichol pyrophosphate to nascent polypeptides. Assembly of the precursor proceeds by highly ordered sequential addition of mannose and glucose to form Glc3Man9GlcNAc2-P-P-dolichol. Yeast mutants in asparagine-linked glycosylation (alg), generated by an 3H-Man suicide technique, were assigned to eight complementation groups which define steps in oligosaccharide-lipid synthesis (Huffaker, T.C., and Robbins, P.W. (1982) J. Biol. Chem. 257, 3203-3210). Alg3 invertase oligosaccharides are resistant to endo-beta-N-acetylglucosaminidase H, and the lipid-oligosaccharide pool yields Man5Glc-NAc2, suggesting its structure may be that from mammalian cells lacking Man-P-dolichol (Chapman, A., et al. (1980) J. Biol. Chem. 255, 4441-4446). To test this supposition, the endoplasmic reticulum form of invertase derepressed in alg3,sec18 yeast at 37 degrees C was isolated as a source of oligosaccharides whose processing beyond glucose and/or mannose trimming, if involved, would be prevented. Man8GlcNAc2 and Man5GlcNAc2 were released by peptide-N-glycosidase F from alg3,sec18 invertase in a 1:5 molar ratio. 1H NMR spectroscopy revealed Man8GlcNAc2 to be the alpha 1,2-mannosidase-trimming product described earlier (Byrd, J. C., Tarentino, A. L., Maley, F., Atkinson, P. H., and Trimble, R. B. (1982) J. Biol. Chem. 257, 14657-14666), while Man5GlcNAc2 was Man alpha 1, 2Man alpha 1,2Man alpha 1,3(Man alpha 1,6)Man beta 1,4GlcNAc beta 1, 4GlcNAc. This provides a structural proof for the lipid-linked Man5GlcNAc2 originally proposed from enzymatic and chemical analyses of the radiolabeled mammalian precursor. Experimental evidence indicates that, unlike the mammalian cell mutants which are unable to synthesize Man-P-dolichol, alg3 yeast accumulate Man5GlcNAc2-P-P-dolichol due to a defective alpha 1,3-mannosyltransferase required for the next step in oligosaccharide-lipid elongation.  相似文献   

12.
Cytochrome c2 (cyt) is the mobile electron donor to the reaction center (RC) in photosynthetic bacteria. The electrostatic interactions involved in the dynamics of docking of cyt onto the RC were examined by double mutant studies of the rates of electron transfer between six modified Rhodobacter sphaeroides RCs in which negatively charged acid residues were replaced with Lys and five modified Rhodobacter capsulatus Cyt c2 molecules in which positively charged Lys residues were replaced with Glu. We measured the second-order rate constant, k2, for electron transfer from the reduced cyt to the oxidized primary donor on the RC, which reflects the energy of the transition state for the formation of the active electron transfer complex. Strong interactions were found between Lys C99 and Asp M184/Glu M95, and between Lys C54 and Asp L261/Asp L257. The interacting residues were found to be located close to each other in the recently determined crystal structure of the cyt-RC complex [Axelrod, H., et al. (2002) J. Mol. Biol. (in press)]. The interaction energies were approximately inversely proportional to the distances between charges. These results support earlier suggestions [Tetreault, M., et al. (2001) Biochemistry 40, 8452-8462] that the structure of the transition state in solution resembles the structure of the cyt-RC complex in the cocrystal and indicate that specific electrostatic interactions facilitate docking of the cyt onto the RC in a configuration optimized for both binding and electron transfer. The specific interaction between Asp M184 and Lys C99 may help to nucleate short-range hydrophobic contacts.  相似文献   

13.
14.
The procedure of Perrella et al. (Perrella, M., Benazzi, L., Cremonesi, L., Vesely, S., Viggiano, G., and Rossi-Bernardi, L. (1983) J. Biol. Chem. 258, 4511-4517) for trapping the intermediate compounds between human hemoglobin and carbon monoxide was validated by quantitatively determining during the approach to equilibrium all the species present in a solution containing large amounts of intermediates. An accurate estimate of the intermediate compounds at 50% carbon monoxide saturation in 0.1 M KCl, pH 7, at 22 degrees C, allowed the calculation, according to Adair's scheme, of the four equilibrium constants. At 50% ligand saturation, the pool of intermediate species was about 12% of the total. A slightly greater concentration of tri-liganded than mono-liganded species was found. Carbon monoxide bound to beta chains in slightly greater excess with respect to alpha chains in both the mono- and tri-liganded species. The symmetrical bi-liganded intermediates, alpha 2 beta CO2 and alpha 2CO beta 2, were absent. The nature of the bi-liganded intermediate found to be present in detectable amounts by our technique has yet to be clarified: it could be either the asymmetrical species (alpha beta) (alpha CO beta CO) and (alpha beta CO) (alpha CO beta) or both of them. Such a finding on the functional heterogeneity among the four possible bi-liganded intermediates is consistent with hypotheses of the existence of more than two quaternary structures in the course of ligand binding to hemoglobin.  相似文献   

15.
Rat blood plasma contains three high molecular weight thiol ester-containing proteinase inhibitors, alpha 1-macroglobulin (alpha 1M), alpha 1-inhibitor III (alpha 1I3), and alpha 2-macroglobulin (alpha 2M). Rat serums have been analyzed using a two-dimensional gel electrophoretic technique which optimizes recovery of high molecular weight proteins. alpha 1M, and (alpha beta)4-tetramer in native solution, separated in the second sodium dodecyl sulfate-containing electrophoretic dimension as a disulfide-linked (alpha beta)2-dimer with an approximate Mr of 360 kDa. alpha 1I3 separated in the gels as a single 190-kDa polypeptide. It is also a monomer in native solution by ultracentrifugation criteria. Native rat alpha 2M is a tetramer, but it separates in the gels as a disulfide-linked dimer with an Mr of approximately 360 kDa. The kinetics of changes in concentration of these proteins during the induction of polyarthritis was also measured by quantitative immunoelectrophoresis. In rats with adjuvant-induced polyarthritis, the concentration of alpha 1I3 dramatically decreases and alpha 2M appears and continues to increase in a biphasic manner for 2 weeks. The alpha 1M concentration remains relatively constant. All three macroglobulins were purified utilizing modern rapid chromatographic techniques, and parallel comparisons of their native physicochemical properties were carried out. The N-terminal sequence of the alpha-chain of rat alpha 1M was also shown to share sequence homology with that of alpha 2M. In agreement, Esnard et al. (Esnard, F., Gutman, N., El Moujahed, A., and Gauthier, F. (1985) FEBS Lett. 182, 125-129) recently reported that alpha 1I3 also contains a thiol ester bond, as do alpha 1M and alpha 2M, since it reacts covalently with [14C]methylamine and is cleaved autolytically at 80 degrees C. We have examined negatively stained preparations of native, trypsin-treated, and methylamine-treated human alpha 2M, rat alpha 2M, and rat alpha 1M in the electron microscope. Trypsin appears to convert globular ring-shaped native molecules to rectangular box-like structures, in agreement with the conclusions of a recent report on human alpha 2M (Tapon-Bretaudiere, J., Bros, A., Couture-Tosi, E., and Delain, E. (1985) EMBO J. 4, 85-89).  相似文献   

16.
Hemoglobin can be specifically carboxymethylated at its NH2-terminal amino groups (i.e. HbNHCH2COO-) to form the derivatives alpha 2Cm beta 2, alpha 2 beta 2Cm, and alpha 2Cm beta 2Cm, where Cm represents carboxymethyl. Previous studies (DiDonato, A., Fantl, W. J., Acharya, A. S., and Manning, J. M. (1983) J. Biol. Chem. 258, 11890-11895) suggested that these derivatives could be used as stable analogues of the corresponding carbamino (Hb-NHCOO-) forms of hemoglobin, adducts that are generated reversibly in vivo when CO2 combines with alpha-amino groups. In this paper we present x-ray diffraction studies of both carbamino hemoglobin and carboxymethylated hemoglobin that verify this proposal and we use the carboxymethylated derivatives to study the functional consequences of placing a covalently bound carboxyl group at the NH2 terminus of each hemoglobin subunit. Our studies also provide additional information concerning the oxygen-linked binding of anions and protons to Val-1 alpha. Difference electron density analysis of deoxy alpha 2Cm beta 2Cm versus the unmodified deoxyhemoglobin tetramer (deoxy alpha 2 beta 2) shows that the covalently bound carboxyl moieties replace inorganic anions that are normally bound to the free NH2-terminal amino groups in crystals of native deoxyhemoglobin grown from solutions of concentrated (2.3 M) ammonium sulfate. In the case of the beta-subunits, the carboxymethyl group replaces an inorganic anion normally bound between the alpha-amino group of Val-1 beta, the epsilon-amino group of Lys-82 beta, and backbone NH groups at the NH2-terminal end of the F'-helix. In the case of the alpha-subunits, the carboxymethyl group replaces an anion that is normally bound between the alpha-amino group of Val-1 alpha and the beta-OH group of Ser-131 alpha. A corresponding difference electron map of carbamino deoxyhemoglobin in low-salt (50 mM KCl) crystals shows that CO2 bound in the form of carbamate occupies the same two anion binding sites. The alkaline Bohr effect of alpha 2Cm beta 2 is only marginally lower (approximately 7%) than that of alpha 2 beta 2. Previous studies (Kilmartin, J. V., 1977) have shown that about 30% of the alkaline Bohr effect is the result of an oxygen-linked change in the pK alpha of Val-1 alpha, and O'Donnell et al., 1979, found that this portion of the Bohr effect is the result of the oxygen-linked binding of chloride to Val-1 alpha.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

17.
Integrins alpha(1)beta(1) and alpha(2)beta(1) are two major collagen receptors on the surface of eukaryotic cells. Binding to collagen is primarily due to an A-domain near the N terminus of the alpha chains. Previously, we reported that recombinant A-domain of alpha(1)beta(1) (alpha(1)A) had at least two affinity classes of binding sites in type I collagen (Rich, R. L., et al. (1999) J. Biol. Chem. 274, 24906-24913). Here, we compared the binding of the recombinant A-domain of alpha(2)beta(1) (alpha(2)A) to type I collagen with that of alpha(1)A using surface plasmon resonance and showed that alpha(2)A exhibited only one detectable class of binding sites in type I collagen, with a K(D) of approximately 10 microm at approximately 3 binding sites per collagen molecule. We further demonstrated that alpha(1)A and alpha(2)A competed with each other for binding to type I collagen in enzyme-linked immunosorbent assay (ELISA), suggesting that the binding sites in collagen for the two A-domains overlap or are adjacent to each other. By using rotary shadowing, the complexes of alpha(1)A- and alpha(2)A-procollagen were visualized. Morphometric analyses indicated three major binding regions (near the N terminus, in the central part, and near the C terminus) along the type I procollagen molecule for both A-domains. The positions of the respective binding regions for alpha(1)A and alpha(2)A were overlapping with or adjacent to each other, consistent with the ELISA results. Analysis of the sequences of type I collagen revealed that GER or GER-like motifs are present at each of the binding regions, and notably, the central region contains the GFOGER sequence, which was previously identified as a high affinity site for both alpha(1)A and alpha(2)A (Knight, C. G., et al. (2000) J. Biol. Chem. 275, 35-40). Peptides containing GLOGERGRO (peptide I, near the N terminus), GFOGERGVQ (peptide II, central), and GASGERGPO (peptide III, near the C terminus) were synthesized. Peptides I and II effectively inhibited the binding of alpha(1)A and alpha(2)A to type I collagen, while peptide III did so moderately. The N-terminal site in type I collagen has the sequence GLOGER in all three chains. Thus, it seems that peptide I represents a newly discovered native high affinity site for alpha(1)A and alpha(2)A.  相似文献   

18.
The conformation of Forssman glycolipid, GalNAc alpha 1-3GalNAc beta 1-3Gal alpha 1-4Gal beta 1-4Glc beta 1-1ceramide, was analysed with the aid of the rotating frame NOE and Hartmann-Hahn spectroscopy. NOE contacts between C-, O-, and N-linked protons were used for distance mapping. The glycosidic bonds that are common to globotriaosylceramide and globoside showed a similar flexibility as found for these compounds [Poppe et al., (1990) Eur. J. Biochem. 189, 313-325; J. Am. Chem. Soc. 112, 7762-7771]. In contrast, the conformational mobility of the terminal GalNAc alpha 1-3GalNAc beta linkage appears to be restrained. A new approach, based on 2D exchange spectroscopy, was proposed for revealing of spatial proximities between exchangeable protons in Me2SO solution.  相似文献   

19.
Integrin-extracellular matrix (ECM) interactions in two-dimensional (2D) culture systems are widely studied (Goldstein and DiMilla, 2002. J Biomed. Mater. Res. 59, 665-675; Koo et al., 2002. J. Cell Sci. 115, 1423-1433). Less understood is the role of the ECM in promoting intercellular cohesion in three-dimensional (3D) environments. We have demonstrated that the alpha5beta1-integrin mediates strong intercellular cohesion of 3D cellular aggregates (Robinson et al., 2003. J. Cell Sci. 116, 377-386). To further investigate the mechanism of alpha5beta1-mediated cohesivity, we used a series of chimeric alpha5beta1-integrin-expressing cells cultured as multilayer cellular aggregates. In these cell lines, the alpha5 subunit cytoplasmic domain distal to the GFFKR sequence was truncated, replaced with that of the integrin alpha4, the integrin alpha2, or maintained intact. Using these cells, alpha5beta1-integrin-mediated cell aggregation, compaction and cohesion were determined and correlated with FN matrix assembly. The data presented demonstrate that cells cultured in the absence of external mechanical support can assemble a FN matrix that promotes integrin-mediated aggregate compaction and cohesion. Further, inhibition of FN matrix assembly blocks the intercellular associations required for compaction, resulting in cell dispersal. These results demonstrate that FN matrix assembly contributes significantly to tissue cohesion and represents an alternative mechanism for regulating tissue architecture.  相似文献   

20.
The mass-driven assembly of spectrin dimers to form tetramers involves two equal head-to-head alpha-beta associations and requires at least 30 degrees C for interconversion to occur readily. In this paper, the properties of tetramer formation were investigated using two complementary univalent peptides (the alpha I domain and beta monomers). Since the alpha I domain lacks an essential nucleation site required for side-to-side (lateral) heterodimer assembly [Speicher et al. (1992) J. Biol. Chem. 267, 14775-14782], these two peptides can only assemble head-to-head at a single site. This head-to-head assembly readily occurs at lower temperatures, indicating the temperature barrier for dimer-tetramer interconversion is caused by a conformational constraint of the dimer. This constraint, a closed hairpin loop, is released when the laterally associated partner is removed. The univalent alpha I-beta binding affinity at 37 degrees C (Ka = 1.4 x 10(5) M-1) is similar to the dimer-tetramer association constant at the same temperature. As the temperature is decreased from 37 to 0 degrees C, the alpha I-beta binding affinity increases about 32-fold. In contrast with head-to-head associations involving dimers, the second-order rate constants of two complementary univalent peptides (i.e., alpha I and beta) are dramatically higher, and the estimated activation energy (about 50 kJ mol-1) is about 5-fold lower. An open dimer conformation is an obligatory high-energy intermediate required for dimer-tetramer interconversion, and opening the dimer hairpin loop contributes about 190 kJ mol-1 to the activation energy for tetramer association.(ABSTRACT TRUNCATED AT 250 WORDS)  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