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.
The blood coagulation protein factor XI (FXI) consists of a pair of disulfide-linked chains each containing four apple domains and a catalytic domain. The apple 4 domain (A4; F272-E362) mediates non-covalent homodimer formation even when the cysteine involved in an intersubunit disulfide is mutated to serine (C321S). To understand the role of non-covalent interactions stabilizing the FXI dimer, equilibrium unfolding of wild-type A4 and its C321S variant was monitored by circular dichroism, intrinsic tyrosine fluorescence and dynamic light scattering measurements as a function of guanidine hydrochloride concentration. Global analysis of the unimolecular unfolding transition of wild-type A4 revealed a partially unfolded equilibrium intermediate at low to moderate denaturant concentrations. The optically detected equilibrium of C321S A4 also fits best to a three-state model in which the native dimer unfolds via a monomeric intermediate state. Dimer dissociation is characterized by a dissociation constant, K(d), of approximately 90 nM (in terms of monomer), which is in agreement with the dissociation constant measured independently using fluorescence anisotropy. The results imply that FXI folding occurs via a monomeric equilibrium intermediate. This observation sheds light on the effect of certain naturally occurring mutations, such as F283L, which lead to intracellular accumulation of non-native forms of FXI. To investigate the structural and energetic consequences of the F283L mutation, which perturbs a cluster of aromatic side-chains within the core of the A4 monomer, it was introduced into the dissociable dimer, C321S A4. NMR chemical shift analysis confirmed that the mutant can assume a native-like dimeric structure. However, equilibrium unfolding measurements show that the mutation causes a fourfold increase in the K(d) value for dissociation of the native dimer and a 1 kcal/mol stabilization of the monomer, resulting in a highly populated intermediate. Since the F283 side-chain does not directly participate in the dimer interface, we propose that the F283L mutation leads to increased dimer dissociation by stabilizing a monomeric state with altered side-chain packing that is unfavorable for homodimer formation.  相似文献   

2.
Coagulation factor XI (FXI) is a covalent homodimer consisting of two identical subunits of 80 kDa linked by a disulfide bond formed by Cys-321 within the Apple 4 domain of each subunit. Because FXI(C321S) is a noncovalent dimer, residues within the interface between the two subunits must mediate its homodimeric structure. The crystal structure of FXI demonstrates formation of salt bridges between Lys-331 of one subunit and Glu-287 of the other subunit and hydrophobic interactions at the interface of the Apple 4 domains involving Ile-290, Leu-284, and Tyr-329. FXI(C321S), FXI(C321S,K331A), FXI(C321S,E287A), FXI(C321S,I290A), FXI(C321S,Y329A), FXI(C321S,L284A), FXI(C321S,K331R), and FXI(C321S,H343A) were expressed in HEK293 cells and characterized using size exclusion chromatography, analytical ultracentrifugation, electron microscopy, and functional assays. Whereas FXI(C321S) and FXI(C321S,H343A) existed in monomer/dimer equilibrium (K(d) approximately 40 nm), all other mutants were predominantly monomers with impaired dimer formation by analytical ultracentrifugation (K(d)=3-38 microm). When converted to the active enzyme, FXIa, all the monomeric mutants activated FIX similarly to wild-type dimeric FXIa. In contrast, these monomeric mutants could not be activated efficiently by FXIIa, thrombin, or autoactivation in the presence of dextran sulfate. We conclude that salt bridges formed between Lys-331 of one subunit and Glu-287 of the other together with hydrophobic interactions at the interface, involving residues Ile-290, Leu-284, and Tyr-329, are essential for homodimer formation. The dimeric structure of FXI is essential for normal proteolytic activation of FXI by FXIIa, thrombin, or FXIa either in solution or on an anionic surface but not for FIX activation by FXIa in solution.  相似文献   

3.
Although bovine beta-lactoglobulin assumes a monomeric native structure at pH 3 in the absence of salt, the addition of salts stabilizes the dimer. Thermodynamics of the monomer-dimer equilibrium dependent on the salt concentration were studied by sedimentation equilibrium. The addition of NaCl, KCl, or guanidine hydrochloride below 1 M stabilized the dimer in a similar manner. On the other hand, NaClO(4) was more effective than other salts by about 20-fold, suggesting that anion binding is responsible for the salt-induced dimer formation, as observed for acid-unfolded proteins. The addition of guanidine hydrochloride at 5 M dissociated the dimer into monomers because of the denaturation of protein structure. In the presence of either NaCl or NaClO(4), the dimerization constant decreased with an increase in temperature, indicating that the enthalpy change (DeltaH(D)) of dimer formation is negative. The heat effect of the dimer formation was directly measured with an isothermal titration calorimeter by titrating the monomeric beta-lactoglobulin at pH 3.0 with NaClO(4). The net heat effects after subtraction of the heat of salt dilution, corresponding to DeltaH(D), were negative, and were consistent with those obtained by the sedimentation equilibrium. From the dependence of dimerization constant on temperature measured by sedimentation equilibrium, we estimated the DeltaH(D) value at 20 degrees C and the heat capacity change (DeltaC(p)) of dimer formation. In both NaCl and NaClO(4), the obtained DeltaC(p) value was negative, indicating the dominant role of burial of the hydrophobic surfaces upon dimer formation. The observed DeltaC(p) values were consistent with the calculated value from the X-ray dimeric structure using a method of accessible surface area. These results indicated that monomer-dimer equilibrium of beta-lactoglobulin at pH 3 is determined by a subtle balance of hydrophobic and electrostatic effects, which are modulated by the addition of salts or by changes in temperature.  相似文献   

4.
Sinha D  Marcinkiewicz M  Lear JD  Walsh PN 《Biochemistry》2005,44(30):10416-10422
Factor XI, unlike other coagulation proteins, is a homodimer of two identical subunits linked by a single disulfide bond formed by Cys321. The present study was undertaken to understand the physiological significance of the dimeric nature of factor XI. We have expressed a mutant FXI/G326C in which the Gly326 residue of factor XI has been mutated to Cys326, reasoning that Cys321 would form an intrachain disulfide bond with Cys326 as in prekallikrein, a plasma protein that exists as a monomer even with 58% amino acid sequence identity and a domain structure very similar to factor XI. No free thiol could be detected in the expressed protein, and it migrated as a monomer on nonreduced SDS-PAGE. In physiological buffer, however, the protein was found to exist in a state of monomer-dimer equilibrium as assessed by gel-filtration chromatography and ultracentrifugation studies (K(d) approximately 36 nM). Functional studies revealed that FXI/G326C was indistinguishable from plasma factor XI in a plasma-clotting assay and in a factor IX activation assay both in the presence and absence of activated platelets even at concentrations at which less than 5% of the mutant exists as dimers. We conclude that, for optimal function in the presence of activated platelets, a preformed dimer of factor XI is not required.  相似文献   

5.
1. The fluorescence polarization, P, of FAD increased on complex formation with the apoenzyme of D-amino acid oxidase [D-amino acid: O2 ocidoreductase (deaminating), EC 1.4.3.3]. The time course of the increase was monophasic. The values of P were extimated to be 0.04, 0.4, and 0.4 for FAD, the enzyme and the enzyme-benzoate complex, respectively. 2. The value of P of the enzyme is dependent on its concentration, indicating that the degrees of dissociation of FAD in the monomer and dimer are different. The dissociation constant was calculated to be 7 times 10-minus 7 M for the monomeric form of the enzyme. This value is far larger than the value for the dimeric form of the enzyme, 1 times 10-minus 8 M, calculated from equilibrium dialysis data. 3. Changes in fluorescence polarization of the enzyme due to changes in solution pH or temperature can be explained in terms of the monomer-dimer equilibrium.  相似文献   

6.
7.
The monomer-dimer association constant of solubilized and delipidated sarcoplasmic reticulum ATPase was measured by large zone elution gel chromatography in the presence of a high concentration (18.6 mM) of the nonionic detergent dodecyloctaethylene glycol monoether (C12E8) and of different ATPase protein concentrations in the range of 0.74 (6.4 nM monomers) to 30 (0.26 microM monomers) microgram/ml. The association equilibrium constant (Ka) obtained from the concentration-dependent dissociation curve was 9.37 X 10(7) M-1 at 24 degrees C. The derived free energy change (delta G0) for the monomer-dimer association was -10.8 kcal/mo, reflecting a high degree of tightness between inter-subunit domains in soluble dimeric ATPase. A steep dissociation curve within a short natural logarithmic span (2.5 units) was obtained when the degree of dissociation increased from 0.1 to 0.9, suggesting that a conformational drift accompanies the dissociation of soluble dimeric ATPase. A unique leading boundary was formed in the large zone chromatographies, indicating a reversible equilibrium which was rapid when compared to the time taken for the chromatographic run. Enzymatic activity was continuously monitored in the eluate, revealing that soluble ATPase at different degrees of dissociation was active.  相似文献   

8.
The monomer-dimer equilibrium for the human immunodeficiency virus type 1 (HIV-1) protease has been investigated under physiological conditions. Dimer dissociation at pH 7.0 was correlated with a loss in beta-sheet structure and a lower degree of ANS binding. An autolysis-resistant mutant, Q7K/L33I/L63I, was used to facilitate sedimentation equilibrium studies at neutral pH where the wild-type enzyme is typically unstable in the absence of bound inhibitor. The dimer dissociation constant (KD) of the triple mutant was 5.8 microM at pH 7.0 and was below the limit of measurement (approximately 100 nM) at pH 4.5. Similar studies using the catalytically inactive D25N mutant yielded a KD value of 1.0 microM at pH 7.0. These values differ significantly from a previously reported value of 23 nM obtained indirectly from inhibitor binding measurements (Darke et al., 1994). We show that the discrepancy may result from the thermodynamic linkage between the monomer-dimer and inhibitor binding equilibria. Under conditions where a significant degree of monomer is present, both substrates and competitive inhibitors will shift the equilibrium toward the dimer, resulting in apparent increases in dimer stability and decreases in ligand binding affinity. Sedimentation equilibrium studies were also carried out on several drug-resistant HIV-1 protease mutants: V82F, V82F/I84V, V82T/I84V, and L90M. All four mutants exhibited reduced dimer stability relative to the autolysis-resistant mutant at pH 7.0. Our results indicate that reductions in drug affinity may be due to the combined effects of mutations on both dimer stability and inhibitor binding.  相似文献   

9.
Hays JL  Watowich SJ 《Biochemistry》2004,43(32):10570-10578
Although oligomerization of receptor tyrosine kinases (RTKs) is necessary for receptor activation and signaling, a quantitative understanding of how oligomerization mediates these critical processes does not exist. We present a comparative thermodynamic analysis of functionally active dimeric and functionally inactive monomeric soluble analogues of the c-MET RTK, which clearly reveal that oligomerization regulates the binding affinity and binding kinetics of the kinase toward ATP and tyrosine-containing peptide substrates. Thermodynamic binding data for oligomeric c-MET were obtained from the dimeric TPR-MET oncoprotein, a functionally active fusion derivative of the c-MET RTK. This naturally occurring oncoprotein contains the cytoplasmic domain of c-MET fused to a coiled coil dimerization domain from the nuclear pore complex. Comparative data were obtained from a soluble monomeric kinase compromising the c-MET cytoplasmic domain (cytoMET). Significantly, under equilibrium binding conditions, the oligomeric phosphorylated kinase showed a significantly lower dissociation constant (K(d,dimer) = 11 microM) for a tyrosine-containing peptide derived from the C-terminal tail of the c-MET RTK when compared to the phosphorylated monomeric kinase cytoMET (K(d,monomer) = 140 microM). Surprisingly, equilibrium dissociation constants measured for the kinase and ATP were independent of the oligomerization state of the kinase (approximately 10 microM). Stopped-flow analysis of peptide substrate binding showed that the association rate constants (k(2)) differed 2-fold and dissociation rate constants (k(-2)) differed 10-fold when phosphorylated TPR-MET was compared to phosphorylated cytoMET. ATP binding abrogated the differences in k(2) rates observed between the two oligomeric states of the c-MET cytoplasmic domain. These results clearly imply that oligomerization induces important thermodynamic and conformational changes in the substrate binding regions of the c-MET protein and provide quantitative mechanistic insights into the necessary role of oligomerization in RTK activation.  相似文献   

10.
It has been previously reported that the structure of a human pancreatic ribonuclease variant, namely PM8, constitutes a dimer by the exchange of an N-terminal domain, although in an aqueous solution it is found mainly as a monomer. First, we investigated the solution conditions that favour the dimerization of this variant. At 29 degrees C in a 20% (v/v) ethanol buffer, a significant fraction of the protein is found in dimeric form without the appearance of higher oligomers. This dimer was isolated by size-exclusion chromatography and the dimerization process was studied. The dissociation constant of this dimeric form is 5 mm at 29 degrees C. Analysis of the dependence of the dimerization process on the temperature shows that unlike bovine pancreatic ribonuclease, a decrease in the temperature shifts the monomer-dimer equilibrium to the latter form. We also show that a previous dissociation of the exchangeable domain from the main protein body does not take place before the dimerization process. Our results suggest a model for the dimerization of PM8 that is different to that postulated for the dimerization of the homologous bovine pancreatic ribonuclease. In this model, an open interface is formed first and then intersubunit interactions stabilize the hinge loop in a conformation that completely displaces the equilibrium between nonswapped and swapped dimers to the latter one.  相似文献   

11.
pH-dependent conformational changes are known to occur in dimeric procaspase-3, and they have been shown to affect the rate of automaturation. We studied the equilibrium unfolding of procaspase-3(C163S) as a function of pH (between pH 8.5 and pH 4) in order to examine these changes in the context of folding and stability. The data show that the procaspase dimer undergoes a pH-dependent dissociation below pH 5, so that the protein is mostly monomeric at pH 4. Consistent with this, the dimer unfolds via a four-state process between pH 8.5 and pH 4.75, in which the native dimer isomerizes to a dimeric intermediate, and the dimeric intermediate dissociates to a monomer, which then unfolds. In contrast, a small protein concentration dependence was observed by circular dichroism, but not by fluorescence emission, at pH 4.5 and pH 4.2. There was no protein-concentration dependence to the data collected at pH 4. Overall, the results are consistent with the redistribution of the population of native dimer (N(2)) to dimeric intermediate (I(2)) to monomeric intermediate (I), as the pH is lowered so that at pH 4, the "native" ensemble resembles the monomeric intermediate (I) observed during unfolding at higher pH. An emerging picture of the monomeric procaspase is discussed. Procaspase-3 is most stable at pH approximately 7 (24-26 kcal/mol), and while the stability decreased with pH, it was observed that dimerization contributes the majority (>70%) of the conformational free energy.  相似文献   

12.
Trigger Factor (TF) is the first chaperone that interacts with nascent chains of cytosolic proteins in Escherichia coli. Although its chaperone activity requires association with ribosomes, TF is present in vivo in a 2-3 fold molar excess over ribosomes and a fraction of it is not ribosome-associated after cell lysis. Here we show that TF follows a three-state equilibrium. Size exclusion chromatography, crosslinking and analytical ultracentrifugation revealed that uncomplexed TF dimerizes with an apparent Kd of 18 microM. Dimerization is mediated by the N-terminal ribosome binding domain and the C-terminal domain of TF, whereas the central peptidyl prolyl isomerase (PPlase) and substrate binding domain does not contribute to dimerization. Crosslinking experiments showed that TF is monomeric in its ribosome-associated state. Quantitative analysis of TF binding to ribosomes revealed a dissociation constant for the TF-ribosome complex of approximately 1.2 microM. From these data we estimate that in vivo most of the ribosomes are in complex with monomeric TF. Uncomplexed TF, however, is in a monomer-dimer equilibrium with approximately two thirds of TF existing in a dimeric state.  相似文献   

13.
Chemokines, like stromal cell-derived factor-1 (SDF1/CXCL12), are small secreted proteins that signal cells to migrate. Because SDF1 and its receptor CXCR4 play important roles in embryonic development, cancer metastasis, and HIV/AIDS, this chemokine signaling system is the subject of intense study. However, it is not known whether the monomeric or dimeric structure of SDF1 is responsible for signaling in vivo. Previous structural studies portrayed the SDF1 structure as either strictly monomeric in solution or dimeric when crystallized. Here, we report two-dimensional NMR, pulsed-field gradient diffusion and fluorescence polarization measurements at various SDF1 concentrations, solution conditions, and pH. These results demonstrate that SDF1 can form a dimeric structure in solution, but only at nonacidic pH when stabilizing counterions are present. Thus, while the previous NMR structural studies were performed under acidic conditions that strongly promote the monomeric state, crystallographic studies used nonacidic buffer conditions that included divalent anions shown here to promote dimerization. This pH-sensitive aggregation behavior is explained by a dense cluster of positively charged residues at the SDF1 dimer interface that includes a histidine side chain at its center. A heparin disaccharide shifts the SDF1 monomer-dimer equilibrium in the same manner as other stabilizing anions, suggesting that glycosaminoglycan binding may be coupled to SDF1 dimerization in vivo.  相似文献   

14.
The subunit interface of the homodimeric hemoglobin from Scapharca inaequivalvis, HbI, is stabilized by a network of interactions that involve several hydrogen-bonded structural water molecules, a hydrophobic patch, and a single, symmetrical salt bridge between residues Lys-30 and Asp-89. Upon mutation of Lys-30 to Asp, the interface is destabilized markedly. Sedimentation equilibrium and velocity experiments allowed the estimate of the dimerization constants for the unliganded (K(1,2D) = 8 x 10(4) M(-1)) and for the CO-bound (K(1,2L) = 1 x 10(3) m(-1)) and oxygenated (K(1,2L) = 70 m(-1)) derivatives. For the oxygenated derivative, the destabilization of the subunit interface with respect to native HbI corresponds to about 8 kcal/mol, an unexpectedly high figure. In the K30D mutant, at variance with the native protein, oxygen affinity and cooperativity are strongly dependent on protein concentration. At low protein concentrations (e.g. 1.2 x 10(-5) m heme), at which the monomeric species becomes significant also in the unliganded derivative, oxygen affinity increases and cooperativity decreases. At protein concentrations where both derivatives are dimeric (e.g. 3.3 x 10(-3) m heme), both cooperativity and oxygen affinity decrease. Taken together, the experimental data indicate that in the K30D mutant, the mechanism of cooperativity is drastically altered and is driven by a ligand-linked monomer-dimer equilibrium rather than being based on a direct heme-heme communication as in native HbI.  相似文献   

15.
Chi SW  Jeong DG  Woo JR  Lee HS  Park BC  Kim BY  Erikson RL  Ryu SE  Kim SJ 《FEBS letters》2011,585(4):664-670
Heat shock protein 33 (Hsp33) from Escherichia coli is a redox-regulated molecular chaperone that protects cells from oxidative stress. To understand the molecular basis for the monomer-dimer switch in the functional regulation of E. coli Hsp33, we generated a constitutively monomeric Hsp33 by introducing the Q151E mutation in the dimeric interface and determined its crystal structure. The overall scaffold of the monomeric Hsp33(1-235) (Q151E) mutant is virtually the same as that of the dimeric form, except that there is no domain swapping. The measurement of chaperone activity to thermally denatured luciferase showed that the constitutively monomeric Hsp33 mutant still retains chaperone activity similar to that of wild-type Hsp33(1-235), suggesting that a Hsp33 monomer is sufficient to interact with slowly unfolded substrate.  相似文献   

16.
The monomer-dimer equilibria of the dimeric phospholipases A2 from Crotalus atrox and Agkistrodon piscivorus piscivorus venoms were examined chromatographically as a function of pH and in the presence versus absence of the essential cofactor, calcium ion. At neutral pH without calcium, the subunits of both enzymes reequilibrated sufficiently slowly that dimer and monomer were separated by size exclusion chromatography. At pH 4.2 and lower, the dimers underwent rapid dissociation and reassociation, eluting as single broad peaks whose position as a function of applied protein concentration could be analyzed to determine association constants using an algorithm that estimates these values based on elution positions. Lowering the pH from 7.0 to 4.2 increased the self-association constant of the C. atrox enzyme by 1 order of magnitude and that of the A. p. piscivorus dimer by a factor of 3. Calcium ion, an essential cofactor of phospholipase A2, converted the kinetic behavior of the dimers at neutral pH from slow to virtually instantaneous on the time scale of the chromatography runs, 40 min. Calcium ion also altered the thermodynamic stability of the enzymes; the association constant of A. p. piscivorus phospholipase A2 in neutral pH buffer was reduced by approximately 2 orders of magnitude, whereas that of C. atrox was increased by a factor of 6. The structural basis for the disparate effects of calcium ion on these two acidic, dimeric venom phospholipases A2 is uncertain. This study illustrates the importance of calcium ion and pH on the solution behavior of the dimeric members of this class of enzymes.  相似文献   

17.
The self-association of alpha-chymotrypsin and its di-isopropyl phosphoryl derivative in in I0.03 sodium phophate buffer, pH7,9, was investigated by velocity sedimentation, equilibrium sedimentation and difference gel chromatography. No differences between the native and chemically modified enzyme were observed in the ultracentrifuge studies, and only a marginal (0.6%) difference in weight-average elution volume was detected by difference gel chromatography of 5g/litre solutions on Sephadex G-75. From quantitative analyses of sedimentation velocity and sedimentation-equilibrium distributions obtained with iPr2P (di-isopropylphosphoryl)-chymotrypsin, the polymerizing system is postulated to involve an indefinite association of dimer (with an isodesmic association constant of 0.68 litre/g) that is formed by a discrete dimerization step with equilibrium constant 0.25 litre/g. In addition to providing the best fit of the experimental results, this model of chymotrypsin polymerization at low ionic strength is also consistent with an earlier observation that dimer formation is a symmetrical head-to-head phenomenon under conditions of higher ionic strength (I0.29, pH7.9) where association is restricted to a monomer-dimer equilibrium. It is proposed that the dimerization process is essentially unchanged by variation in ionic strength at pH7.9, and that higher polymers are formed by an entirely different mechanism involving largely electrostatic interactions between dimeric species.  相似文献   

18.
The anthranilate phosphoribosyltransferase from Sulfolobus solfataricus (ssAnPRT) forms a homodimer with a hydrophobic subunit interface. To elucidate the role of oligomerisation for catalytic activity and thermal stability of the enzyme, we loosened the dimer by replacing two apolar interface residues with negatively charged residues (mutations I36E and M47D). The purified double mutant I36E+M47D formed a monomer with wild-type catalytic activity but reduced thermal stability. The single mutants I36E and M47D were present in a monomer-dimer equilibrium with dissociation constants of about 1 μM and 20 μM, respectively, which were calculated from the concentration-dependence of their heat inactivation kinetics. The monomeric form of M47D, which is populated at low subunit concentrations, was as thermolabile as monomeric I36E+M47D. Likewise, the dimeric form of I36E, which was populated at high subunit concentrations, was as thermostable as dimeric wild-type ssAnPRT. These findings show that the increased stability of wild-type ssAnPRT compared to the I36E+M47D double mutant is not caused by the amino acid exchanges per se but by the higher intrinsic stability of the dimer compared to the monomer. In accordance with the negligible effect of the mutations on catalytic activity and stability, the X-ray structure of M47D contains only minor local perturbations at the dimer interface. We conclude that the monomeric double mutant resembles the individual wild-type subunits, and that ssAnPRT is a dimer for stability but not for activity reasons.  相似文献   

19.
1. Six different hemoglobin (Hb) fractions were isolated and characterized from the larvae of Chironomus thummi thummi using column chromatographic procedures. 2. Chromatographic and sedimentation-analytic studies (sedimentation coefficients of 2.0 +/- 0.2 (S)) have shown three Hb fractions to exist basically in a monomeric form. The molecular weight of component M-2 was determined by sedimentation equilibrium technique to be 15,470 +/- 400. The dimeric Hb was found to have sedimentation coefficients of 3.0 +/- 0.1 (S) in the weakly acidic pH region. In alkaline milieu, the reversible dissociation proceeds into the monomeric molecules (S20, W = 1.9 +/- 0.1 (S)). Molecular weights vary between pH 5.7 and 9.8 not only with hydrogen ion concentration, but also with protein concentration in correspondence with a dissociation-association equilibrium consisting of monomers and dimers. 3. For the Hb fraction M-2, a friction ratio of f/fo = 1.03 was calculated, suggesting an almost spherical shape of this protein. In contrast, the dimeric component appears to have a much more asymmetric structure (f/fo = 1.19). 4. The indivdual MetHb fractions bind the ligands: fluoride, imidazole and azide with different affinities.  相似文献   

20.
SDF-1alpha is a member of the chemokine family implicated in various reactions in the immune system. The interaction of SDF-1alpha with its receptor, CXCR4, is responsible for metastasis of a variety of cancers. SDF-1alpha is also known to play a role in HIV-1 pathogenesis. The structures of SDF-1alpha determined by NMR spectroscopy have been shown to be monomeric while X-ray structures are dimeric. Biochemical data and in vivo studies suggest that dimerization is likely to be important for the function of chemokines. We report here the dynamics of SDF-1alpha determined through measurement of main chain (15)N NMR relaxation data. The data were obtained at several concentrations of SDF-1alpha and used to determine a dimerization constant of approximately 5 mM for a monomer-dimer equilibrium. The dimerization constant was subsequently used to extrapolate values for the relaxation data corresponding to monomeric SDF-1alpha. The experimental relaxation data and the extrapolated data for monomeric SDF-1alpha were analyzed using the model free approach. The model free analysis indicated that SDF-1alpha is rigid on the nano- to picosecond timescale with flexible termini. Several residues involved in the dimer interface display slow micro- to millisecond timescale motions attributable to chemical exchange such as monomer-dimer equilibrium. NMR relaxation measurements are shown to be applicable for studying oligomerization processes such as the dimerization of SDF-1alpha.  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