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 125 毫秒
1.
S100 proteins constitute a large subfamily of the EF-hand superfamily of calcium binding proteins. They possess one classical EF-hand Ca2+-binding domain and an atypical EF-hand domain. Most of the S100 proteins form stable symmetric homodimers. An analysis of literature data on S100 proteins showed that their physiological concentrations could be much lower than dissociation constants of their dimeric forms. It means that just monomeric forms of these proteins are important for their functioning. In the present work, thermal denaturation of apo-S100P protein monitored by intrinsic tyrosine fluorescence has been studied at various protein concentrations within the region from 0.04–10 μM. A transition from the dimeric to monomeric form results in a decrease in protein thermal stability shifting the mid-transition temperature from 85 to 75 °C. Monomeric S100P immobilized on the surface of a sensor chip of a surface plasmon resonance instrument forms calcium dependent 1 to 1 complexes with human interleukin-11 (equilibrium dissociation constant 1.2 nM). In contrast, immobilized interleukin-11 binds two molecules of dimeric S100P with dissociation constants of 32 nM and 288 nM. Since effective dissociation constant of dimeric S100P protein is very low (0.5 μM as evaluated from our data) the sensitivity of the existing physical methods does not allow carrying out a detailed study of S100P monomer properties. For this reason, we have used molecular dynamics methods to evaluate structural changes in S100P upon its transition from the dimeric to monomeric state. 80-ns molecular dynamics simulations of kinetics of formation of S100P, S100B and S100A11 monomers from the corresponding dimers have been carried out. It was found that during the transition from the homo-dimer to monomer form, the three S100 monomer structures undergo the following changes: (1) the helices in the four-helix bundles within each monomer rotate in order to shield the exposed non-polar residues; (2) almost all lost contacts at the dimer interface are substituted with equivalent and newly formed interactions inside each monomer, and new stabilizing interactions are formed; and (3) all monomers recreate functional hydrophobic cores. The results of the present study show that both dimeric and monomeric forms of S100 proteins can be functional.  相似文献   

2.
The S100 proteins comprise 25 calcium-signalling members of the EF-hand protein family. Unlike typical EF-hand signalling proteins such as calmodulin and troponin-C, the S100 proteins are dimeric, forming both homo- and heterodimers in vivo. One member of this family, S100B, is a homodimeric protein shown to control the assembly of several cytoskeletal proteins and regulate phosphorylation events in a calcium-sensitive manner. Calcium binding to S100B causes a conformational change involving movement of helix III in the second calcium-binding site (EF2) that exposes a hydrophobic surface enabling interactions with other proteins such as tubulin and Ndr kinase. In several S100 proteins, calcium binding also stabilizes dimerization compared to the calcium-free states. In this work, we have examined the guanidine hydrochloride (GuHCl)-induced unfolding of dimeric calcium-free S100B. A series of tryptophan substitutions near the dimer interface and the EF2 calcium-binding site were studied by fluorescence spectroscopy and showed biphasic unfolding curves. The presence of a plateau near 1.5 M GuHCl showed the presence of an intermediate that had a greater exposed hydrophobic surface area compared to the native dimer based on increased 4,4-dianilino-1,1′-binaphthyl-5,5′-disulfonic acid fluorescence. Furthermore, 1H-15N heteronuclear single quantum coherence analyses as a function of GuHCl showed significant chemical shift changes in regions near the EF1 calcium-binding loop and between the linker and C-terminus of helix IV. Together these observations show that calcium-free S100B unfolds via a dimeric intermediate.  相似文献   

3.
Proaerolysin, the proform of the channel-forming protein aerolysin, is secreted as a dimer by Aeromonas sp. The protein also exists as a dimer in the crystal, as well as in solution, at least at concentrations in the region of 500 microg/ml. Recently it has been argued that proaerolysin becomes monomeric at concentrations below 100 microg/ml and that only the monomeric form of the protoxin can bind to cell surface receptors (Fivaz, M., Velluz, M.-C., and van der Goot, F. G. (1999) J. Biol. Chem. 274, 37705-37708). Here we show, using non-denaturing polyacrylamide electrophoresis, chemical cross-linking, and analytical ultracentrifugation, that proaerolysin remains dimeric at the lowest concentrations of the protein that we measured (less than 5 microg/ml) and that the dimeric protoxin is quite capable of receptor binding.  相似文献   

4.
Canonical glutathione (GSH) transferases are dimeric proteins with subunits composed of an N-terminal GSH binding region (domain 1) and a C-terminal helical region (domain 2). The stabilities of several GSH transferase dimers are dependent upon two groups of interactions between domains 1 and 2 of opposing subunits: a hydrophobic ball-and-socket motif and a buried charge cluster motif. In rGSTM1-1, these motifs involve residues F56 and R81, respectively. The structural basis for the effects of mutating F56 to different residues on dimer stability and function has been reported (Codreanu et al. (2005) Biochemistry 44, 10605-10612). Here, we show that the simultaneous disruption of both motifs in the F56S/R81A mutant causes complete dissociation of the dimer to a monomeric protein on the basis of gel filtration chromatography and multiple-angle laser light scattering. The fluorescence and far-UV CD properties of the double mutant as well as the kinetics of amide H/D exchange along the polypeptide backbone suggest that the monomer has a globular structure that is similar to a single subunit in the native protein. However, the mutant monomer has severely impaired catalytic activity, suggesting that the dimer interface is vital for efficient catalysis. Backbone amide H/D exchange kinetics in the F56S and F56S/R81A mutants indicate that a reorganization of the loop structure between helix alpha2 and strand beta3 near the active site is responsible for the decreased catalytic activity of the monomer. In addition, the junction between the alpha4 and alpha5 helices in F56S/R81R shows decreased H/D exchange, indicating another structural change that may affect catalysis. Although the native subunit interface is important for dimer stability, urea-induced unfolding of the F56S/R81A mutant suggests that the interface is not essential for the thermodynamic stability of individual subunits. The H/D exchange data reveal a possible molecular basis for the folding cooperativity observed between domains 1 and 2.  相似文献   

5.
The quaternary structure of bovine seminal ribonuclease, the only dimeric protein in the superfamily of ribonucleases, is maintained both by noncovalent forces and by two intersubunit disulfides. The available monomeric derivatives of the enzyme may not be reassembled into dimers. They are catalytically active, but do not retain certain properties of the dimeric enzyme, such as: (i) the ability to respond cooperatively to increasing substrate concentrations in the rate-limiting reaction step; and (ii) the antitumor and immunosuppressive actions. In this report we describe the preparation of stable monomers of seminal ribonuclease which can be reassociated into covalent dimers indistinguishable from the native protein. With this procedure a hybrid dimer was constructed, made up of a native subunit associated to a subunit catalytically inactivated by selective alkylation of the active site His-119. This dimer was found to have enzymic properties typical of monomeric ribonucleases, such as a hyperbolic saturation curve in the hydrolytic rate-limiting step of the reaction. However, the hybrid dimer was one order-of-magnitude more active than the dimeric enzyme.  相似文献   

6.
The unfolding of triosephosphate isomerase (TIM) from Trypanosoma brucei (TbTIM) induced by guanidine hydrochloride (GdnHCl) was characterized. In contrast to other TIMs, where unfolding is a two or three state process, TbTIM showed two intermediates. The solvent exposure of different regions of the protein in the unfolding process was characterized spectroscopically with mutant proteins in which tryptophans (W) were changed to phenlylalanines (F). The midpoints of the transitions measured by circular dichroism, intrinsic fluorescence, and catalytic activity, as well as the increase in 1-aniline 8-naphthalene sulfonate fluorescence, show that the native state was destabilized in the W12F and W12F/W193F mutants, relative to the wild-type enzyme. Using the hydrodynamic profile for the unfolding of a monomeric TbTIM mutant (RMM0-1TIM) measured by size-exclusion chromatography as a standard, we determined the association state of these intermediates: D*, a partially expanded dimer, and M*, a partially expanded monomeric intermediate. High-molecular-weight aggregates were also detected. At concentrations over 2.0 M GdnHCl, the hydrodynamic properties of TbTIM and RMM0-1TIM are the same, suggesting that the dimeric intermediate dissociates and the unfolding proceeds through the denaturation of an expanded monomeric intermediate. The analysis of the denaturation process of the TbTIM mutants suggests a sequence for the gradual exposure of W residues: initially the expansion of the native dimer to form D* affects the environments of W12 and W159. The dissociation of D* to M* and further unfolding of M* to U induces the exposure of W170. The role of protein concentration in the formation of intermediates and aggregates is discussed considering the irreversibility of this unfolding process.  相似文献   

7.
Low molecular weight protein tyrosine phosphatase (LMW-PTP) dimerizes in the phosphate-bound state in solution with a dissociation constant of K(d)=1.5(+/-0.1)mM and an off-rate on the order of 10(4)s(-1). 1H and 15N NMR chemical shifts identify the dimer interface, which is in excellent agreement with that observed in the crystal structure of the dimeric S19A mutant. Two tyrosine residues of each molecule interact with the active site of the other molecule, implying that the dimer may be taken as a model for a complex between LMW-PTP and a target protein. 15N relaxation rates for the monomeric and dimeric states were extrapolated from relaxation data acquired at four different protein concentrations. Relaxation data of satisfactory precision were extracted for the monomer, enabling model-free analyses of backbone fluctuations on pico- to nanosecond time scales. The dimer relaxation data are of lower quality due to extrapolation errors and the possible presence of higher-order oligomers at higher concentrations. A qualitative comparison of order parameters in the monomeric and apparent dimeric states shows that loops forming the dimer interface become rigidified upon dimerization. Qualitative information on monomer-dimer exchange and intramolecular conformational exchange was obtained from the concentration dependence of auto- and cross-correlated relaxation rates. The loop containing the catalytically important Asp129 fluctuates between different conformations in both the monomeric and dimeric (target bound) states. The exchange rate compares rather well with that of the catalyzed reaction step, supporting existing hypotheses that catalysis and enzyme dynamics may be coupled. The side-chain of Trp49, which is important for substrate specificity, exhibits conformational dynamics in the monomer that are largely quenched upon formation of the dimer, suggesting that binding is associated with the selection of a single side-chain conformer.  相似文献   

8.
Dimeric procaspase-3 unfolds via a four-state equilibrium process.   总被引:2,自引:0,他引:2  
K Bose  A C Clark 《Biochemistry》2001,40(47):14236-14242
We have examined the folding and assembly of a catalytically inactive mutant of procaspase-3, a homodimeric protein that belongs to the caspase family of proteases. The caspase family, and especially caspase-3, is integral to apoptosis. The equilibrium unfolding data demonstrate a plateau between 3 and 5 M urea, consistent with an apparent three-state unfolding process. However, the midpoint of the second transition as well as the amplitude of the plateau are dependent on the protein concentration. Overall, the data are well described by a four-state equilibrium model in which the native dimer undergoes an isomeration to a dimeric intermediate, and the dimeric intermediate dissociates to a monomeric intermediate, which then unfolds. By fitting the four-state model to the experimental data, we have determined the free energy change for the first step of unfolding to be 8.3 +/- 1.3 kcal/mol. The free energy change for the dissociation of the dimeric folding intermediate to two monomeric intermediates is 10.5 +/- 1 kcal/mol. The third step in the unfolding mechanism represents the complete unfolding of the monomeric intermediate, with a free energy change of 7.0 +/- 0.5 kcal/mol. These results show two important points. First, dimerization of procaspase-3 occurs as a result of the association of two monomeric folding intermediates, demonstrating that procaspase-3 dimerization is a folding event. Second, the stability of the dimer contributes significantly to the conformational free energy of the protein (18.8 of 25.8 kcal/mol).  相似文献   

9.
The OpuA system of Bacillus subtilis is a member of the substrate-binding-protein-dependent ABC transporter superfamily and serves for the uptake of the compatible solute glycine betaine under hyperosmotic growth conditions. Here, we have characterized the nucleotide-binding protein (OpuAA) of the B.subtilis OpuA transporter in vitro. OpuAA was overexpressed heterologously in Escherichia coli as a hexahistidine tag fusion protein and purified to homogeneity by affinity and size exclusion chromatography (SEC). Dynamic monomer/dimer equilibrium was observed for OpuAA, and the K(D) value was determined to be 6 microM. Under high ionic strength assay conditions, the monomer/dimer interconversion was diminished, which enabled separation of both species by SEC and separate analysis of both monomeric and dimeric OpuAA. In the presence of 1 M NaCl, monomeric OpuAA showed a basal ATPase activity (K(M)=0.45 mM; k(2)=2.3 min(-1)), whereas dimeric OpuAA showed little ATPase activity under this condition. The addition of nucleotides influenced the monomer/dimer ratio of OpuAA, demonstrating different oligomeric states during its catalytic cycle. The monomer was the preferred species under post-hydrolysis conditions (e.g. ADP/Mg(2+)), whereas the dimer dominated the nucleotide-free and ATP-bound states. The affinity and stoichiometry of monomeric or dimeric OpuAA/ATP complexes were determined by means of the fluorescent ATP-analog TNP-ATP. One molecule of TNP-ATP was bound in the monomeric state and two TNP-ATP molecules were detected in the dimeric state of OpuAA. Binding of TNP-ADP/Mg(2+) to dimeric OpuAA induced a conformational change that led to the decay of the dimer. On the basis of our data, we propose a model that couples changes in the oligomeric state of OpuAA with ATP hydrolysis.  相似文献   

10.
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   

11.
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   

12.
Monomeric and dimeric PufX-containing core complexes have been purified from membranes of wild-type Rhodobacter sphaeroides. Reconstitution of both samples by detergent removal in the presence of lipids leads to the formation of two-dimensional crystals constituted of dimeric core complexes. Two-dimensional crystals were further analyzed by cryoelectron microscopy and atomic force microscopy. A projection map at 26-A resolution reveals that core complexes assemble in an "S"-shaped dimeric complex. Each core complex is composed of one reaction center, 12 light-harvesting 1 alpha/beta-heterodimers, and one PufX protein. The light-harvesting 1 assemblies are open with a gap of density of approximately 30-A width and surround oriented reaction centers. A maximum density is found at the dimer junction. Based on the projection map, a model is proposed, in which the two PufX proteins are located at the dimer junction, consistent with the finding of dimerization of monomeric core complexes upon reconstitution. This localization of PufX in the core complex implies that PufX is the structural key for the dimer complex formation rather than a channel-forming protein for the exchange of ubiquinone/ubiquinol between the reaction center and the cytochrome bc1 complex.  相似文献   

13.
Decorin, the prototypical small leucine-rich proteoglycan, binds to collagen and thereby regulates collagen assembly into fibrils. The crystal structure of the decorin core protein revealed a tight dimer formed by the association of two monomers via their concave faces (Scott, P. G., McEwan, P. A., Dodd, C. M., Bergmann, E. M., Bishop, P. N., and Bella, J. (2004) Proc. Natl. Acad. Sci. U.S.A. 101, 15633–15638). Whether decorin binds collagen as a dimer has been controversial. Using analytical ultracentrifugation, we determined a dissociation constant of 1.37 ± 0.30 μm for the mouse decorin dimer. Dimerization could be abolished by engineering glycosylation sites into the dimer interface; other interface mutants remained dimeric. The monomeric mutants were as stable as wild-type decorin in thermal unfolding experiments. Mutations on the concave face of decorin abolished collagen binding regardless of whether the mutant proteins retained the ability to dimerize or not. We conclude that the concave face of decorin mediates collagen binding and that the dimer therefore must dissociate to bind collagen.  相似文献   

14.
Ferguson PL  Shaw GS 《Biochemistry》2002,41(11):3637-3646
Human S100B(beta beta) is a small intracellular EF-hand calcium-binding protein that consists of two noncovalently associated 91-residue beta monomers. The three-dimensional structures of S100B reveal the dimer interface consists of four alpha-helices (I, I' and IV, IV') packed in an X-type bundle. In this study, guanidine hydrochloride denaturation and dynamic light scattering were used to assess the impact of single (L3A, L3S, M7A, I11A, F14A) and double (L3A/I11A and L3A/F14A) substitution mutations in helix I on the stability and dimerization propensity of S100B. The free energy of unfolding (Delta G(u)) of wild-type apo-S100B was determined to be 72.4 +/- 4.0 kJ mol(-1), consistent with it being the most stable calcium-binding protein to date. The order of stability of the mutants in their apo form is S100B > L3A > L3S > I11A > M7A approximately L3A/I11A > F14A > L3A/F14A. Further, there is a strong correlation between the stability and the cooperativity of unfolding. Each mutation proved to be more stable in its calcium form compared to its apo form. The calcium-bound L3S substitution proved to be significantly more stable than calcium-saturated S100B, whereas the L3A, I11A, and L3A/I11A mutants are only slightly more stable than the wild-type protein. The F14A and L3A/F14A mutants are significantly reduced in stability, even in the presence of calcium.  相似文献   

15.
Mutations at many different sites in the gene encoding human Cu,Zn superoxide dismutase (SOD) are known to be causative agents in amyotrophic lateral sclerosis (ALS). One explanation for the molecular basis of this pathology is the aggregation of marginally soluble, partially structured states whose populations are enhanced in the protein variants. As a benchmark for testing this hypothesis, the equilibrium and kinetic properties of the reversible folding reaction of a metal-free variant of SOD were investigated. Reversibility was achieved by replacing the two non-essential cysteine residues with non-oxidizable analogs, C6A/C111S, to produce apo-AS-SOD. The metal-free pseudo-wild-type protein is folded and dimeric in the absence of chemical denaturants, and its equilibrium folding behavior is well described by an apparent two-state mechanism involving the unfolded monomer and the native dimer. The apparent free energy of folding in the absence of denaturant and at standard state is -20.37(+/- 1.04) kcal (mol dimer)(-1). A global analysis of circular dichroism kinetic traces for both unfolding and refolding reactions, combined with results from small angle X-ray scattering and time-resolved fluorescence anisotropy measurements, supports a sequential mechanism involving the unfolded monomer, a folded monomeric intermediate, and the native dimer. The rate-limiting monomer folding reaction is followed by a near diffusion-limited self-association reaction to form the native dimer. The relative population of the folded monomeric intermediate is predicted not to exceed 0.5% at micromolar concentrations of protein under equilibrium and both strongly unfolding and refolding conditions for metal-free pseudo-wild-type SOD.  相似文献   

16.
Arginine kinase (AK) catalyzes the reversible phosphorylation of arginine by MgATP to form a high-energy compound phosphoarginine (Parg) and MgADP in forward reaction in invertebrates. To detect the different catalytical mechanisms among Stichopus-AK (dimer) and Limulus-AK (monomer) and Torpedo creatine kinase (dimeric CK) and to reveal the structural role of the C-terminal domain loop (C-loop) of dimeric AK, six single-site mutants, E314D, E314Q, E314V, F315A, F315H and F315Y were constructed as well as two multi-site variants, S312R/F315H/V319E (formed by substituting the C-loop of monomeric AK for that of dimeric AK, termed the AAloop) and S312G/E314V/F315D/E317A/S318A/G321S (formed by substituting the C-loop of dimeric CK for that of dimeric AK, termed the ACloop). The AK activity of the three mutants at Glu314 decreased significantly, from 60- to 500-fold. The ACloop showed only slight AK activity, unlike the same construction in Limulus-AK. In addition, all Phe315 mutants including the AAloop which retained Glu314 had modest AK activity (5–84% of the wild type). All the results above suggested that Glu314 played a more significant role in catalysis in dimeric AK than in the monomer. In addition, ANS profiles indicated that the tolerance of the three Glu314 mutants to denaturant decreased slightly compared with wild type AK. Though monomeric AK has a His residue at site 315, mutants F315H and the AAloop could not resist any perturbation of denaturant, and the mutants showed a Gibbs free energy of about 2.7 kJ/mol lower than wild type AK. Therefore Phe315 in dimeric AK has a different role from His315 in monomeric AK. This might contribute to the stabilization of the native conformation, while His315 in Limulus AK directly binded to the carboxylate of arginine. Taking all the results above together, we suggested a unique mechanism in dimeric AK, different from both monomeric AK and dimeric CK.  相似文献   

17.
Marlatt NM  Shaw GS 《Biochemistry》2007,46(25):7478-7487
S100B is a 21 kDa member of the S100 calcium-binding protein family. This protein comprises a symmetric homodimer with each subunit having two EF-hands arranged from four alpha-helices (I-IV). S100B binds calcium and undergoes a conformation change leading to the exposure of hydrophobic surface residues that enable the protein to interact with biological target molecules. The most significant structural change that occurs during calcium binding results in a change in the orientation of helix III with respect to helices II and IV. In this work, the calcium-sensitive conformational change has been studied by utilizing fast 1H-15N HSQC experiments and water-transfer methods to follow the amide exchange in apo-S100B and Ca-S100B at 35 degrees C. In apo-S100B, the protection factors are 2-3 orders of magnitude lower for helix III than for helix I, II, or IV. In addition, the exchange stability measured here for the dimer interface helices (I, I', IV, and IV'), in the absence of calcium, is similar to the stability obtained from chemical denaturation experiments. When calcium binds, significant decreases in the protection factors for helices I and IV indicate a modification in the stability of the dimer interface has occurred. In contrast, helix II protection factors increase slightly, which is consistent with a decreased level of surface exposure of this helix. These data have been compared with those of the monomeric S100 protein, calbindin D9k, to illustrate that upon calcium binding there is a balance maintained between the amide exchange rates in helices II and III, although largely the rates are dissimilar for each of these proteins. This distinguishing feature may be important for the calcium-induced conformational change in S100B, where calcium binding is transmitted to the dimer-forming helices.  相似文献   

18.
The dimeric yeast protein Ure2 shows prion-like behaviour in vivo and forms amyloid fibrils in vitro. A dimeric intermediate is populated transiently during refolding and is apparently stabilized at lower pH, conditions suggested to favour Ure2 fibril formation. Here we present a quantitative analysis of the effect of pH on the thermodynamic stability of Ure2 in Tris and phosphate buffers over a 100-fold protein concentration range. We find that equilibrium denaturation is best described by a three-state model via a dimeric intermediate, even under conditions where the transition appears two-state by multiple structural probes. The free energy for complete unfolding and dissociation of Ure2 is up to 50 kcal mol(-1). Of this, at least 20 kcal mol(-1) is contributed by inter-subunit interactions. Hence the native dimer and dimeric intermediate are significantly more stable than either of their monomeric counterparts. The previously observed kinetic unfolding intermediate is suggested to represent the dissociated native-like monomer. The native state is stabilized with respect to the dimeric intermediate at higher pH and in Tris buffer, without significantly affecting the dissociation equilibrium. The effects of pH, buffer, protein concentration and temperature on the kinetics of amyloid formation were quantified by monitoring thioflavin T fluorescence. The lag time decreases with increasing protein concentration and fibril formation shows pseudo-first order kinetics, consistent with a nucleated assembly mechanism. In Tris buffer the lag time is increased, suggesting that stabilization of the native state disfavours amyloid nucleation.  相似文献   

19.
P D Jeffrey 《Biochemistry》1979,18(12):2508-2513
Examination and measurement of electron micrographs of negatively stained hemocyanin molecules from Cherax destructor show that the predominant aggregated forms, the 16S and 24S components, are typical structures for arthropod hexamers and dodecamers, respectively. In Cherax hemocyanin the hexamers are formed from the monomeric (Mr congruent to 75,000) subunits, M1 and M2, while the dodecamers contain in addition a dimeric (Mr congruent to 150,000) subunit, M3'. Studies of the composition of solutions of the subunits M1 and m2 to which calcium ions have been added at pH 7.8 show that, under these conditions, reassembly occurs to particles indistinguishable from native hexamers. It is noteworthy that dodecamers are not seen since this confirms the previous suggestion that incorporation of the dimeric subunit in the assembly process is necessary for their formation. The results obtained from Cherax hemocyanin are related to those of previous structural studies of arthropod hemocyanins. In particular, the possible controlling role of certain specific subunits in arthropod hemocyanin oligomers containing more than one kind of subunit is illustrated with a model for the Cherax dodecamer, in which the dimeric subunit is shared between the two halves of the molecule.  相似文献   

20.
Deu E  Kirsch JF 《Biochemistry》2007,46(19):5810-5818
The guanidine hydrochloride (GdnHCl) mediated denaturation pathway for the apo form of homodimeric Escherichia coli aspartate aminotransferase (eAATase) (molecular mass = 43.5 kDa/monomer) includes a partially folded monomeric intermediate, M* [Herold, M., and Kirschner, K. (1990) Biochemistry 29, 1907-1913; Birolo, L., Dal Piaz, F., Pucci, P., and Marino, G. (2002) J. Biol. Chem. 277, 17428-17437]. The present investigation of the urea-mediated denaturation of eAATase finds no evidence for an M* species but uncovers a partially denatured dimeric form, D*, that is unpopulated in GdnHCl. Thus, the unfolding process is a function of the employed denaturant. D* retains less than 50% of the native secondary structure (circular dichroism), conserves significant quaternary and tertiary interactions, and unfolds cooperatively (mD*<==>U = 3.4 +/- 0.3 kcal mol-1 M-1). Therefore, the following equilibria obtain in the denaturation of apo-eAATase: D <==> 2M 2M* <==> 2U in GdnHCl and D <==> D* <==> 2U in urea (D = native dimer, M = folded monomer, and U = unfolded state). The free energy of unfolding of apo-eAATase (D <==> 2U) is 36 +/- 3 kcal mol-1, while that for the D* 2U transition is 24 +/- 2 kcal mol-1, both at 1 M standard state and pH 7.5.  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