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.
Dialysis kinetics measurements have been made to study the effect of ionic strength on the dimerization of lysozyme in acidic solutions that lead to the growth of tetragonal lysozyme crystals. Using glutaraldehyde cross-linked dimers of lysozyme, we have determined that both monomers and dimers can escape from 25,000 molecular weight cutoff dialysis membranes with velocity constants of 5.1 x 10(-7) and 1.0 x 10(-7) s(-1) for the monomer and dimer species, respectively. The flux from 25K MWCO membranes has been measured for lysozyme in pH 4.0 buffered solutions of 1, 3, 4, 5, and 7% NaCl over a wide range of protein concentrations. Assuming that dimerization is the first step in crystallization, a simple monomer to dimer equilibrium was used to model the flux rates. Dimerization constants calculated at low protein concentrations were 265, 750, 1212, and 7879 M(-1) for 3, 4, 5, and 7% NaCl, respectively. These values indicate that dimerization increases with the ionic strength of the solution suggesting that aggregation is moderated by electrostatic interactions. At high protein concentrations and high supersaturation, the dimerization model does not describe the data well. However, the Li model that uses a pathway of monomer <-> dimer <-> tetramer <-> octamer <-> 16-mer fits the measured flux data remarkably well suggesting the presence of higher order aggregates in crystallizing solutions.  相似文献   

2.
Bovine carbonmonoxy hemoglobin investigated with light scattering studies is found to dissociate from its native tetramer structure into dimers and monomers. The values of the hydrated tetramer radius, RT = 32.1 Å, and the fractional dissociation vs pH, have been obtained at different ionic strengths from the autocorrelation function of the scattered light. The results suggest that a relevant contribution to Hb dissociation is due to electrostatic effects and, by means of a model derived by Tanford, it has been possible to predict the behavior of dissociation. Among the findings of this approach, we recall the estimates of the electrostatic energy contributions to Hb dissociation, up to ? GRT, and the predicted charge of tetrameric Hb vs pH, which agrees very well with the experimental data. © 1994 John Wiley & Sons, Inc.  相似文献   

3.
The molecular weights of different aggregational states of phosphoenolpyruvate carboxylase purified from the leaves of Zea mays have been determined by measurement of the molecular diameter using a Malvern dynamic light scattering spectrometer. Using these data to identify the monomer, dimer, tetramer, and larger aggregate(s) the effect of pH and various ligands on the aggregational equilibria of this enzyme have been determined. At neutral pH the enzyme favored the tetrameric form. At both low and high pH the tetramer dissociated, followed by aggregation to a "large" inactive form. The order of dissociation at least at low pH appeared to be two-step: from tetramer to dimers followed by dimer to monomers. The monomers then aggregate to a large aggregate, which is inactive. The presence of EDTA at pH 8 protected the enzyme against both inactivation and large aggregate formation. Dilution of the enzyme at pH 7 at room temperature results in driving the equilibrium from tetramer to dimer. The presence of malate with EDTA stabilizes the dimer as the predominant form at low protein concentrations. The presence of the substrate phosphoenolpyruvate alone and with magnesium and bicarbonate induced formation of the tetramer, and decreased the dissociation constant (Kd) of the tetrameric form. The inhibitor malate, however, induced dissociation of the tetramer as evidenced by an increase in the Kd of the tetramer.  相似文献   

4.
The pathway by which the tetrameric protein transthyretin dissociates   总被引:1,自引:0,他引:1  
Foss TR  Wiseman RL  Kelly JW 《Biochemistry》2005,44(47):15525-15533
The homotetrameric protein transthyretin (TTR) must undergo rate-limiting dissociation to its constituent monomers in order to enable partial denaturation that allows the process of amyloidogenesis associated with human pathology to ensue. The TTR quaternary structure contains two distinct dimer interfaces, one of which creates the two binding sites for the natural ligand thyroxine. Tetramer dissociation could proceed through three distinct pathways; scission into dimers along either of the two unique quaternary interfaces followed by dimer dissociation represents two possibilities. Alternatively, the tetramer could lose monomers sequentially. To elucidate the TTR dissociation pathway, we employed two different TTR constructs, each featuring covalent attachment of proximal subunits. We demonstrate that tethering the A and B subunits of TTR with a disulfide bond (as well as the symmetrically disposed C and D subunits) allows urea-mediated dissociation of the resulting (TTR-S-S-TTR)(2) construct, affording (TTR-S-S-TTR)(1) retaining a stable 16-stranded beta-sheet structure that is equivalent to the dimer not possessing a thyroid binding site. In contrast, linking the A and C subunits employing a peptide tether (TTR-L-TTR)(2) affords a kinetically stable quaternary structure that does not dissociate or denature in urea. Both tethered constructs and wild-type TTR exhibit analogous stability based on guanidine hydrochloride denaturation curves. The latter denaturant can denature the tetramer, unlike urea, which can only denature monomeric TTR; hence urea requires dissociation to monomers to function. Under native conditions, the (TTR-S-S-TTR)(2) construct is able to dissociate and incorporate subunits from labeled WT TTR homotetramers at a rate equivalent to that exhibited by WT TTR. In contrast, the (TTR-L-TTR)(2) construct is unable to exchange any subunits, even after 180 h. All of the data presented herein and elsewhere demonstrate that the pathway of TTR tetramer dissociation occurs by scission of the tetramer along the crystallographic C(2) axis affording AB and CD dimers that rapidly dissociate into monomers. Determination of the mechanism of dissociation provides an explanation for why small molecules that bind at the AB/CD dimer-dimer interface impose kinetic stabilization upon TTR and disease-associated variants thereof.  相似文献   

5.
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   

6.
Cyanide binding to Chromatium vinosum ferricytochrome c′ has been studied to further investigate possible allosteric interactions between the subunits of this dimeric protein. Cyanide binding to C. vinosum cytochrome c′ appears to be cooperative. However, the cyanide binding reaction is unusual in that the overall affinity of cyanide increases as the concentration of cytochrome c′ decreases and that cyanide binding causes the ligated dimer to dissociate to monomers as shown by gel-filtration chromatography. Therefore, the cyanide binding properties of C. vinosum ferricytochrome c′ are complicated by a cyanide-linked dimer to monomer dissociation equilibrium of the complexed protein. The dimer to monomer dissociation constant is 20-fold smaller than that for CO linked dissociation constant of ferrocytochrome c′. Furthermore, the pH dependence of both the intrinsic equilibrium binding constant and the dimer to monomer equilibrium dissociation constant was investigated over the pH range of 7.0 to 9.2 to examine the effect of any ionizable groups. The equilibrium constants did not exhibit a significant pH dependence over this pH range.  相似文献   

7.
G R Parr  G G Hammes 《Biochemistry》1976,15(4):857-862
The kinetics of dissociation and reassembly of rabbit skeletal muscle phosphofructokinase has been studied using fluorescence, stopped-flow fluorescence and enzyme activity measurements. The dissociation of the fully active tetramer in 0.8 M guanidine hydrochloride (0.1 M potassium phosphate, pH 8.0) occurs in three kinetic phases as measured by changes in the protein fluorescence emission intensity: dissociation of tetramer to dimer with a relaxation time of a few milliseconds; dissociation of dimer to monomer with a relaxation time of a few seconds; and a conformational change of the monomer with a relaxation time of a few minutes. All three phases exhibit first-order kinetics; ATP (0.05 mM) retards the second step but does not influence the rate of the other two processes. The rate of the second process increases with decreasing temperature; this may be due to the involvement of hydrophobic interactions in the stabilization of the dimeric enzyme. A further unfolding of the monomer polypeptide chain occurs at higher guanidine concentrations, and the relaxation time associated with this process was found to be 83 ms in 2.5 M guanidine, 0.1 M potassium phosphate (pH 8.0) at 23 degrees C. The phosphofructokinase monomers were reassembled from 0.8 M guanidine chloride by 1:10 dilution of the guanidine hydrochloride concentration and yielded a protein with 70-94% of the original activity, depending on the protein concentration. The reactivation process follows second-order kinetics; ATP (5 mM) increases the rate of reactivation without altering the reaction order, while fructose 6-phosphate does not influence the rate of reaction. The rate-determining step is probably the association of monomers to form the dimer.  相似文献   

8.
How p53 binds DNA as a tetramer.   总被引:8,自引:1,他引:7       下载免费PDF全文
K G McLure  P W Lee 《The EMBO journal》1998,17(12):3342-3350
The p53 tumor suppressor protein is a tetramer that binds sequence-specifically to a DNA consensus sequence consisting of two consecutive half-sites, with each half-site being formed by two head-to-head quarter-sites (--><-- --><--). Each p53 subunit binds to one quarter-site, resulting in all four DNA quarter-sites being occupied by one p53 tetramer. The tetramerization domain forms a symmetric dimer of dimers, and two contrasting models have the two DNA-binding domains of each dimer bound to either consecutive or alternating quarter-sites. We show here that the two monomers within a dimer bind to a half-site (two consecutive quarter-sites), but not to separated (alternating) quarter-sites. Tetramers bind similarly, with the two dimers within each tetramer binding to pairs of half-sites. Although one dimer within the tetramer is sufficient for binding to one half-site in DNA, concurrent interaction of the second dimer with a second half-site in DNA drastically enhances binding affinity (at least 50-fold). This cooperative dimer-dimer interaction occurs independently of tetramerization and is a primary mechanism responsible for the stabilization of p53 DNA binding. Based on these findings, we present a model of p53 binding to the consensus sequence, with the tetramer binding DNA as a pair of clamps.  相似文献   

9.
Double mixing stopped-flow experiments have been performed to study the stability of asymmetric hemoglobin (Hb) hybrids, consisting of a deoxy and a liganded dimer. The doubly liganded [deoxy/cyano-met] hybrid (species 21) was reported to have an enhanced stability, with tetramer to dimer dissociation requiring over 100 seconds, based on a method that required an incubation of over two days. However, kinetic experiments revealed rapid ligand binding to species 21, as for triply liganded tetramers, which dissociate within a few seconds. For the present study, [deoxy dimer/azido-met dimer] hybrids are formed within 200 ms by stopped-flow mixing of dithionite with a solution containing oxyHb and azido-metHb. The dithionite scavenges oxygen, thus transforming oxyHb to deoxyHb, and the [oxy dimer/azido-met dimer] hybrid to the asymmetric [deoxy/azido-met] hybrid (species 21). After a variable aging time of the asymmetric hybrids, their allosteric state is probed by CO binding in a second mixing. As previously observed the freshly produced asymmetric hybrids bind CO rapidly as for R-state Hb. As the hybrids are aged from 0.1 to 10 seconds, the fraction of slow CO binding increases, consistent with a dissociation of the asymmetric hybrid to form the more stable deoxy Hb tetramer which reacts slowly with CO. Control experiments showed a predominantly slow phase for deoxy Hb, and fast rebinding for the symmetric hybrids.The kinetic data can be simulated with a tetramer to dimer dissociation rate for species 21 of 1.5/second at 100 mM NaCl (pH 7.2) and 1.9/second at 180 mM NaCl (pH 7.4). These values are similar to those reported for liganded Hb, as opposed to deoxy (T-state) tetramers which dissociate over four orders of magnitude more slowly. As expected from simulations of dimer exchange, the observed transition rate depends on the initial fractions of oxy- and metHb; this effect is not consistent with a slow R to T transition. These results, showing a lifetime of about one second for species 21, do not support the symmetry rule which is based on an enhanced stability of the asymmetric hybrid.  相似文献   

10.
The number of artificial protein supramolecules has been increasing; however, control of protein oligomer formation remains challenging. Cytochrome c′ from Allochromatium vinosum (AVCP) is a homodimeric protein in its native form, where its protomer exhibits a four‐helix bundle structure containing a covalently bound five‐coordinate heme as a gas binding site. AVCP exhibits a unique reversible dimer–monomer transition according to the absence and presence of CO. Herein, domain‐swapped dimeric AVCP was constructed and utilized to form a tetramer and high‐order oligomers. The X‐ray crystal structure of oxidized tetrameric AVCP consisted of two monomer subunits and one domain‐swapped dimer subunit, which exchanged the region containing helices αA and αB between protomers. The active site structures of the domain‐swapped dimer subunit and monomer subunits in the tetramer were similar to those of the monomer subunits in the native dimer. The subunit–subunit interactions at the interfaces of the domain‐swapped dimer and monomer subunits in the tetramer were also similar to the subunit–subunit interaction in the native dimer. Reduced tetrameric AVCP dissociated to a domain‐swapped dimer and two monomers upon CO binding. Without monomers, the domain‐swapped dimers formed tetramers, hexamers, and higher‐order oligomers in the absence of CO, whereas the oligomers dissociated to domain‐swapped dimers in the presence of CO, demonstrating that the domain‐swapped dimer maintains the CO‐induced subunit dissociation behavior of native ACVP. These results suggest that protein oligomer formation may be controlled by utilizing domain swapping for a dimer–monomer transition protein.  相似文献   

11.
A theoretical study of aldolase tetramer stability, conducted by finite difference Poisson-Boltzmann (FDPB) and modified Tanford-Kirkwood (MTK) techniques using the atomic coordinates of human aldolase, is described. A method for calculating the interaction energy between subunits is proposed. An analysis of the contribution of different energy terms to the stability and oligomeric equilibria (monomer ⇔ dimer ⇔ tetramer) of aldolase is made. It is shown that the loss of solvation energy and electrostatic interactions at very high and low pH-s destabilise the oligomers. These energy terms are compensated over a wide pH range by the stabilization energy due to hydrophobic interactions. It is shown that the aldolase tetramer is energetically more preferable than other oligomers in the pH range from 5 to 11. Subunit-subunit interactions within the tetramer suggest one dimeric form to be the most stable of the possible sub-parts. For this reason the tetramer can be thought of as a “dimer of dimers”. A comparison between our theoretical results and available experimental data shows that the dissociation of the aldolase tetramer below pH 3–4 cooperatively leads to acid denaturation. A second dissociation is predicted to occur at high pH (>12) in addition to the well known acidic dissociation. The analysis suggests that a mutation of His20 or Arg257 to a neutral residue could decrease the pH of the acidic dissociation by approximately 1 pH unit. Received: 16 February 1998 / Revised version: 8 April 1998 / Accepted: 19 April 1998  相似文献   

12.
D A Yphantis  T Arakawa 《Biochemistry》1987,26(17):5422-5427
Recombinant DNA derived human interferon gamma (IFN-gamma) from Escherichia coli was examined by equilibrium ultracentrifugation. Short-column equilibrium experiments at pH 6.9 in 0.1 M ammonium acetate buffer gave a z-average molecular weight of 33,500 +/- 1400 at infinite dilution, corresponding to 1.98 +/- 0.08 times the formula weight. Long- (2.6 mm) column experiments at pH 7.5 in 0.04 M imidazole buffer gave a molecular weight of 33,400 +/- 500. Under the latter conditions IFN-gamma behaves somewhat nonideally, with the departure from ideality accounted for by an effective (Donnan) charge of about 6+. No association of this dimer to form tetramer or higher polymers was observed, with the association constant for formation of tetramer from dimer K24 found to be less than 34 L mol-1. Similarly, no dissociation to monomers was observable, with the dissociation constant to monomer K21 being less than 5 X 10(-8) mol L-1. At pH 3.55 in 0.02 M buffer (acetate plus acetic acid), there was virtually complete dissociation of the dimer to monomer. Extreme nonideality was seen in this low ionic strength system, and the effective charge on the protein was estimated to be about 11+. The reduced molecular weight M(1 -upsilon rho) of the monomer was found to be about 4.09 +/- 0.20 kg mol-1; this corresponds to a molecular weight of 16,410 +/- 820, with the Scatchard definition of components. A small amount of a polymer with a molecular weight of about 0.5 X 10(6) was detected under these conditions.  相似文献   

13.
The cytosolic NADP+-dependent malic enzyme (c-NADP-ME) has a dimer-dimer quaternary structure in which the dimer interface associates more tightly than the tetramer interface. In this study, the urea-induced unfolding process of the c-NADP-ME interface mutants was monitored using fluorescence and circular dichroism spectroscopy, analytical ultracentrifugation and enzyme activities. Here, we demonstrate the differential protein stability between dimer and tetramer interface interactions of human c-NADP-ME. Our data clearly demonstrate that the protein stability of c-NADP-ME is affected predominantly by disruptions at the dimer interface rather than at the tetramer interface. First, during thermal stability experiments, the melting temperatures of the wild-type and tetramer interface mutants are 8–10°C higher than those of the dimer interface mutants. Second, during urea denaturation experiments, the thermodynamic parameters of the wild-type and tetramer interface mutants are almost identical. However, for the dimer interface mutants, the first transition of the urea unfolding curves shift towards a lower urea concentration, and the unfolding intermediate exist at a lower urea concentration. Third, for tetrameric WT c-NADP-ME, the enzyme is first dissociated from a tetramer to dimers before the 2 M urea treatment, and the dimers then dissociated into monomers before the 2.5 M urea treatment. With a dimeric tetramer interface mutant (H142A/D568A), the dimer completely dissociated into monomers after a 2.5 M urea treatment, while for a dimeric dimer interface mutant (H51A/D90A), the dimer completely dissociated into monomers after a 1.5 M urea treatment, indicating that the interactions of c-NADP-ME at the dimer interface are truly stronger than at the tetramer interface. Thus, this study provides a reasonable explanation for why malic enzymes need to assemble as a dimer of dimers.  相似文献   

14.
Among proteins utilized as sweeteners, neoculin and miraculin are taste-modifying proteins that exhibit pH-dependent sweetness. Several experiments on neoculin have shown that His11 of neoculin is responsible for pH dependence. We investigated the molecular mechanism of the pH dependence of neoculin by molecular dynamics (MD) calculations. The MD calculations for the dimeric structures of neoculin and His11 mutants showed no significant structural changes for each monomer at neutral and acidic pH levels. The dimeric structure of neoculin dissociated to form isolated monomers under acidic conditions but was maintained at neutral pH. The dimeric structure of the His11Ala mutant, which is sweet at both neutral and acidic pH, showed dissociation at both pH 3 and 7. The His11 residue is located at the interface of the dimer in close proximity to the Asp91 residue of the other monomer. The MD calculations for His11Phe and His11Tyr mutants demonstrated the stability of the dimeric structures at neutral pH and the dissociation of the dimers to isolated monomers. The dissociation of the dimer caused a flexible backbone at the surface that was different from the dimeric interface at the point where the other monomer interacts to form an oligomeric structure. Further MD calculations on the tetrameric structure of neoculin suggested that the flexible backbone contributed to further dissociation of other monomers under acidic conditions. These results suggest that His11 plays a role in the formation of oligomeric structures at pH 7 and that the isolated monomer of neoculin at acidic pH is responsible for sweetness.  相似文献   

15.
Apoferritin from horse spleen is composed of 24 subunits that undergo partial dissociation after chemical modification with 2,3-dimethylmaleic anhydride (DMMA), yielding dimeric, trimeric, and tetrameric intermediates, stable at pH 8.5 and 0 degrees C. Deacylation at neutral pH and elevated temperature provides a means to initiate reassembly by appropriate shifts of the solvent conditions. In order to monitor the pathway of self-assembly, starting from different intermediates of dissociation, dimers, trimers, and tetramers were isolated and investigated with respect to their capacity to accomplish reassociation. Intrinsic protein fluorescence, gel permeation chromatography, and analytical ultracentrifugation were applied to characterize the intermediate and final stages of association. The assembly of both the dimer and trimer yields greater than 85% of the native tetracosamer; the overall rate, starting from the dimer, exceeds the one starting from the trimer. Under comparable conditions, the tetramer exhibits only partial reassociation via the dimer and monomer; the corresponding dissociation reaction determines the observed slower rate. Significant assembly intermediates are "structured monomers", dimers, trimers, and dodecamers. Polymerization of the dimer via the tetramer, octamer, etc., does not occur on the pathway of assembly. The results confirm the assembly scheme proposed previously on the basis of cross-linking and spectroscopic experiments [Gerl, M., & Jaenicke, R. (1987) Eur. Biophys. J. 15, 103-109]. Comparison of structural models involving the different subunit interactions responsible for the sequential association supports the monomer----dimer----trimer----hexamer----dodecamer----tetracosamer mechanism of apoferritin self-assembly.  相似文献   

16.
A reduction in pH is known to induce the disassociation of the tetrameric form of transthyretin and favor the formation of amyloid fibers. Using continuum electrostatic techniques, we calculate the titration curves and the stability of dimer and tetramer formation of transthyretin as a function of pH. We find that the tetramer and the dimer become less stable than the monomer as the pH is lowered. The free energy difference is 13.8 kcal/mol for dimer formation and 27 kcal/mol for tetramer formation, from the monomers, when the pH is lowered from 7 to 3.9. Similar behavior is observed for both the wild-type and the mutant protein. Certain residues (namely Glu-72, His-88, His-90, Glu-92, and Tyr-116), play an important role in the binding process, as seen by the considerable pK(1/2) change of these residues upon dimer formation.  相似文献   

17.
Tetramerization of the human p53 tumor suppressor protein is required for its biological functions. However, cellular levels of p53 indicate that it exists predominantly in a monomeric state. Since the oligomerization of p53 involves the rate-limiting formation of a primary dimer intermediate, we engineered a covalently linked pair of human p53 tetramerization (p53tet) domains to generate a tandem dimer (p53tetTD) that minimizes the energetic requirements for forming the primary dimer. We demonstrate that p53tetTD self-assembles into an oligomeric structure equivalent to the wild-type p53tet tetramer and exhibits dramatically enhanced oligomeric stability. Specifically, the p53tetTD dimer exhibits an unfolding/dissociation equilibrium constant of 26 fM at 37 degrees C, or a million-fold increase in stability relative to the wild-type p53tet tetramer, and resists subunit exchange with monomeric p53tet. In addition, whereas the wild-type p53tet tetramer undergoes coupled (i.e. two-state) dissociation/unfolding to unfolded monomers, the p53tetTD dimer denatures via an intermediate that is detectable by differential scanning calorimetry but not CD spectroscopy, consistent with a folded p53tetTD monomer that is equivalent to the p53tet primary dimer. Given its oligomeric stability and resistance against hetero-oligomerization, dimerization of p53 constructs incorporating the tetramerization domain may yield functional constructs that may resist exchange with wild-type or mutant forms of p53.  相似文献   

18.
The red cells of the arcid clam Anadara ovalis contain two electrophoretically distinct hemoglobins: Hb Major (Hb Ma) and Hb Minor (Hb Mi). The major component consists of two electrophoretically indistinguishable tetramers each composed of two heterodimers; the minor hemoblogin is a homodimer whose subunits are different from the tetramer. Functionally, Hb Ma has a higher P50, exhibits a concentration dependent oxygen affinity, has significant ligand cooperativity (n = 2.0), lacks a Bohr effect and is unaffected by ATP. HB Mi has a P50 which is lower and independent of hemoglobin concentration, shows appreciable cooperativity (n = 1.4) and exhibits no heterotropic effects. Both Hb Ma and Mi are resistant to dissociation in the presence of 1.0 M NaI, NaCl and guanidine-HCl but dissociate to monomers when converted to the aquamet but not the cyanmet derivative. The dissociation is completely inhibited by mercaptoethanol. The large number of reactive -SH groups (10-13 per tetramer) suggests that the monomerization is mediated by intra-subunit disulfide bridge formation.  相似文献   

19.
20.
Cycling of intracellular pH has recently been shown to play a critical role in ischemia-reperfusion injury. Ischemia-reperfusion also leads to mitochondrial matrix acidification and dysfunction. However, the mechanism by which matrix acidification contributes to mitochondrial dysfunction, oxidative stress, and the resultant cellular injury has not been elucidated. We observe pH-dependent equilibria between monomeric, dimeric, and a previously undescribed tetrameric form of pig heart lipoamide dehydrogenase (LADH), a mitochondrial matrix enzyme. Dynamic light scattering studies of native LADH in aqueous solution indicate that lowering pH favors a shift in average molecular mass from higher oligomeric states to monomer. Sedimentation velocity of LADH entrapped in reverse micelles reveals dimer and tetramer at both pH 5.8 and 7.5, but monomer was observed only at pH 5.8. Enzyme activity measurements in reverse Aerosol OT micelles in octane indicate that LADH dimer and tetramer possess lipoamide dehydrogenase and diaphorase activities at pH 7.5. Upon acidification to pH 5.8 only the LADH monomer is active and only the diaphorase activity is observed. These results indicate a correlation between pH-dependent changes in the LADH reaction specificity and its oligomeric state. The acidification of mitochondrial matrix that occurs during ischemia-reperfusion injury is sufficient to alter the structure and enzymatic specificity of LADH, thereby reducing mitochondrial defenses, increasing oxidative stress, and slowing the recovery of energy metabolism. Matrix acidification may also disrupt the quaternary structure of other mitochondrial protein complexes critical for cellular homeostasis and survival.  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