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.
2.
Arc repressor is tetrameric when bound to operator DNA   总被引:10,自引:0,他引:10  
B M Brown  J U Bowie  R T Sauer 《Biochemistry》1990,29(51):11189-11195
The Arc repressor of bacteriophage P22 is a member of a family of DNA-binding proteins that use N-terminal residues in a beta-sheet conformation for operator recognition. Here, Arc is shown to bind to its operator site as a tetramer. When mixtures of Arc (53 residues) and an active variant of Arc (78 residues) are used in gel retardation experiments, five discrete protein-DNA complexes are observed. This result is as expected for operators bearing heterotetramers containing 4:0, 3:1, 2:2, 1:3, and 0:4 ratios of the two proteins. Direct measurements of binding stoichiometry support the conclusion that Arc binds to a single 21-base-pair operator site as a tetramer. The Arc-operator binding reaction is highly cooperative (Hill constant = 3.5) and involves at least two coupled equilibria. In the first reaction, two unfolded monomers interact to form a folded dimer (Bowie & Sauer, 1989a). Rapid dilution experiments indicate that the Arc dimer is the kinetically significant DNA-binding species and allow an estimate of the equilibrium dissociation constant for dimerization [K1 = 5 (+/- 3) x 10(-9) M]. The rate of association of Arc-operator complexes shows the expected second-order dependence on the concentration of free Arc dimers, with k2 = 2.8 (+/- 0.7) x 10(18) M-2 s-1. The dissociation of Arc-operator complexes is a first-order process with k-2 = 1.6 (+/- 0.6) x 10(-4) s-1. The ratio of these kinetic constants [K2 = 5.7 (+/- 2.3) x 10(-23) M2] provides an estimate for the equilibrium constant for dissociation of the DNA-bound tetramer to two free Arc dimers and the operator. An independent determination of this complex equilibrium constant [K2 = 7.8 (+/- 4.8) x 10(-23) M2] was obtained from equilibrium binding experiments.  相似文献   

3.
Reactivation of tetrameric porcine skeletal muscle lactic dehydrogenase after dissociation and extensive unfolding of the monomers by 6 M guanidine hydrochloride (Gdn . HCl) is characterized by sigmoidal kinetics, indicating a complex mechanism involving rate-limiting folding and association steps. For analysis of the association reactions, chemical cross-linking with glutaraldehyde may be used [Hermann, R., Jaenicke, R., & Rudolph, R. (1981) Biochemistry 20, 2195-2201]. The data clearly show that the formation of a dimeric intermediate is determined by a first-order folding reaction of the monomers with k1 = (8.0 +/- 0.1) x 10(-4) s-1. The rate constant of the association of dimers to tetramers which represents the second rate-limiting step on the pathway of reconstitution after guanidine denaturation, was then determined by reactivation and cross-linking experiments after dissociation in 0.1 M H3PO4 containing 1 M Na2SO4. The rate constant for the dimer association (which is the only rate-limiting step after acid dissociation) was k2 = (3.0 +/- 0.5) x 10(4) M-1 s-1. On the basis of the given two rate constants, the complete reassociation pattern of porcine lactic dehydrogenase after dissociation and denaturation in 6 M Gdn . HCl can be described by the kinetic model (formula: see text).  相似文献   

4.
Naphthyridine dimer composed of two naphthyridine chromophores and a linker connecting them strongly, and selectively, binds to the guanine-guanine mismatch in duplex DNA. The kinetics for the binding of the G-G mismatch to the naphthyridine dimer was investigated by surface plasmon resonance assay. The sensor surface was prepared by immobilizing naphthyridine dimer through a long poly(ethylene oxide) linker with the ligand density of 9.1 x 10(-12) fmolnm(-2). The kinetic analyses revealed that the binding of the G-G mismatch was sequence dependent on the flanking base pairs, and the G-G mismatches flanking at least one G-C base pair bound to the surface via a two-step process with a 1:1 DNA-ligand stoichiometry. The first association rate constant for the binding of the G-G mismatch in the 5'-CGG-3'/3'-GGC-5' sequence to the naphthyridine dimer-immobilized sensor surface was 3.2 x 10(3)M(-1)s(-1) and the first dissociation rate constant was 1.4 x 10(-2)s(-1). The association and dissociation rate constants for the second step were insensitive to the flanking sequences, and were almost of the same order of magnitude as the first dissociation rate constant. This indicates that the second step had only a small energetic contribution to the binding. The association constant calculated from kinetic parameters was 2.7 x 10(5)M(-1), which is significantly smaller than the apparent association constants obtained from experiments in solution. Electrospray ionization time-of-flight (ESI-TOF) mass spectrometry on the complex produced from the G-G mismatch and naphthyridine dimer showed the formation of the 1:1 complex and a 1:2 DNA-ligand complex in solution. The latter complex became the dominant complex when a six-fold excess of naphthyridine dimer was added to DNA.  相似文献   

5.
1. The hemocyanin of the bivalve, Yoldia limatula (Say) was found by light-scattering to have a mol. wt of 8.0 +/- 0.6 x 10(6). Mass measurements by scanning transmission electron microscopy (STEM) gave a particle mass of 8.25 +/- 0.42 x 10(6) for the native particle and 4.09 +/- 0.20 x 10(6) for the half-molecule. 2. The hemocyanin subunits fully dissociated in 8.0 M urea and 6.0 M GdmCl at pH 8.0, and at pH 11.0, 0.01 M EDTA have mol. wts of 4.38 x 10(5), 4.22 x 10(5) and 4.71 x 10(5), close to one-twentieth of the parent molecular weight of Y. limatula hemocyanin and most gastropod hemocyanins. 3. Analyses of the urea dissociation transitions studied at pH 8.0, 1 x 10(-2) M Mg2+, 1 x 10(-2) M Ca2+ and pH 8.0, 3 x 10(-3) M Ca2+ suggest few hydrophobic amino acid groups, of the order of 10 to 15 at the contact areas of each half-molecule or decamer. 4. The further dissociation of the decamers to dimers and the dimers to monomers indicates the presence of a larger number of amino acid groups of ca 35-40/dimer and 100-120/monomer. 5. This suggests hydrophobic stabilization of the dimer to dimer and monomer to monomer contacts within the decamers, as observed with other molluscan hemocyanins.  相似文献   

6.
M J Chen  K H Mayo 《Biochemistry》1991,30(26):6402-6411
Platelet factor 4 (PF4) monomers (7800 daltons) form dimers and tetramers in varying molar ratios under certain solution conditions [Mayo, K. H., & Chen, M. J. (1989) Biochemistry 28, 9469]. The presence of a simplified aromatic region (one Tyr and two His) and resolved monomer, dimer, and tetramer Y60 3,5 ring proton resonances makes study of PF4 aggregate association/dissociation thermodynamics and kinetics possible. PF4 protein subunit association/dissociation equilibrium thermodynamic parameters have been derived by 1H NMR (500MHz) resonance line-fitting analysis of steady-state Y60 3,5 ring proton resonance monomer-dimer-tetramer populations as a function of temperature from 10 to 40 degrees C. Below 10 degrees C and above 40 degrees C, resonance broadening and overlap severely impaired analysis. Enthalpic and entropic contributions to dimer association Gibb's free energy [-5.1 kcal/mol (30 degrees C)] are +2.5 +/- 1 kcal/mol and +26 +/- 7 eu, respectively, and for tetramer association Gibb's free energy [-5.7 kcal/mol (30 degrees C)], they are -7.5 +/- 1 kcal/mol and -7 +/- 3 eu, respectively. These thermodynamic parameters are consistent with low dielectric medium electrostatic/hydrophobic interactions governing dimer formation and hydrogen bonding governing tetramer formation. Association/dissociation kinetic parameters, i.e., steady-state jump rates, have been derived from exchange-induced line-width increases and from 1H NMR (500 MHz) saturation-transfer and spin-lattice (Tl) relaxation experiments. From dissociation jump rates and equilibrium constants, association rate constants were estimated. For dimer and tetramer equilibria at 30 degrees C, unimolecular dissociation rate constants are 35 +/- 10 s-1 for dimer dissociation and 6 +/- 2 s-1 for tetramer dissociation.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
The formation of deoxyhemoglobin was examined by measuring the heme spectral change that accompanies the aggregation of isolated alpha and beta chains. At low hemeconcentrations (less than 10(-5) M), tetramer formation can be described by two consecutive, second order reactions representing the aggregation of monomers followed by the association of alphabeta dimers. At neutral pH, the rates of monomer and dimer aggregation are roughly the same, approximately 5 X 10(5) M(-1) X(-1) at 20 degrees. Raising or lowering the pH results in a uniform decrease of both aggregation rates due presumably to repulsion of positively charged subunits at acid pH and repulsion of negatively charged subunits at alkaline pH. Addition of p-hydroxymercuribenzoate to alpha chains lowers the rate of monomer aggregation whereas addition of mercurials to the beta subunits appears to lower both the rate of monomer and the rate of dimer aggregation. At high heme concentrations (greater than 10(-5) M) or in the presence of organic phosphates, the rate of chain aggregation becomes limited, in part, by the slow dissociation of beta chain tetramers. In the case of inositol hexaphosphate, the rate of hemoglobin formation exhibits a bell-shaped dependence on phosphate concentration. When intermediate concentrations of inositol hexaphosphate (approximately 10(-4 M) are preincubated with beta subunits, a slow first order time course is observed and exhibits a half-time of about 8 min. As more inositol hexaphosphate is added, the chain aggregation reaction begins to occur more rapidly. Eventually at about 10(-2) M inositol hexaphospate, the time course becomes almost identical to that observed in the absence of phosphates. The increase in the velocity of the chain aggregation reaction at high phosphate concentrations suggests strongly that inositol hexaphosphate binds to beta monomers and, if added in sufficiently large amounts, promotes beta4 dissociation. A quantitative analysis of these results showed that the affinity of beta monomers for inositol hexaphosphate is the same as that of alphabeta dimers. Only when tetramers are formed, either alpha2beta2 or beta4, is a marked increase in affinity for inositol hexaphosphate observed.  相似文献   

8.
Coiled coils consist of two or more amphipathic a-helices wrapped around each other to form a superhelical structure stabilized at the interhelical interface by hydrophobic residues spaced in a repeating 3-4 sequence pattern. Dimeric coiled coils have been shown to often form in a single step reaction in which association and folding of peptide chains are tightly coupled. Here, we ask whether such a simple folding mechanism may also apply to the formation of a three-stranded coiled coil. The designed 29-residue peptide LZ16A was shown previously to be in a concentration-dependent equilibrium between unfolded monomer (M), folded dimer (D), and folded trimer (T). We show by time-resolved fluorescence change experiments that folding of LZ16A to D and T can be described by 2M (k1)<==>(k(-1)) D and M + D (k2)<==>(k(-2)) T. The following rate constants were determined (25 degrees C, pH 7): k1 = 7.8 x 10(4) M(-1) s(-1), k(-1) = 0.015 s(-1), k2 = 6.5 x 10(5) M(-1) s(-1), and k(-2) = 1.1 s(-1). In a separate experiment, equilibrium binding constants were determined from the change with concentration of the far-ultraviolet circular dichroism spectrum of LZ16A and were in good agreement with the kinetic rate constants according to K(D) = k1/2k(-1) and K(T) = k2/k(-2). Furthermore, pulsed hydrogen-exchange experiments indicated that only unfolded M and folded D and T were significantly populated during folding. The results are compatible with a two-step reaction in which a subpopulation of association competent (e.g., partly helical) monomers associate to dimeric and trimeric coiled coils.  相似文献   

9.
Kinetic studies on Cro repressor-operator DNA interaction   总被引:14,自引:0,他引:14  
The six operators of phage lambda and their consensus sequence were synthesized as 21 base-pair DNAs and their interactions with Cro repressor were studied using a filter binding assay. The measured equilibrium dissociation constants suggest that Cro has the highest affinity to the consensus operator (KD = 1.2 X 10(-12) M) and then the OR3 operator (KD = 2.0 X 10(-12) M), after that the affinity becomes lower in the following order: OR1, OL1, OL2, OL3, OR2. The competition experiments show that Cro forms the most stable complex with the consensus operator (t1/2 = 150 min), which is followed by the complex with OR3 (t1/2 = 70 min), OR1, OL1, OL2, OL3 and OR2. The association rate constants (ka) were also measured. They are approximately the same (2 X 10(8) to 4 X 10(8) m-1 s-1) for the consensus, OR3, OR2 and OR1 operators. These experiments have thus shown that the sequence difference in the operator affects the dissociation (KD and kd) but not the association (ka) process. The operators' binding strengths relative to OR1 are 14 (for consensus operator), 7.6 (OR3), 0.73 (OL1), 0.42 (OL2), 0.16 (OL3) and 0.1 (OR2). Seven different lengths of OR-containing DNA fragments were prepared. Measurement of kinetic parameters shows that the affinity of Cro to operator DNA (measured by KD) is essentially constant and independent of the DNA length, while the association and dissociation rate constants increase as the DNA length increases. This is consistent with the idea that Cro locates and leaves its operator via a two-step mechanism. It appears that Cro binds first at an arbitrary site on DNA, then is transferred to its operator site by a facilitated mechanism. The process is reversed when Cro dissociates from the operator. Most of our data fit to the theoretical expression formulated by Berg, Winter & von Hippel for the sliding mechanism. We conclude that Cro slides along the DNA to locate and leave the operator.  相似文献   

10.
As an effort to elucidate the quaternary structure of cyclomaltodextrinase I-5 (CDase I-5) as a function of pH and salt concentration, the dissociation/association processes of the enzyme were investigated under various pH and salt conditions. Previous crystallographic analysis of CDase I-5 indicated that it existed exclusively as a dodecamer at pH 7.0, forming an assembly of six 3D domain-swapped dimeric subunits. In the present study, analytical ultracentrifugation analysis suggested that CDase I-5 was present as a dimer in the pH range of 5.0-6.0, while the dodecameric form was predominant at pH values above 6.5. No dissociation of the dodecamer was observed at pH 7.0 and the above. Gel filtration chromatography showed that CDase I-5 dissociated into dimers at a rate of 8.58 x 10(-2) h(-1) at pH 6.0. A mutant enzyme with three histidine residues (H49, H89, and H539) substituted with valines dissociated into dimers faster than the wild-type enzyme at both pH 6.0 and 7.0. The tertiary structure indicated that the effect of pH on dissociation of the oligomer was mainly due to the protonation of H539. Unlike the pH-dependent process, the dissociation of wild-type CDase I-5 proceeded very fast at pH 7.0 in the presence of 0.2-1.0 M of KCl. Stopped-flow spectrophotometric analysis at various concentrations of KCl showed that the rate constants of dissociation (kd) from dodecamers into dimers were 5.96 s(-1) and 7.99 s(-1) in the presence of 0.2 M and 1.0 M of KCl, respectively.  相似文献   

11.
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   

12.
Compound (1), which consists of an oxine and a pyridinium group, was synthesized as a metal-responsive DNA binding ligand. Two 1s coordinate to a Cu(II) to form a stable dimer (1(2)-Cu), even in the presence of DNA. The binding of 1 with sonicated calf thymus DNA was enhanced by ca. 10(3) times after forming the dimer; the binding constants were estimated to be 3.2 x 10(4)M(-1) and 2.4 x 10(7)M(-1) in the absence and the presence, respectively, of a half mole of Cu(II). The enormous acceleration of the binding is partly attributed to the generation of a dicationic charge by the formation of the dimer. High cooperativity between dimers could be also responsible; dimers would gather along the duplex as a template to form 1D spiral aggregates.  相似文献   

13.
c-Myc, the protein product of protooncogene c-myc, functions in cell proliferation, differentiation, and neoplastic disease. In this study, recombinant c-Myc and Max proteins, encompassing DNA binding (basic region) and dimerization (helix-loop-helix/leucine zipper) domain of human origin, were expressed in bacteria as Myc87 and Max85. Myc87 was purified under denatured conditions and was renatured again. The dissociation constant for the protein dimers and for dimer/DNA complexes were not detectable by isothermal titration calorimetry because of the low degree of solubility of Myc87 and Max85. Therefore, we set up equations which were used to determine the dissociation constants from the proportion of protein-DNA complexes. The dimer dissociation constants in TBS were 5.90(+/-0.54)x10(-7)M for Max85/Max85 homodimer, 6.85(+/-0.25)x10(-3)M for Myc87/Myc87 homodimer, and 2.55(+/-0.29)x10(-8)M for Myc87/Max85 heterodimer, and the DNA-binding dissociation constants in TBS were 1.33(+/-0.21)x10(-9)M for Max85/Max85/DNA, 2.27(+/-0.08)x10(-12)M for Myc87/Myc87/DNA, and 4.43(+/-0.37)x10(-10)M for Myc87/Max85/DNA. In addition, we revealed that linoleic acid which is known as an inhibitor for the formation of Max/Max/DNA complex reduced the affinity of Max homodimer for DNA. This result indicates that linoleic acid may bind to the DNA-binding region of Max homodimer.  相似文献   

14.
15.
16.
Dimer-tetramer and monomer-dimer-tetramer equilibria of tetrameric hemoglobins and their single chains in the CO form, respectively, were evaluated using the microbatch multiangle light-scattering (MALS) analysis system. The molecular weights of human Hb A and Hb F in the CO form were dependent on concentration. The dissociation constants to dimers of Hb A and Hb F were 2.58 x 10(-6) and 0.66 x 10(-6), respectively. Equilibration of single globin chains, including alpha, beta, and gamma chains, was also evaluated by the same method. The dissociation constants of alpha-chain dimers to monomers, of beta-chain tetramers to monomers, and of gamma-chain tetramers to dimers were 14 x 10(-6), 25 x 10(-17), and 6.86 x 10(-6) M, respectively. These results indicate that the MALS analysis system can not only determine molecular weight but also characterize protein-protein interactions of multi-subunit proteins.  相似文献   

17.
The finding that exchange of tubulin subunits between tubulin dimers (alpha-beta + alpha'beta' <--> alpha'beta + alphabeta') does not occur in the absence of protein cofactors and GTP hydrolysis conflicts with the assumption that pure tubulin dimer and monomer are in rapid equilibrium. This assumption underlies the many physical chemical measurements of the K(d) for dimer dissociation. To resolve this discrepancy we used surface plasmon resonance to determine the rate constant for dimer dissociation. The half-time for dissociation was approximately 9.6 h with tubulin-GTP, 2.4 h with tubulin-GDP, and 1.3 h in the absence of nucleotide. A Kd equal to 10(-11) M was calculated from the measured rate for dissociation and an estimated rate for association. Dimer dissociation was found to be reversible, and dimer formation does not require GTP hydrolysis or folding information from protein cofactors, because 0.2 microM tubulin-GDP incubated for 20 h was eluted as dimer when analyzed by size exclusion chromatography. Because 20 h corresponds to eight half-times for dissociation, only monomer would be present if dissociation were an irreversible reaction and if dimer formation required GTP or protein cofactors. Additional evidence for a 10(-11) M K(d) was obtained from gel exclusion chromatography studies of 0.02-2 nM tubulin-GDP. The slow dissociation of the tubulin dimer suggests that protein tubulin cofactors function to catalyze dimer dissociation, rather than dimer assembly. Assuming N-site-GTP dissociation is from monomer, our results agree with the 16-h half-time for N-site GTP in vitro and 33 h half-life for tubulin N-site-GTP in CHO cells.  相似文献   

18.
The enzyme complex prothrombinase plays a pivotal role in fibrin clot development through the production of thrombin, making this enzyme complex an attractive target for therapeutic regulation. This study both functionally and structurally characterizes a potent, highly selective, active site directed inhibitor of human factor Xa and prothrombinase, PD0313052, and identifies structurally conserved residues in factor Xa and prothrombinase. Analyses of the association and dissociation of PD0313052 with human factor Xa identified a reversible, slow-onset mechanism of inhibition and a simple, single-step bimolecular association between factor Xa and PD0313052. This interaction was governed by association (k(on)) and dissociation (k(off)) rate constants of (1.0 +/- 0.1) x 10(7) M(-1) s(-1) and (1.9 +/- 0.5) x 10(-3) s(-1), respectively. The inhibition of human factor Xa by PD0313052 displayed significant tight-binding character described by a Ki* = 0.29 +/- 0.08 nM. Similar analyses of the inhibition of human prothrombinase by PD0313052 also identified a slow-onset mechanism with a Ki* = 0.17 +/- 0.03 nM and a k(on) and k(off) of (0.7 +/- 0.1) x 10(7) M(-1) s(-1) and (1.7 +/- 0.8) x 10(-3) s(-1), respectively. Crystals of factor Xa and PD0313052 demonstrated hydrogen bonding contacts within the S1-S4 pocket at residues Ser195, Asp189, Gly219, and Gly216, as well as interactions with aromatic residues within the S4 pocket. Overall, these data demonstrate that the inhibition of human factor Xa by PD0313052 occurs via a slow, tight-binding mechanism and indicate that active site residues of human factor Xa, including the catalytic Ser195, are effectively unaltered following assembly into prothrombinase.  相似文献   

19.
The foldon domain constitutes the C-terminal 30 amino acid residues of the trimeric protein fibritin from bacteriophage T4. Its function is to promote folding and trimerization of fibritin. We investigated structure, stability and folding mechanism of the isolated foldon domain. The domain folds into the same trimeric beta-propeller structure as in fibritin and undergoes a two-state unfolding transition from folded trimer to unfolded monomers. The folding kinetics involve several consecutive reactions. Structure formation in the region of the single beta-hairpin of each monomer occurs on the submillisecond timescale. This reaction is followed by two consecutive association steps with rate constants of 1.9(+/-0.5)x10(6)M(-1)s(-1) and 5.4(+/-0.3)x10(6)M(-1)s(-1) at 0.58 M GdmCl, respectively. This is similar to the fastest reported bimolecular association reactions for folding of dimeric proteins. At low concentrations of protein, folding shows apparent third-order kinetics. At high concentrations of protein, the reaction becomes almost independent of protein concentrations with a half-time of about 3 ms, indicating that a first-order folding step from a partially folded trimer to the native protein (k=210 +/- 20 s(-1)) becomes rate-limiting. Our results suggest that all steps on the folding/trimerization pathway of the foldon domain are evolutionarily optimized for rapid and specific initiation of trimer formation during fibritin assembly. The results further show that beta-hairpins allow efficient and rapid protein-protein interactions during folding.  相似文献   

20.
1. The hemocyanin of the chiton, Stenoplax conspicua, has a molecular weight determined by light-scattering of 4.2 X 10(6) daltons, (dt) and a sedimentation coefficient of 60 S. 2. The fully dissociated subunits in 6.0 and 8.0 M urea, and at pH 8.9-10 in the absence of divalent ions, have molecular weights of 4.15-4.30 x 10(5) and 4.17-4.75 x 10(5) dt, which is close to one-tenth of the molecular weight of the parent hemocyanin assembly. 3. The pH dependence of the molecular weights from pH 4.5 to 11 exhibit bell-shaped transition profiles, best accounted for by a three-species, decamer to dimer to monomer scheme of subunit dissociation, with one acidic and one basic ionizing group per dimer and 5-8 acidic and basic groups per monomer. 4. In the absence of stabilizing divalent ions S. conspicua hemocyanin is relatively unstable. At pH 7.4 in the presence of 0.01 M EDTA, it is predominantly in the dimeric state, characterized by a sedimentation constant of 18 S. It is also more readily dissociated to monomers at high pHs (8-9 and above) than are the C. stelleri and A. granulata hemocyanins. 5. Urea and GdmCl are effective dissociating agents of S. conspicua hemocyanin. The urea dissociation profile obtained at pH 8.5, 0.01 M Mg2+, 0.01 M Ca2+, and analyzed by means of the decamer-dimer-monomer scheme of subunit dissociation gave estimates of about 30 amino acid groups (Napp) at the dimer contacts within the hemocyanin decamers and about 120 groups per monomer within each dimer, suggesting hydrophobic stabilization of hemocyanin assembly.  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