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 dissociation of allophycocyanin trimers to monomers was examined under a variety of conditions. For alkyl ureas and alcohols the dissociation increased as the straight-chain alkyls increased in length. The effect of branching chains was smaller. Tetrapropylammonium chloride was found to be a very effective agent for trimer dissociation when compared to ureas and alcohols with similar or longer alkyl chains. An explanation for these observations is that the hydrocarbons have an affinity for nonpolar regions in the contact areas between monomers in a trimeric structure. A comparison among several inorganic salts demonstrated that the chaotropic salts (NaSCN greater than NaClO4 much greater than NaNO3 greater than NaBr) fostered increased trimer dissociation, while nonchaotropes (KF, (NH4)2SO4, K phosphate, and NaCl) produced no measurable amounts of monomer. Allophycocyanin dissolved in D2O was much more stable against dissociation than when dissolved in H2O. All the above observations were consistent with hydrophobic forces being the dominant source of trimer stabilization. The equilibrium constant for the dissociation of trimers to monomers was calculated to be about 6 X 10(-16) mol2 liter-2. Calculations were made of the apparent total number of amino acids (40) in the two contact regions on each monomer. An absorption change analogous but not necessarily identical to a conversion of allophycocyanin II to III was noted when (NH4)2SO4 was present. When allophycocyanin's nonexchangeable hydrogens were replaced by deuteriums, it was more readily dissociated to monomers.  相似文献   

2.
The dissociation of porcine erythrocyte catalase [EC 1.11.1.6] into subunits on denaturation with alkali, GuHCl and urea was investigated by following the changes in hydrodynamic properties, absorption and CD spectra in the Soret region and inactivation of the enzyme. It was found that dissociation proceeded in an "all or none" manner from the native tetramer (molecular weight, ca. 250,000) into identical 1/4-sized monomers (molecular weight, ca. 54,000 with alkali, 65,000 with urea and 71,000 with GuHCl) as estimated by ultracentrifugal analyses. On this dissociation, the sedimentation coefficient decreased from about 11S to 5.1 - 3.7S, and absorption spectra in the Soret region decreased to about 40% of the native level and showed a broad band around 365-375 nm and a shoulder around 415-420 nm; these changes were accompanied by complete loss of enzyme activity. The change in enzyme activity correlated well with that of absorption and CD spectra in the Soret region, depending on denaturation time, alkaline pH used and concentration of both denaturants. The reassociated catalase obtained by removing urea by dialysis was characterized by recovery of distinct CD bands in the Soret and near ultraviolet regions, although the partial refolding of alpha-helical conformation occurred without recovery of enzyme activity. These results indicate that the conformational changes and dissociation process of catalase into subunits can be monitored spectrophotometrically in relation to enzyme activity, and that subtle conformations near the heme groups and polypeptide backbone play an important role in maintaining full enzyme activity of the catalase molecule.  相似文献   

3.
The subunit structure, dissociation, and unfolding of the hemoglobin of the earthworm, Lumbricus terrestris, were investigated by light scattering molecular weight methods and changes in optical rotatory dispersion (at 233 nm) and absorption in the Soret region. Urea and the alkylureas, methyl-, ethyl-, propyl-, and butylurea, were employed as the reagents to cause both dissociation and unfolding of the protein. Analysis of the light scattering data suggests that the dissociation patterns as a function of hemoglobin concentration in the various dissociating solvents can be described in quantitative terms, either as an equilibrium mixture consisting of parent duodecamers and hexamers of 3 x 10(6) and 1.5 x 10(6) molecular weight (in 1-3 M urea, 1-2 M methyl- and ethylurea, and 1 M propylurea), as a mixture of hexamers and monomers, the latter with a molecular weight of 250000 (i.e., in 4 M urea), or as a mixture of all three species of duodecamers, hexamers, and monomers, seen in 2 M propylurea. Parallel studies by optical rotation and absorption measurements indicate that there is little or no unfolding of the subunits at urea and alkylurea concentrations where complete dissociation to hexamers and extensive dissociation to monomers can be achieved. Further splitting of the monomers (A subunits) to smaller fragments of one-third to one-quarter of the molecular weight of the monomers (B subunits) is seen in the presence of 7 and 8 M urea (pH 7) and in alkaline urea to propylurea solutions. Analysis of the dissociation data of duodecamers to monomers, based on equations used in studies of the urea and amide dissociation of human hemoglobin A from our laboratory, suggests few urea and alkylurea binding sites at the areas of hexamer contacts in the associated duodecameric form of L. terrestris hemoglobin. This suggests that hydrophobic interactions are not the dominant forces that govern the state of association of L. terrestris hemoglobin relative to polar and ionic interactions. The unfolding effects of the ureas, at concentrations above the dissociation transitions, are closely similar to their effects on other globular proteins, suggesting that hydrophobic interactions play an important role in the maintenance of the folded conformation of the subunits. Use of the Peller-Flory equation, with binding constants based on free energy transfer data of hydrophobic amino acid side chains and denaturation data used in previous denaturation studies, gave a relatively good acount of the observed denaturation midpoints obtained with the various ureas supporting these conclusions.  相似文献   

4.
The effects of the neutral salts of the Hofmeister series, NaCl, NaClO4, MgCl2, NaI, and also guanidine hydrochloride (Gdn-HCl)on the subunit organization and the state of association of Lumbricus terrestris hemoglobin were examined by light scattering molecular weight measurements. The subunit dissociation of the parent duodecameric structure of 3 x 10(6) molecular weight by various salts is similar in pattern to the sequential splitting of the associated protein to half-molecules of hexamers of 1.5 x 10(6) molecular weight, followed by further dissociation at higher reagent concentration to monomers of 250000 molecular weight. Duodecamer to hexamer dissociation is observed in 0.4 M MgCl2, 1-2 M NaCl, and 1 M Gdn-HCl, while hexamer to monomer dissociation is seen in the presence of 1 M MgCl2. All three species of duodecamers, hexamers, and monomers seem to be present in 1 M NaClO4. Further splitting of the monomers of A subunits to smaller B fragments of one-third to one-quarter molecular weight is observed in 1 M NaI solutions. Optical rotation in the peptide region and absorption measurements in the Soret region indicate the salt dissociation of Lumbricus terrestris hemoglobin is not accompanied by major changes in the folding of the subunits, except in the case of the strong protein denaturant, Gdn-HCl. Relative to the dissociation effects of the urea series of compounds reported in the preceding paper (Herskovits and Harrington, 1975), the neutral salts appear to be much more effective dissociating agents for L. terrestris hemoglobin. This suggests that polar and ionic interactions are relatively more important for the maintenance of the protein than hydrophobic interactions. This conclusion is also supported by calculations of the possible effects of binding of NaClO4, based on the Setschenow constants of the literature describing the interaction of salts with the peptide and hydrophobic alkyl group of the average amino acid found in proteins, on the standard free energy of dissociation of the duodecamer to hexamer.  相似文献   

5.
The sarcoplasmic calcium-binding protein (SCP) of the sandworm Nereis possesses three Ca2(+)-Mg2+ sites but no Ca2(+)-specific site. Binding of Mg2+, but not of Ca2+, displays a marked positive cooperativity. The apparent cooperativity of Ca2+ binding in the presence of Mg2+ results from the allostery in Mg2+ dissociation. Binding of the first Ca2+ or Mg2+ induces all the conformational change, monitored by Trp fluorescence. In displacement reactions the conformational changes occur in the step SCP.Mg3----SCP.Ca1Mg2. Stopped-flow experiments indicate that Trp fluorescence changes upon Ca2(+)-binding are instantaneous whereas Mg2(+)-binding involves a fast pre-equilibrium (Keq = 28 M-1), followed by two slow consecutive conformational changes with k1 = 13.5 s-1 and k2 = 0.21 s-1. The fluorescence change after dissociation of Ca2+ from SCP is monophasic with k = 0.02 s-1; that after Mg2+ dissociation is biphasic with k1 = 0.8 s-1 and k2 = 0.1 s-1. Trp life time measurements also indicate that Ca2(+)- and Mg2(+)-induced conformational changes are completely different. Displacement of bound Ca2+ by Mg2+ can be described by two consecutive reactions in which the first (without fluorescence change) corresponds to the dissociation of the last Ca2+ (k1 = 2.4 s-1) and the second (k2 = 0.45 s-1) to the final conformational change observed upon direct Mg2+ binding. Displacement of bound Mg2+ by Ca2+ follows the kinetic scheme of simple competition; the conformational rate constant approaches asymptotically (up to the limit of 129 s-1) the dissociation rate of Mg2+ as the concentration of Ca2+ increases. In summary, after fast dissociation of Ca2+ or Mg2+, Nereis SCP slowly converts to the metal-free configuration, but in Ca2(+)-Mg2+ exchange reactions, the conformational changes are nearly as fast as the cation dissociation reactions.  相似文献   

6.
The severe acute respiratory syndrome coronavirus enters cells through the activities of a spike protein (S) which has receptor-binding (S1) and membrane fusion (S2) regions. We have characterized four sequential states of a purified recombinant S ectodomain (S-e) comprising S1 and the ectodomain of S2. They are S-e monomers, uncleaved S-e trimers, cleaved S-e trimers, and dissociated S1 monomers and S2 trimer rosettes. Lowered pH induces an irreversible transition from flexible, L-shaped S-e monomers to clove-shaped trimers. Protease cleavage of the trimer occurs at the S1-S2 boundary; an ensuing S1 dissociation leads to a major rearrangement of the trimeric S2 and to formation of rosettes likely to represent clusters of elongated, postfusion trimers of S2 associated through their fusion peptides. The states and transitions of S suggest conformational changes that mediate viral entry into cells.  相似文献   

7.
Kinetics and mechanism of bilirubin binding to human serum albumin   总被引:3,自引:0,他引:3  
The kinetics of bilirubin binding to human serum albumin at pH 7.40, 4 degrees C, was studied by monitoring changes in bilirubin absorbance. The time course of the absorbance change at 380 nm was complex: at least three kinetic events were detected including the bimolecular association (k1 = 3.8 +/- 2.0 X 10(7) M-1 S-1) and two relaxation steps (52 = 40.2 +/- 9.4 s-1 and k3 = 3.8 +/- 0.5 s-1). The presence of the two slow relaxations was confirmed under pseudo-first order conditions with excess albumin. Curve-fitting procedures allowed the assignment of absorption coefficients to the intermediate species. When the bilirubin-albumin binding kinetics was observed at 420 nm, only the two relaxations were seen; apparently the second order association step was isosbestic at this wavelength. The rate of albumin-bound bilirubin dissociation was measured by mixing the pre-equilibrated human albumin-bilirubin complex with bovine albumin. The rate constant for bilirubin dissociation measured at 485 nm was k-3 = 0.01 s-1 at 4 degrees C. A minimum value of the equilibrium constant for bilirubin binding to human albumin determined from the ratio k1/k-3 is therefore approximately 4 X 10(9) M-1.  相似文献   

8.
Calcium dissociation from the C-terminal and N-terminal halves of calmodulin, intact bovine brain calmodulin and the respective phenoxybenzamine complexes or melittin complexes was measured directly by stopped-flow fluorescence with the calcium chelator Quin 2 and, when possible, also by protein fluorescence using endogenous tyrosine fluorescence by mixing with EGTA. Calcium dissociation from the C-terminal half of calmodulin, which contains only the two high-affinity calcium-binding sites, and from intact calmodulin was monophasic, with good correlation of the rates of calcium dissociation obtained by the two methods. The apparent rates with Quin 2 and endogenous tyrosine fluorescence were 13.4 s-1 and 12.8 s-1, respectively, in the C-terminal half and 10.5 s-1 and 10.8 s-1, respectively, in intact calmodulin (pH 7.0, 25 degrees C, 100 mM KCl). Alkylation of the C-terminal half resulted in a biphasic calcium dissociation (Quin 2: kobs 1.90 s-1 and 0.73 s-1 respectively; tyrosine: kobs 1.65 s-1 and 0.61 s-1 respectively). Alkylation of intact calmodulin resulted in a four-phase calcium dissociation measured with Quin 2 (kobs 85.3 s-1, 11.1 s-1, 1.92 s-1 and 0.59 s-1); the latter two phases are assumed to represent calcium release from high-affinity sites since they correspond to the biphasic tyrosine fluorescence change in intact alkylated calmodulin (kobs 2.04 s-1 and 0.53 s-1 respectively) and the rate parameters determined in the C-terminal half. Evidently perturbation of the calcium-binding sites by alkylation reduces the rate of calcium dissociation and allows a distinction to be made between dissociation from each of the two high-affinity sites as well as the distinct conformational change on dissociation of each calcium. Alkylation of the N-terminal half resulted in biphasic calcium release with rates (kobs 153 s-1 and 10.9 s-1 respectively) similar to those observed in intact alkylated calmodulin. The rates of calcium dissociation from calmodulin-melittin or fragment-melittin complexes, measured with Quin 2, were slower and monophasic in the C-terminal half (kobs 1.12 s-1), biphasic in the N-terminal half (kobs 140 s-1 and 26.8 s-1 respectively) and triphasic in intact calmodulin (kobs 126 s-1, 12.1 s-1 and 1.38 s-1). Calmodulin antagonists thus increase the apparent calcium affinity of high and low-affinity sites mainly due to a reduced calcium 'off rate', presumably because of conformation restrictions.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

9.
The catalytic subunit of aspartate transcarbamoylase from Escherichia coli reacts readily with 2,4,6-trinitrobenzenesulfonate, resulting in the loss of enzymatic activity. Substrates and substrate analogs protect the enzyme in a competitive manner, indicating that the loss of activity is due to modification of active-site residues. This conclusion was confirmed by fractionating tryptic digests of the modified protein followed by the identification of active-site lysines 83 and 84 as the modified residues. When three trinitrophenyl groups are incorporated per catalytic trimer, 70% of the activity is lost. The modified protein retains the sedimentation velocity and electrophoretic properties of the native catalytic subunit and can associate with regulatory subunit to form a holoenzyme-like molecule. The trinitrophenylated catalytic trimers have two strong absorption bands at 345 and 420 nm which serve as sensitive spectral probes in difference-spectroscopy experiments. Results from such experiments show that 1) the modified trimeric enzyme binds active-site ligands; 2) dissociation of the trimer into compact, highly structured monomers gives a spectral response distinguishable from that observed when the chains are completely unfolded; and 3) even though dissociation of the trimers to folded monomers causes the complete loss of enzyme activity, the resulting monomers still retain the ability to bind the bisubstrate analog N-(phosphonacetyl)-L-aspartate. These results indicate that the active site must be at least partially formed in the absence of any quaternary structure.  相似文献   

10.
The kinetics of assembly have been monitored spectrophotometrically for normal and variant human oxyhemoglobins in 0.1 M Tris, 0.1 M NaCl, 1 mM Na2EDTA, pH 7.4, at 21.5 degrees C. Oxyhemoglobin versus oxy chain static difference spectra were performed and revealed subtle but significant absorption changes in both the visible and Soret regions. Kinetic experiments were performed by rapidly mixing equivalent (in heme) concentrations of alpha and beta A chains and following the change in absorbance at 583 nm with time. Over a protein concentration range of 10-100 microM in heme prior to mixing, these time courses were homogeneous and followed first-order kinetics, yielding a value of 0.069 s-1 for the apparent rate constant of dissociation of oxygenated beta A chain tetramers. Under these conditions, the overall assembly of oxyhemoglobins S (beta 6Glu----Val) and N-Baltimore (beta 95Lys----Glu) were also governed by the rates of dissociation of their respective oxygenated beta S and beta N-Baltimore chain tetramers with the apparent first-order rate constants of 0.044 and 0.15 s-1, respectively. In the Soret region, the alpha, beta monomer combination reaction could be observed if the protein concentration (heme basis) was lowered and if protein nonequivalency (beta chain exceeded alpha chain concentration) mixing experiments were performed. A kinetic oxyhemoglobin A, oxy-alpha, oxy-beta A monomer difference spectrum could be generated, and simple second-order kinetics were observed (415 nm) yielding rate constants of 2.3, 3.3, and 4.8 X 10(5) M-1 s-1 for the assembly of oxyhemoglobins S, A, and N-Baltimore, respectively. To our knowledge, this is the first kinetic study to reveal significant differences between the rate of association of alpha and beta monomers of hemoglobin A and those of two distinctly charged hemoglobin variants.  相似文献   

11.
The hemoglobin of Biomphalaria glabrata was purified to homogeneity by gel filtration column followed by anion exchange chromatography. The dissociation products were analyzed by a 5–15% gradient polyacrylamide gel electrophoresis containing sodium dodecyl sulfate (SDS-PAGE) giving a band of 270 kDa and a band of 180 kDa after reduction with β-mercaptoethanol. The same profile was obtained in a 3.5% agarose gel electrophoresis containing SDS (SDS-AGE) but showed additional bands of higher molecular weight. These bands were proposed to be monomers, dimers and trimers, since they showed a good correlation in a plot of Rf versus log Mr. After partial reduction in a two-dimensional SDS-AGE, the proposed monomers and dimers produced two and four bands, respectively, likely indicating one to four chains crosslinked by disulfide bridges. Digestion with four different proteases yielded several equivalent fragments with molecular weights multiples of its minimum molecular weight (17.7 kDa). The circular dichroism spectrum of the protein showed a characteristic high α-helix content (70%). It was proposed that this hemoglobin is a pentamer with a molecular weight of aproximately 1.8×103 kDa, assembled by five 360-kDa subunits, each formed by two 180-kDa chains linked in pairs by disulfide bridges and each of these chains, in turn, comprised by ten heme binding domains linked in tandem. These data are compared to the published information for other planorbid extracellular hemoglobins.  相似文献   

12.
Hemerythrin from Siphonosoma cumanense has a trimeric structure consisting of identical subunits, which have no cooperativity nor Bohr effect on oxygen-binding. The trimer was dissociated into its monomers by the modification of the SH group of its cysteines with p-chloromercuriphenylsulfonic acid (PCMPS), which was monitored by stopped-flow of both spectrophotomeric and small angle X-ray scattering methods. The results showed that the process involved sequential modification of the SH groups, dissociation into monomers, and auto-oxidation of ferrous iron in the active center. The modification of the SH groups with PCMPS followed second-order kinetics with a rate constant of 1.8 M-1.s-1. The dissociation and auto-oxidation followed first-order kinetics with rate constants of 4 X 10(-3) s-1 and 5 X 10(-4) s-1, respectively. The obtained rate of auto-oxidation was much faster than that in the native state. These findings lead to the conclusion that the trimeric state of S. cumanense hemerythrin is necessary to prevent auto-oxidation.  相似文献   

13.
Pregnant mare serum gonadotropin exhibits a dissociation at acid pH as shown by the drop of s20,w values from 3.52 S at pH 8.1 to 2.52 S at pH 2.0. The dissociation is accompanied by an absorbance change with a maximum at 287 nm and a parallel loss of both follicle-stimulating hormone (FSH) and luteinizing hormone (LH) activities as followed by radioreceptor assays. The apparent pKa of the acid transition is 3.45 with an extremely slow and temperature-dependent rate at pH 2.0 (1.8 . 10(-4) s-1 at 37 degrees C). By gel filtration the molecular weight of the active hormone is estimated to be 45 000 (rather than the previously reported 53 000-64 000). The active conformation of the hormone includes beta sheet structure (34%) as for other gonadotropin hormones with a minor but significative amount of alpha-helix. Four tyrosine residues were titrated, two of pKa = 10.3 and two of pKa = 11 out of a total of seven tyrosines. The parallel changes in FSH and LH activities during the preparation and the acid transition suggest that the two biological activities are intrinsic properties of the same molecular entity.  相似文献   

14.
The reversible inactivation and dissociation of the allosteric phosphofructokinase from Escherichia coli has been studied in relatively mild conditions, i.e., in the presence of the chaotropic agent KSCN. At moderate KSCN concentration, the loss of enzymatic activity involves two separated phases: first, a rapid dissociation of part of the tetramer into dimers, second, a slower displacement of the dimer-tetramer equilibrium upon further dissociation of the dimer into monomers. These two reactions can no longer be distinguished above 0.3 M KSCN since complete inactivation occurs in a single reaction. Different changes are observed for the fluorescence and the activity of the enzyme in KSCN: the fluorescence is not affected by the dissociation into dimers which is responsible for inactivation. The decrease in fluorescence reflects the change in environment of the unique tryptophan residue, Trp 311, during the dimer to monomer dissociation. This residue belongs to the interface containing the regulatory site, and its native fluorescence indicates that this interface is still present in the dimer. The substrate fructose 6-phosphate protects phosphofructokinase from inactivation by binding to the tetramer and prevents its dissociation into dimers. The presence of phosphoenolpyruvate prevents the slow dissociation of the dimer into monomers, which shows the ability of the dimer to bind the inhibitor. Two successive processes can be observed during reassociation of the protein upon KSCN dilution. First, a fast reaction (k1 = 2 x 10(5) M-1.s-1) is accompanied by a fluorescence increase and results in the formation of the dimeric species.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
The main light harvesting complex of photosystem II in plants, LHCII, exists in a trimeric state. To understand the biological significance of trimerization, a comparison has been made been LHCII trimers and LHCII monomers prepared by treatment with phospholipase. The treatment used caused no loss of chlorophyll, but there was a difference in carotenoid composition, together with the previously observed alterations in absorption spectrum. It was found that, when compared to monomers, LHCII trimers showed increased thermal stability and a reduced structural flexibility as determined by the decreased rate and amplitude of fluorescence quenching in low-detergent concentration. It is suggested that LHCII should be considered as having two interacting domains: the lutein 1 domain, the site of fluorescence quenching [Wentworth et al. (2003) J. Biol. Chem. 278, 21845-21850], and the lutein 2 domain. The lutein 2 domain faces the interior of the trimer, the differences in absorption spectrum and carotenoid binding in trimers compared to monomers indicating that the trimeric state modulates the conformation of this domain. It is suggested that the lutein 2 domain controls the conformation of the lutein 1 domain, thereby providing allosteric control of fluorescence quenching in LHCII. Thus, the pigment configuration and protein conformation in trimers is adapted for efficient light harvesting and enhanced protein stability. Furthermore, trimers exhibit the optimum level of control of energy dissipation by modulating the development of the quenched state of the complex.  相似文献   

16.
In vitro folding pathway of phage P22 tailspike protein.   总被引:10,自引:0,他引:10  
A Fuchs  C Seiderer  R Seckler 《Biochemistry》1991,30(26):6598-6604
The intracellular chain folding and association pathway of the thermostable, trimeric phage P22 tailspike endorhamnosidase has been the subject of a previous detailed study employing temperature-sensitive folding mutants. Recently, reconstitution of native tailspikes from completely unfolded polypeptides has been accomplished, providing a model system to compare protein folding pathways in vivo and in vitro. The in vitro reconstitution pathway of the protein after dilution from guanidine hydrochloride or acid-urea solutions at 10 degrees C was characterized by spectroscopic and hydrodynamic techniques, and may be summarized as an ordered sequence of folding, association, and folding reactions. Multiphasic folding of monomers was indicated by changes in circular dichroism and fluorescence, with a rate constant of k = 1.6 X 10(-3) s-1 for the slowest phase observed spectroscopically. Trimerization of structured monomers was followed by size-exclusion HPLC and was completed within 1.5 h at a protein concentration of 20 micrograms/mL. Although at this time trimers did not exchange subunits, they were readily dissociable by dodecyl sulfate in the cold. Formation of native, detergent-resistant trimers was only completed after 3 days of reconstitution at 10 degrees C. The reconstitution pathway of the tailspike protein closely resembles its intracellular maturation path. Thus, the in vitro reconstitution system, as a valid model of chain folding and association in vivo, should provide the tools to localize the steps or intermediates on the pathway that are the targets of temperature-sensitive folding mutations.  相似文献   

17.
We compared the physicochemical characteristics of alpha 2-macroglobulin (alpha 2M) monomers produced by limited reduction and carboxamidomethylation to those of the naturally occurring monomeric alpha-macroglobulin homologue rat alpha 1-inhibitor 3 (alpha 1 I3). Unlike alpha 1 I3, alpha 2 M monomers fail to inhibit proteolysis of the high molecular weight substrate hide powder azure by trypsin. In contrast to alpha 1 I3, which remains monomeric after reacting with proteinase, alpha 2 M monomers reassociate to higher molecular weight species (dimers, trimers, and tetramers) after reacting with proteinase. Reaction of alpha 2 M monomers at molar ratios of proteinase to alpha 2M monomers as low as 0.3:1 leads to extensive reassociation and is accompanied by complete bait-region and thiolester bond cleavage. During the reaction of alpha 2M monomers with proteinases, the proteinase binds to the reassociating alpha 2M subunits but is not inhibited. Of significance, all the bound proteinase was covalently linked to the reassociated alpha 2M species. Treatment of alpha 2M monomers with methylamine results in thiolester bond cleavage but minimal reassociation. Treatment of alpha 2M monomers with methylamine followed by proteinase results in complete bait-region cleavage and is accompanied by marked reassociation of alpha 2M monomers to higher molecular weight species. However, no proteinase is associated with these higher molecular weight forms. We infer that bait-region cleavage is more important than thiolester bond cleavage in driving alpha 2M monomers to reassociate. Despite many similarities between alpha 1I3 and alpha 2M monomers, significant differences must exist with respect to proteinase orientation within the inhibitor to account for the failure of alpha 2M monomers to protect large molecular weight substrates from proteolysis by bound proteinase, in contrast to the naturally occurring monomeric homologue rat alpha 1 I3.  相似文献   

18.
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   

19.
The extracellular hemoglobin of Glossoscolex paulistus (HbGp) is constituted of subunits containing heme groups, monomers and trimers, and nonheme structures, called linkers, and the whole protein has a minimum molecular mass near 3.1 × 106 Da. This and other proteins of the same family are useful model systems for developing blood substitutes due to their extracellular nature, large size, and resistance to oxidation. HbGp samples were studied by dynamic light scattering (DLS). In the pH range 6.0-8.0, HbGp is stable and has a monodisperse size distribution with a z-average hydrodynamic diameter (Dh) of 27 ± 1 nm. A more alkaline pH induced an irreversible dissociation process, resulting in a smaller Dh of 10 ± 1 nm. The decrease in Dh suggests a complete hemoglobin dissociation. Gel filtration chromatography was used to show unequivocally the oligomeric dissociation observed at alkaline pH. At pH 9.0, the dissociation kinetics is slow, taking a minimum of 24 h to be completed. Dissociation rate constants progressively increase at higher pH, becoming, at pH 10.5, not detectable by DLS. Protein temperature stability was also pH-dependent. Melting curves for HbGp showed oligomeric dissociation and protein denaturation as a function of pH. Dissociation temperatures were lower at higher pH. Kinetic studies were also performed using ultraviolet-visible absorption at the Soret band. Optical absorption monitors the hemoglobin autoxidation while DLS gives information regarding particle size changes in the process of protein dissociation. Absorption was analyzed at different pH values in the range 9.0-9.8 and at two temperatures, 25°C and 38°C. At 25°C, for pH 9.0 and 9.3, the kinetics monitored by ultraviolet-visible absorption presents a monoexponential behavior, whereas for pH 9.6 and 9.8, a biexponential behavior was observed, consistent with heme heterogeneity at more alkaline pH. The kinetics at 38°C is faster than that at 25°C and is biexponential in the whole pH range. DLS dissociation rates are faster than the autoxidation dissociation rates at 25°C. Autoxidation and dissociation processes are intimately related, so that oligomeric protein dissociation promotes the increase of autoxidation rate and vice versa. The effect of dissociation is to change the kinetic character of the autoxidation of hemes from monoexponential to biexponential, whereas the reverse change is not as effective. This work shows that DLS can be used to follow, quantitatively and in real time, the kinetics of changes in the oligomerization of biologic complex supramolecular systems. Such information is relevant for the development of mimetic systems to be used as blood substitutes.  相似文献   

20.
Abstract The effects of solar and artificial ultraviolet radiation on photosynthetic oxygen production and phycobiliprotein composition were investigated in the freshwater flagellate, Cyanophora paradoxa . The phycobiliproteins of the cell acting as photosynthetic accessory pigments were found to be readily affected by even short exposure to ultraviolet radiation, while the membrane-bound chlorophyll protein complexes were hardly impaired as demonstrated by SDS polyacrylamide gel electrophoresis, isoelectric focussing and fast protein liquid chromatography (FPLC). The phycobilisomes are dissembled from the outside inwards and go from higher molecular weight components to hexamers ( αβ )6, trimers ( αβ )3 and finally monomers (αβ). Fluorescence spectra indicated that the energy transfer from accessory pigments to the photosystems was impaired by ultraviolet radiation. Photosynthetic oxygen production was affected on a much faster timescale than changes in the absorption spectra or at the protein level.  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