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.
G Voordouw  R S Roche 《Biochemistry》1975,14(21):4667-4673
The functional properties of the four calcium ions, bound by thermolysin, appear to be very similar to those of the single calcium ion bound by thermomycolase (G. Voordouw and R.S. Roche (1975), Biochemistry, preceding paper in this issue). Hence when the free calcium ion concentration is varied in the range where the calcium double-site dissociates (G. Voordouw and R.S. Roche (1974), Biochemistry 13, 5017), no changes are observed in the sedimentation coefficient or the peptide circular dichroism. Differences in molar ellipticity and molar extinction coefficient occur in the aromatic ultraviolet region, which parallel the occupancy of the calcium binding double site. The difference spectrum, characterized by a main band at 290 nm and a somewhat smaller band at 283 nm, is interpreted as due to the transfer of a partially buried tryptophan residue to the aqueous solvent upon dissociation of the two calcium ions from the double site. This is most likely Trp-186, which is in between Asp-185 and Glu-187, two chelating amino acids of this site. From the calcium dependence of the rate constant for autolytic degradation we conclude, as for thermomycolase, that only conformers devoid of bound calcium ion serve as substrates in the reaction. This rate constant increases about 1000-fold, when the double site dissociates. Hydrogen-tritium exchange studies show the presence of a large stable strcutural core, comprising about 32% of all the peptide hydrogens present. These do not exchange-in after 24 hr at 25degreesC, pH 9.0, ionic strenth 0.1. The exchange-out of 60 slow hydrogens was found to be independent of the free calcium ion concentration in the range 2.0-8.0 X 10(-4) M, where all four calcium-binding sites are saturated. The calcium dependence of the first-order rate constant for thermal denaturation at 80degreesC, pH 7.0, indicates that thermolysin is stabilized by only one calcium ion under these conditions. These observations are rationalized in terms of a calcium-binding model for thermolysin and the known three-dimensional structure of the enzyme and its calcium-binding sites.  相似文献   

2.
G Voordouw  C Milo  R S Roche 《Biochemistry》1976,15(17):3716-3724
The total kinetic thermal stability of a protein molecule, expressed as the total free energy of activation in thermal denaturation reactions, can be separated into an intrinsic contribution of the polypeptide chain and a contribution due to the binding of calcium ions. The theory for this procedure is applied to thermal denaturation data, obtained at the pH of optimum stability, for the serine proteases, thermomycolase and subtilisin types Carlsberg and BPN', and for the zinc metalloendopeptidases, thermolysin and neutral protease A. The results, obtained from Arrhenius plots at high and low free calcium ion concentrations, reveal a considerable variation in the calcium ion contribution to the total kinetic thermal stability of the various enzymes. In the serine protease group, at 70 degrees C, the stability is largest for thermomycolase, mainly due to a relatively high intrinsic contribution. For the metalloendopeptidases the total kinetic thermal stability is largest for thermolysin, the difference between thermolysin and neutral protease A being dominated by bound calcium ion contributions. The intrinsic kinetic thermal stability of the polypeptide chain of thermolysin is considerably smaller than that of any of the serine proteases and is probably of the same order of magnitude as that of neutral protease A. Thus, the well known total kinetic thermal stability of thermolysin is due mainly to a single calcium ion (Voordouw, G., and Roche, R. S. (1975), Biochemistry 14, 4667) that binds with high affinity even at very high temperatures (K congruent to 6 X 10(7) M-1 at 80 degrees C).  相似文献   

3.
A protein which showed high affinity for calcium ions was isolated from bull seminal vesicle secretion and seminal plasma. Its calcium-binding activity depended on the ionic strength and pH of the medium. The dissociation constant was 7-7 X 10(-7) M and there were 14 binding sites per protein molecule. The molecular weight of calcium-binding protein from bull seminal vesicle secretion, estimated by the gel filtration method, was 110,000. The protein may be involved in the regulation of the calcium ion level in seminal plasma.  相似文献   

4.
Avian anterior (ALD) and posterior (PLD) latissimus dorsi muscle homogenates were tested for autolysis of proteins at pH values for which muscle proteases have been described (pH 4.0, 7.0, and 8.5). The action of MgCl2 and CaCl2 at physiological concentrations on these autolytic activities was measured to determine if changes in ion concentration could initiate proteolysis. At pH 7.0, it was found that alterations in CaCl2 concentrations within physiological limits in the presence of a MgCl2 concentration below physiological levels could activate or inhibit autolysis. Changes in magnesium concentration also could activate autolysis. This calcium effect was greatest for the PLD muscle. The possible role of calcium compartmental changes in initiation of contractile protein breakdown following denervation is discussed.  相似文献   

5.
Vertical starch-gel electrophoresis at pH 8.6 revealed extensive hemoglobin multiplicity with several distinct cathodal and anodal hemoglobin components. Anodal hemoglobin components are present throughout the life cycle of the king salmon. Additional cathodal components are found in the adult fish. Cathodal hemoglobin components exhibited a higher oxygen affinity (P50 = 10.2 mm at 13 degrees C, pH 7.3) than the anodal hemoglobin components (P50 = 21.8 mmHg at 13 degrees C). Oxygen binding of the anodal hemoglobins are sensitive to pH, temperature, organic phosphates (ATP and GTP), as well as, ionic strength; binding of oxygen to the cathodal hemoglobins is independent of pH and not affected by organic phosphates. Anodal hemoglobin components are less resistant to thermal denaturation over the pH 6.0 to 8.0 range. Isothermal urea denaturation of separated anodal and cathodal hemoglobin fractions of the king salmon indicate inherent differences in the stabilization energies of these hemoglobins. Autoxidation of these hemoglobins occurs around pH 7.0 and below, as well as, in the presence of increasing Cl- concentrations.  相似文献   

6.
Since 1H-NMR spectra of the calcium bound form (holo) and the calcium free form (apo) of equine lysozyme have an overall similarity, the folded structure of apo equine lysozyme seems to be similar to the holo structure at 25 degrees C and pH 7.0, even at low ionic strengths except for subtle conformational change. However, calcium titration experiments showed that a number of resonances change by a slow exchange process. The changes saturated at one calcium ion per one lysozyme molecule, and no more change was observed by further addition of calcium ions. This shows that just one calcium ion binds to equine lysozyme. To make assignments for these changed proton resonances, two-dimensional 1H-NMR studies, correlated spectroscopy (COSY), two-dimensional homonuclear Hartmann-Hahn spectroscopy (HOHAHA) and nuclear Overhauser effect spectroscopy (NOESY) were carried out. A structural model of equine lysozyme based on the crystal structure of human lysozyme was estimated and used to assign some resonances in the aromatic and beta-sheet regions. It was possible to use some proton signals as a probe to determine the specific conformational change induced by calcium ions. The calcium binding constant KCa was estimated from calcium titration experiments in which changes in the proton signal were monitored. The log KCa value was found to be on the order of 6-7, which is in agreement with the calcium binding constant determined by fluorescence probes. This means that the protons are affected by specific calcium binding.  相似文献   

7.
Human MMP-26 (matrix metalloproteinase-26) (also known as endometase or matrilysin-2) is a putative biomarker for human carcinomas of breast, prostate and other cancers of epithelial origin. Calcium modulates protein structure and function and may act as a molecular signal or switch in cells. The relationship between MMPs and calcium has barely been studied and is absent for MMP-26. We have investigated the calcium-binding sites and the role of calcium in MMP-26. MMP-26 has one high-affinity and one low-affinity calcium binding site. High-affinity calcium binding was restored at physiologically low calcium conditions with a calcium-dissociation constant of 63 nM without inducing secondary and tertiary structural changes. High-affinity calcium binding protects MMP-26 against thermal denaturation. Mutants of this site (D165A or E191A) lose enzymatic activity. Low-affinity calcium binding was restored at relatively high calcium concentrations and showed a K(d2) (low-affinity calcium-dissociation constant) value of 120 microM, which was accompanied with the recovery of enzymatic activity reversibly and tertiary structural changes, but without secondary structural rearrangements. Mutations at the low-affinity calcium-binding site (C3 site), K189E or D114A, induced enhanced affinity for the Ca2+ ion or an irreversible loss of enzymatic activity triggered by low-affinity calcium binding respectively. Mutation at non-calcium-binding site (V184D at C2 site) showed that C2 is not a true calcium-binding site. Observations from homology-modelled mutant structures correlated with these experimental results. A human breast cancer cell line, MDA-MB-231, transfected with wild-type MMP-26 cDNA showed a calcium-dependent invasive potential when compared with controls that were transfected with an inactive form of MMP-26 (E209A). Calcium-independent high invasiveness was observed in the K189E mutant MDA-MB-231 cell line.  相似文献   

8.
The stoichiometry of calcium binding to specific sites (i.e., those producing enzyme activation) was found to be 8-10 nmol/mg protein in native sarcoplasmic reticulum vesicles, and 13.9-15.4 nmol/mg of ATPase purified by non-ionic detergent solubilization and anion exchange chromatography. Parallel measurements of phosphoenzyme yielded levels of 4.0-4.9 and 6.0-7.7 nmol/mg of protein in the two preparations, respectively, demonstrating that each 115 kDa ATPase chain includes one catalytic site and two calcium binding sites. The apparent association constant, K = (6 +/- 2) X 10(5) M-1, and the binding cooperativity, nH = 1.9, were unchanged when measurements were carried out with native sarcoplasmic reticulum vesicles and when the membrane surface charge was altered by lipid substitution with phosphatidylcholine or phosphatidylserine, at neutral pH in the presence of 10 mM MgCl2 and 80 mM KCl. On the other hand, the apparent association constant was increased in the absence of Mg2+ or, to a lesser extent, in the absence of monovalent cations. It was also observed that the cooperative character of the calcium binding isotherms was reduced in low ionic-strength media. Analysis of the electrostatic effects indicates that the calcium-binding domain is shielded from the membrane phospholipid surface charge by virtue of its location within the ATPase protein. The effects of various electrolytes are attributed to monovalent-cation binding in the calcium-binding domain. The apparent loss of cooperativity of the calcium binding isotherms at low ionic strength is attributed to a progressive displacement of the titration curve which is minimal at low degrees of saturation and becomes larger at higher degrees of saturation. This behavior is described quantitatively by the progressive effect of calcium binding on an electrostatic potential generated by localized protein charge densities within, or near, the calcium-binding domain.  相似文献   

9.
Calcium binding by rabbit skeletal myosin, thin filaments and myofibrils was measured in solutions with and without 2 mM MgATP and with ionic strengths adjusted with KCl to 0.05, 0.10 and 0.14 M. Free Mg2+ was held constant at 1 mM, pH at 7.0 and temperature at 25 °C. In the presence of MgATP, the relation between free Ca2+ and myofibrillar bound calcium shifted to the left as ionic strength was decreased from 0.14 to 0.05 M. In the absence of MgATP, myofibrillar calcium binding was enhanced over a wide range of free Ca2+ concentration, but calcium binding was no longer a function of ionic strength. Similarly, calcium binding by thin filaments and myosin was unaffected by changes in ionic strength from 0.05 to 0.14 M. In view of evidence that cross-bridge connections between thick and thin filaments increase as ionic strength decreases, our results suggest that these connections enhance myofibrillar calcium binding. These results thus confirm previous data of Bremel and Weber (Bremel, R. D. and Weber, A. (1972) Nature New Biol. 238, 97–101) who first showed that nucleotide-free cross-bridge connections enhance thin filament calcium binding.  相似文献   

10.
Ligand requirements for Ca2+ binding to EGF-like domains.   总被引:3,自引:0,他引:3  
Site-specific mutagenesis studies of the first epidermal growth factor-like (EGF-like) domain of human clotting factor IX suggest that the calcium-binding site present in this domain (dissociation constant Kd = 1.8 mM at pH 7.5 and ionic strength I = 0.15) involved the carboxylate residues Asp47, Asp49 and Asp64. To further characterize the ligands required for calcium binding to EGF-like domains, two new mutations, Asp47----Asn and Asp49----Asn, were introduced into the domain by peptide synthesis. 1H-NMR spectroscopy was used to obtain the dissociation constants for calcium binding to these mutations. Calcium binding to the Asp49----Asn modified domain is only mildly affected (Kd = 6 mM, I = 0.15), whereas binding to the Asp47----Asn modified domain is severely reduced (Kd = 42 mM, I = 0.15). From these data, it is proposed that the anionic oxygen atoms of the side chains of residues 47 and 64 are essential for calcium binding, whereas the side chain ligand for calcium at residue 49 can be a carboxyamide oxygen. As a control, the introduction of the modification Glu78----Asp in a region of the domain not believed to be involved in calcium binding had very little effect on the Kd for calcium (Kd = 2.6 mM, I = 0.15). Finally, the effect of an Asp47----Gly substitution found in the natural haemophilia B mutant, factor IXAlabama, was investigated. This peptide has a markedly reduced affinity for calcium (Kd = 37 mM, I = 0.15), suggesting that the defect in factor IXAlabama is due to impaired calcium binding to its first EGF-like domain.  相似文献   

11.
In this work, we explored the acid-induced unfolding pathway of non-porin outer membrane protein (OMP), an immunogenic protein from Salmonella Typhi, by monitoring the conformational changes over a pH range of 1.0–7.0 by circular dichroism, intrinsic fluorescence, ANS binding, acrylamide quenching, and dynamic light scattering. The spectroscopic measurements showed that OMP in its native state at pH 7.0 exists in more stable and compact conformation. In contrast, at pH 2.0, OMP retains substantial amount of secondary structure, disrupted side chain interactions, increased hydrodynamic radii, and nearly four-fold increase in ANS fluorescence with respect to the native state, indicating that MG state exists at pH 2.0. Quenching of tryptophan fluorescence by acrylamide further confirmed the accumulation of a partially unfolded state between native and unfolded state. The effect of pH on the conformation and thermostability of OMP points towards its heat resistance at neutral pH (T m?~?69 °C at pH 7.0, monitored by change in MRE222 nm). Acid unfolded state was also characterized by the lack of a cooperative thermal transition. All these results suggested that acid-induced unfolded state of OMP at pH 2.0 represented the molten globule state. The chemical denaturation studies with GuHCl and urea as denaturants showed dissimilar results. The chemical unfolding experiments showed that in both far-UV CD and fluorescence measurements, GuHCl is more efficient than urea. GuHCl is characterized by low C m (~1 M), while urea is characterized by high C m (~3 M). The fully unfolded states were reached at 2 M GuHCl and 4 M urea concentration, respectively. This study adds to several key considerations of importance in the development of therapeutic agents against typhoid fever for clinical purposes.  相似文献   

12.
The Fourier-transform infrared spectroscopy (FTIR) technique with a diamond anvil cell has been applied for examination of the pressure-induced changes occurring in the secondary structure of the alpha-lactalbumin. This is the first high-pressure FTIR study of a calcium-binding protein which simultaneously takes into account spectral changes in both the calcium-ion-binding carboxyl groups' band and the amide I/I' vibrational band. Spectral behavior of three kinds of the protein: the undeuterated holoform, the fully deuterated holoform, and the undeuterated apoform was compared in the pressure range from 0.1 MPa up to 740 MPa. We found that the binding of calcium remarkably stabilizes the alpha-lactalbumin against pressure as it is followed approximately by a 200-MPa increase of the value of pressure at which denaturation occurs. A quantitative analysis of the band of antisymmetrical stretching vibrations of the calcium-binding carboxyl groups revealed that the pressure-induced changes in the calcium-binding loop occur in two stages. Binding of the calcium ion seemingly increases the pressure-stability of the calcium-binding loop to a higher degree than the pressure-stability of the secondary structure of the alpha-lactalbumin. We have also discussed in detail the complex pressure-enhanced H/D exchange in the alpha-lactalbumin. Finally, we have proposed a new assignment of major peaks in the helical region of the amide I/I' spectral band of the partially deuterated alpha-lactalbumin.  相似文献   

13.
A method is described for the purification of native hexokinases P-I and P-II from yeast using preparative isoelectric focussing to separate the isozymes. The binding of glucose to hexokinase P-II, and the effect of this on the monomer--dimer association--dissociation reaction have been investigated quantitatively by a combination of titrations of intrinsic protein fluorescence and equilibrium ultracentrifugation. Association constants for the monomer-dimer reaction decreased with increasing pH, ionic strength and concentration of glucose. Saturating concentrations of glucose did not bring about complete dissociation of the enzyme showing that both sites were occupired in the dimer. At pH 8.0 and high ionic strength, where the enzyme existed as monomer, the dissociation constant of the enzyme-glucose complex was 3 X 10(-4) mol 1(-1) and was independent of the concentration of enzyme. Binding to the dimeric form at low pH and ionic strength (I=0.02 mol 1(-1), pH less than 7.5) was also independent of enzyme concentration (in the range 10-1000 mug ml-1) but was much weaker. The process could be described by a single dissociation constant, showing that the two available sites on the dimer were equivalent and non-cooperative; values of the intrinsic dissociation constant varied from 2.5 X 10(-3) mol 1(-1) at pH 7.0 to 6 X 10(-3) at pH 6.5. Under intermediate conditions (pH 7.0, ionic strength=0.15 mol 1(-1)), where monomer and dimer coexisted, the binding of glucose showed weak positive cooperatively (Hill coefficient 1.2); in addition, the binding was dependent upon the concentration of enzyme in the direction of stronger binding at lower concentrations. The results show that the phenomenon of half-sites reactivity observed in the binding of glucose to crystalline hexokinase P-II does not occur in solution; the simplest explanation of our finding the two sites to be equivalent is that the dimer results from the homologous association of two identical subunits.  相似文献   

14.
The specificity of thermomycolase toward glucagon and the oxidized A and B chains of insulin was investigated. Extensive digestion of glucagon occurred when conducted at pH 7.0 and 45 degrees C for 40 min, whereas hydrolysis of only three peptide bonds occurred at pH 7.0 and 28 degrees C for 5 min. A similar situation was observed for the oxidized B chain of insulin, which exhibited only a single major cleavage after 5 min at 25 degrees C. No well-defined specificity for particular amino acid residues was evident, but ready hydrolysis of peptide bonds occurred within sequences containing non-polar residues. This endoproteinase must therefore possess an extended hydrophobic binding site for polypeptides. Thermomycolase hydrolysed acetylalanylalanylalanine methyl ester and elastin-Congo Red at 22 and 8.5 times the rate of porcine elastase respectively. A limited degradation of native collagen and significant hydrolysis of benzyloxycarbonyl-Gly-Pro-Leu-Gly-Pro were suggestive of some collagenase-like activity. No keratinase activity was apparent.  相似文献   

15.
The reduction of horse heart cytochrome c with ascorbate in the absence of urea and in its presence, 0 to 8 M, pH 7.0, has been investigated using a stopped flow technique and the absorptivity at 550 nm as the monitoring probes, and by using the rate of oxidizability with molecular oxygen. Reduction is found to be consistent with a mechanism involving (i) a urea-dependent equilibrium step between an ascorbate-reducible and an irreducible form, with a [urea]1/2 of 7.5 M and a reversion rate constant of 0.05 +/- 0.02 s-1, (ii) the binding of ascorbate to cytochrome c, with a binding constant of 5.9 M-1 in the absence of urea which decreases to a value of 2.7 M-1 above 5.5 M urea, and (iii) a reduction step, with a urea-independent rate constant of 2.9 +/- 0.3 s-1. This scheme is interpreted in terms of an electron-transfer pathway involving neither the classical "adjacent" attack nor attack at the exposed heme edge, i.e. "remote" attack, but rather, through an alternate pathway involving binding at some site other than the heme crevice opening and a migration path of rather low electron-transfer efficiency. The urea-linked ascorbate reduction step is th X2 in equilibrium D step of the urea denaturation mechanism (Myer, Y. P., MacDonald L. H., Verma, B. C., and Pande, A. J. (1980) Biochemistry 19, 199-207), and the 9 M urea form, D, is the irreducible form. Form X2 and the other intermediate form, X1, are found to be reducible directly by ascorbate, and not through reversion to the native form of the protein. both the integrity of the heme crevice and the polypeptide-organized structures are of little importance as far as ascorbate reducibility is concerned, but the integrity of the structural and protein functional changes reflecting the X2 in equilibrium D step of the mechanism directly or indirectly determines the reducibility of the protein.  相似文献   

16.
Structural changes of troponin C on calcium binding were studied by hydrogen ion titration, circular dichroism, and fluorescence measurements. The potentiometric titration curves in the carboxyl region are shifted towards lower pH with calcium binding. The intrinsic pK of the carboxyl groups at the calcium binding sites decreases by 0.8 pK unit on calcium binding; on the other hand, magnesium ions have little effect on the intrinsic pK of the carboxyl groups. The intrinsic pK of the imidazole group is not affected by calcium binding. The value of w, an electrostatic interaction factor, is identical for calcium-free and calcium-bound troponin C and is about half of the value calculated assuming a compact sphere. The results of difference titration on the calcium binding indicate that the pH of troponin C solution increases on addition of CaCl2 up to 2 mol of Ca2+ per mol of troponin C and then decreases on further addition of CaCl2. The pH increase is depressed in the presence of MgCl2, in the low pH region, or at high ionic strength. The pH increase is also observed on addition of MgCl2. The ellipticity at 222 nm was measured under the same conditions as the difference titration measurements, and the relation between the pH change and the conformational change of troponin C on calcium binding is discussed based on the results obtained. The number of calcium binding sites and the binding constants estimated by analysis of these difference titration curves were in agreement with the results of Potter and Gergely (22). No magnesium binding site was observed. The tyrosine fluorescence measurements indicated that the binding site near tyrosine-109 is one of the high affinity sites.  相似文献   

17.
The effect of pH, mental ions, and denaturing reagents on the thermal stability of thermophilic alpha-amylase [EC 3.2.1.1] were examined. The enzyme was most stable at around pH 9.2, which is coincident with the isoelectric point of the enzyme. The stability of the enzyme was increased by the addition of calcium, strontium, and sodium ions. The addition of calcium ions markedly stabilized the enzyme. The protective effects of calcium and sodium ions were additive. At room temperature, no detectable destruction of the helical structure of the enzyme was observed after incubation for 1 hr in the presence of 1% sodium dodecylsulfate, 8 M urea or 6 M guanidine-HC1. The addition of 8 M urea or 6 M guanidine-HC1 lowered the thermal denaturation temperature of the enzyme. The enzyme contained one atom of tightly bound intrinsic calcium per molecule which could not be removed by electrodialysis unless the enzyme was denatured. The rate constants of inactivation and denaturation reactions in the absence and presence of calcium ions were measured and thermodynamic parameters were determined. The presence of calcium ions caused a remarkable decrease in the activation entropy.  相似文献   

18.
Systematic photometric studies are presented to analyze the thermal denaturation behaviour with and without formamide of metaphase chromosome suspensions in comparison to DNA solutions. Temperature dependent hyperchromicity measurements at 256 nm and 313 nm were performed using an appropriately designed computer-controlled photometer device. Due to an upright optical axis, this allowed absorbance measurements with negligible sedimentation effects not only for solutions of pure DNA, but also for particle suspensions of isolated metaphase chromosomes. This device has a temperature resolution of +/- 0.5 degrees C and an optical sensitivity of 10(-3) to 10(-4) optical density. For calf thymus DNA the reduction of the melting point with the increase of formamide in the solution was measured at pH 7.0 and pH 3.2. The good correlation of the theoretical approximation to experimental data indicated the suitability of the apparatus to quantitatively describe DNA conformation changes induced by thermal denaturation. For metaphase chromosome preparations of Chinese hamster culture cells, absorbance changes were measured between 20 degrees C and 95 degrees C with a temperature gradient of 1 degrees C/min. These measurements were performed at pH 7.0 and at pH 3.2. The denaturation profiles (= first derivative of the absorbance curve) resulted in a highly variable peak pattern at 256 nm and 313 nm indicating complex conformation changes. A statistical evaluation of the temperature values of the peak maxima resulted in temperature ranges typical for chromosomal conformation changes during thermal treatment. Especially the range of highest temperature values was independent from pH modifications. For pH 3.2 the influence of formamide on the denaturation behaviour of metaphase chromosome preparations was analyzed. In contrast to pure DNA solutions, a reduction of the "melting point" (i.e. the maximum temperature at which a conformation change takes place) was not found. However, the denaturation behaviour depended on the duration of formamide treatment before the measurement.  相似文献   

19.
This paper describes the sequence homology of calcium-binding proteins belonging to the troponin C superfamily. Specifically, this similarity has been examined for 276 twelve-residue calcium-binding loops. It has been found that, in the calcium-binding loop, several residues appear invariant, regardless of the species of origin or the affinity of the protein. These residues are Asp at position 1 (+X of the coordinating position of the calcium), Asp or Asn at position 3 (+Y), Gly at position 6, Ile at position 8, and Glu at position 12 (-Z). It has also been found that conservation of certain residues can vary in similar sites in similar proteins. For example, position 3 (+Y) in site 3 of troponin C is always an Asn, whereas in calmodulin the residue is always Asp. This study also examined the calcium-binding affinities of peptide fragments comprising the loop, helix-loop, loop-helix, and helix-loop-helix. These were compared with larger enzymatic or chemically generated protein fragments in an effort to understand the various contributions to the calcium-binding affinity of a single-site versus a two-site domain as found in troponin C and calmodulin. Based on free energy differences, it was found that a 34-residue helix-loop-helix peptide represents about 60% of the binding affinity found in the intact protein. Cooperativity with a second calcium binding site accounted for the remaining 40% of the affinity.  相似文献   

20.
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   

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