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.
Properties of talin from chicken gizzard smooth muscle   总被引:9,自引:0,他引:9  
This paper describes the structural and biochemical characterization of talin, a protein localized to various cellular sites where bundles of actin filaments attach to the plasma membrane. By sodium dodecyl sulfate-polyacrylamide gel electrophoresis, the protein has a molecular mass of 225,000 +/- 5,000 daltons. Hydrodynamic measurements at protein concentrations less than 0.72 mg/ml indicate a monomeric protein with a native molecular mass of 213,000 +/- 15,000 daltons. Sedimentation equilibrium experiments indicate self-association at protein concentrations of 0.72 mg/ml and higher. The data suggest that this self-association is a simple monomer:dimer equilibrium over the range of concentrations observed. At low protein concentrations where talin is a monomer, the Stokes radius and sedimentation coefficient vary with ionic strength. Under low ionic strength conditions (5-20 mM NaCl), talin has a Stokes radius of 6.5 nm and a sedimentation value of 9.4, suggesting an asymmetric globular molecule; whereas under high ionic strength conditions (200 mM NaCl), the Stokes radius increases to 7.7 nm and the sedimentation coefficient decreases to 8.8, suggesting a more elongated protein. This conformation change is confirmed by electron microscopy which reveals a more globular protein at low ionic strength which unfolds to become an elongated flexible molecule as the ionic strength is increased to physiological and higher levels. The amino acid composition of talin indicates a low level of aromatic residues, consistent with its relatively low extinction coefficient, talin has an isoelectric point between pH 6.7 and 6.8 based on isoelectric focusing. The detailed purification of talin is described.  相似文献   

2.
The catalytic domains of activated C1r and C1s, comprising the C-terminal region of the A chain (gamma), disulphide-linked to the B chain, were obtained by limited proteolysis of the native proteases with chymotrypsin and plasmin, respectively, and studied by small angle neutron scattering. For activated C1s (gamma-B), a molar mass of 45,000 +/- 5000 g/mol, and a relatively large radius of gyration (Rg) of 28 +/- 1 A were determined, excluding a single globular domain. The corresponding values for activated C1r (gamma-B)2 (90,000 g/mol, Rg = 34 +/- 1 A) are consistent with a dimer involving the loose packing of two (gamma-B) subunits. Various models of the dimer are discussed in the light of neutron scattering and other data.  相似文献   

3.
The structure of protein SI of Thermus thermophilus (M = 61 kDa) in solution at low and moderate ionic strengths (0 M and 100 mM NaCl, respectively) has been studied by small-angle X-ray and neutron scattering. It was found that protein S1 has a globular conformation under both ionic conditions. The modelling of different packing of six homologous domains of S1 on the basis of the NMR-resolved structure of one domain showed that the best fit of calculated scattering patterns from such complexes to experimental ones is observed at a compact package of the domains. The calculated value of the radius of gyration of the models is 28-29 angtroms, which is characteristic for globular proteins with a molecular mass of about 60 kDa. It was found that protein S1 has a tendency to form associates, and the type of the associate depends on ionic strength. These associates have, in general, two or three monomers at a moderate ionic strength, while at a low ionic strength the number of monomers exceeds three and they are packed in a compact manner. Strongly elongated associates were observed in neutron experiments at a moderate ionic strength in heavy water. The association of protein molecules was also confirmed by the data of dynamic light scattering. From these data, the translational diffusion coefficient of protein S1 at a moderate ionic strength was calculated to be (D20,w = (2.7 +/- 0.1) x 10(-7)cm2/s). This value is essentially smaller than the expected value (D20,w = (5.8 - 6.0) x 10(-7)cm2/s) for the S1 monomer in the globular conformation, indicating the association of protein molecules under equilibrium conditions.  相似文献   

4.
Models of fibronectin.   总被引:4,自引:0,他引:4       下载免费PDF全文
The radius of gyration of human plasma fibronectin was determined by light scattering both under conditions in which the molecule is in an extended conformation (ionic strength 1.01 M, pH 8) and close to its native, more compact conformation (ionic strength 0.16 M, pH 8). These values were found to be 17.5 +/- 0.8 nm and 10.7 +/- 0.9 nm respectively, for a constant mol. wt of 533,000 +/- 8000, in excellent agreement with the value of 520,000 deduced from its known composition. A set of models, each made of two identical, end-to-end joined chains of 28 beads, was then constructed, and their calculated physico-chemical parameters were compared with those available for the whole fibronectin molecule and for some of its proteolytic fragments in both conformations. Two possible models for the circulating form are presented here: in both, the fibronectin molecule is in a compact, tangled conformation, with the amino-terminal end of one chain folded over to the carboxy end of itself or of the other chain either in a hairpin or in a circular fashion. With the exception of the carboxy-terminal fibrin(ogen)-binding domains, all the domains appear to be well exposed to the solvent, and thus free to interact with potential ligands.  相似文献   

5.
C Saez  N M Thielens  E S Bjes  A F Esser 《Biochemistry》1999,38(21):6807-6816
The association of terminal complement proteins was investigated by analytical ultracentrifugation and multi-angle laser light scattering. Native C8 and C9 formed a heterodimer in solution of physiological ionic strength with a free-energy change DeltaG degrees of -8.3 kcal/mol and a dissociation constant Kd of 0.6 microM (at 20 degrees C) that was ionic strength- and temperature-dependent. A van't Hoff plot of the change in Kd was linear between 10 and 37 degrees C and yielded values of DeltaH degrees = -12.9 kcal/mol and DeltaS degrees = -15.9 cal mol-1 deg-1, suggesting that electrostatic forces play a prominent role in the interaction of C8 with C9. Native C8 also formed a heterodimer with C5, and low concentrations of polyionic ligands such as protamine and suramin inhibited the interaction. Suramin induced high-affinity trimerization of C8 (Kd = 0.10 microM at 20 degrees C) and dimerization of C9 (Kd = 0.86 microM at 20 degrees C). Suramin-induced C8 oligomerization may be the primary reason for the drug's ability to prevent complement-mediated hemolysis. Analysis of sedimentation equilibria and also of the fluorescence enhancement of suramin when bound to protein provided evidence for two suramin-binding sites on each C9 and three on each C8 in the oligomers. Oligomerization could be reversed by high suramin concentrations, but 8-aminonaphthalene-1,3,6- trisulfonate (ANTS2- ), which mimics half a suramin molecule, could not compete with suramin binding and oligomerization suggesting that the drug also binds nonionically to the proteins.  相似文献   

6.
The association of native C1 with physiologically relevant proteins was studied by ultracentrifugation. 125I-C1 was centrifuged through numerous sucrose density gradients, each of which contained a different concentration of monomeric (19S) IgM throughout the gradient. The s-rate of C1 (16S) increased with increasing IgM input to a maximum of 32S. In the absence of C1q, the C1r2s2 subunit did not bind to the Ig. In gradients containing physiologic concentrations of IgM (1.3 mg/ml) at 0.14 M ionic strength, the observed s-rate of C1 was 21S. In the presence of 13 mg/ml IgG, C1 sedimented with an s-rate of 19S. Thus, under physiologic conditions, a significant fraction of native C1 is reversibly bound to monomeric Ig. SDS-PAGE analyses show that this interaction does not lead to C1 activation. The interaction of native C1 with C1 inhibitor (C1-In) was studied by ultracentrifugation at physiologic ionic strength. Purified 125I-C1-In alone sedimented with an s-rate of 4S. However in the presence of excess native C1, one-third of the C1-In co-sedimented with C1 at a 16S position. For these studies, 100 microM nitrophenylguanidinobenzoate (NPGB) was present throughout the sucrose density gradient to prevent C1 activation during centrifugation. As the concentration of NPGB was increased, the percent of 125I-C1-In at 16S decreased, indicating that C1-In was binding (reversibly) to the C1 active site region(s), which is at least partially accessible in uncleaved C1. In controls, when NPGB was omitted or activated C1 was used, the s-rate of 125I-C1-In was only 12S due to the release of C1rC1s(C1-In)2 from activated C1. Thus, under physiologic conditions native C1 is reversibly bound to C1-In.  相似文献   

7.
Maltose binding protein (MBP) is a large, monomeric two domain protein containing 370 amino acids. In the absence of denaturant at neutral pH, the protein is in the native state, while at pH 3.0 it forms a molten globule. The molten globule lacks a tertiary circular dichroism signal but has secondary structure similar to that of the native state. The molten globule binds 8-anilino-1-naphthalene sulfonate (ANS). The unfolding thermodynamics of MBP at both pHs were measured by carrying out a series of isothermal urea melts at temperatures ranging from 274-329 K. At 298 K, values of deltaGdegrees , deltaCp, and Cm were 3.1+/-0.2 kcal mol(-1), 5.9+/-0.8 kcal mol(-1) K(-1) (15.9 cal (mol-residue)(-1) K(-1)), and 0.8 M, respectively, at pH 3.0 and 14.5+/-0.4 kcal mol(-1), 8.3+/-0.7 kcal mol(-1) K(-1) (22.4 kcal (mol-residue)(-1) K(-1)), and 3.3 M, respectively, at pH 7.1. Guanidine hydrochloride denaturation at pH 7.1 gave values of deltaGdegrees and deltaCp similar to those obtained with urea. The m values for denaturation are strongly temperature dependent, in contrast to what has been previously observed for small globular proteins. The value of deltaCp per mol-residue for the molten globule is comparable to corresponding values of deltaCp for the unfolding of typical globular proteins and suggests that it is a highly ordered structure, unlike molten globules of many small proteins. The value of deltaCp per mol-residue for the unfolding of the native state is among the highest currently known for any protein.  相似文献   

8.
Chen YR  Clark AC 《Biochemistry》2003,42(20):6310-6320
We have characterized the equilibrium and kinetic folding of a unique protein domain, caspase recruitment domain (CARD), of the RIP-like interacting CLARP kinase (RICK) (RICK-CARD), which adopts a alpha-helical Greek key fold. At equilibrium, the folding of RICK-CARD is well described by a two-state mechanism representing the native and unfolded ensembles. The protein is marginally stable, with a DeltaG(H)()2(O) of 3.0 +/- 0.15 kcal/mol and an m-value of 1.27 +/- 0.06 kcal mol(-1) M(-1) (30 mM Tris-HCl, pH 8, 1 mM DTT, 25 degrees C). While the m-value is constant, the protein stability decreases in the presence of moderate salt concentrations (below 200 mM) and then increases at higher salt concentrations. The results suggest that electrostatic interactions are stabilizing in the native protein, and the favorable Coulombic interactions are reduced at low ionic strength. Above 200 mM salt, the results are consistent with Hofmeister effects. The unfolding pathway of RICK-CARD is complex and contains at least three non-native conformations. The refolding pathway of RICK-CARD also is complex, and the data suggest that the unfolded protein folds via two intermediate conformations prior to reaching the native state. Overall, the data suggest the presence of kinetically trapped, or misfolded, species that are on-pathway both in refolding and in unfolding.  相似文献   

9.
Atropinesterase from Pseudomonas putida has been investigated by means of different ultracentrifugation methods under native and denaturing conditions. The following quantities were determined: sedimentation coefficient, translational diffusion and friction coefficient, partial specific volume and molecular weight. From these data the size, shape and hydration of the enzyme molecule in solution were estimated. The results suggest that atropinesterase is a globular protein which consists of a single polypeptide chain with a molecular weight of about 30,000. In solution under non-denaturing conditions, it occurs mainly as a dimer which hydrodynamically behaves as a rigid impenetrable particle. Calculations based on the spheroid model indicate that this particle resembles a hydrated sphere with a diameter of 6.1 +/- 0.2 nm and a hydration of 0.4 +/- 0.1 g of H2O/g of protein rather than a significantly less hydrated ellipsoid of revolution. Under denaturing conditions dissociation into monomers takes place. The effects of sodium dodecyl sulphate (SDS) on size and shape suggest that dimerization results from side-by-side association of two elongated monomers rather than from end-to-end association. Approximately 57 molecules of SDS are bound per dimer before dissociation occurs concomitant with the additional binding of about 19 molecules of detergent.  相似文献   

10.
In the presence of methylamine and dilute lauryl sulfate (pH 8.0), the human C3 and C4 complement proteins dimerize almost completely. Under these conditions, the related complement protein C5 does not show any tendency to form dimers. This is shown by x-ray and neutron scattering at 9 degrees C and 0.15 M ionic strength. The radii of gyration of the C3 and C4 dimers are very similar, 7.7 and 7.4 nm, and the cross-sectional radii of gyration are the same, 3.4 nm. The scattering curves of the C3 and C4 dimers as well as their Fourier transforms, the p(r)-curves, can be explained by scattering from a model consisting of an elongated elliptic cylinder with semiaxes 6.5 and 2.1 nm and length of 23 nm. This elongated elliptic cylinder model is consistent with the elliptic cylinder model of C4 (Osterberg, R., Eggertsen, G., Lundwall, A., and Sj?quist, J. (1984) Int. J. Biol. Macromol. 6, 195-198) provided that the protein molecules dimerize via their cross-sectional surfaces. Also, the model is consistent with the model of the related protein, alpha 2-macroglobulin, where the four subunits are supposed to form pairwise dimers of an elliptic cylindrical form (Osterberg, R., and Malmensten, B. (1984) Eur. J. Biochem. 143, 541-544).  相似文献   

11.
Members of the spectrin protein family can be found in many different cells and organisms. In all cases studied, the major functional role of these proteins is believed to be structural rather than enzymatic. All spectrin proteins are highly elongated and consist mainly of homologous repeats that constitute rigid segments connected in tandem. It is commonly believed that the details of the spectrin function depend critically on the flexibility of the links between the segments. Here we report on a work addressing this question by studying the transient electric birefringence of recombinant spectrin fragments consisting of segments 14, 14-15, 14-16, and 14-17, respectively, from Drosophila alpha-spectrin. Transient electric birefringence depends sharply on both molecular length and flexibility. We found that the birefringence relaxation time of segment 14 measured at 4 degrees C, but scaled to what is expected at 20 degrees C, equals 16 ns (+/-15%) at pH 7.5 and ionic strength 6 mM. This is consistent with this single segment being rigid, 5 nm long and having an axial ratio equal to about two. Under the same conditions, segments 14-15, 14-16 and 14-17 show relaxation times of 45, 39 and 164 ns (all +/-20%), respectively, scaled to what is expected at 20 degrees C. When the temperature is increased to 37 degrees C the main relaxation time for each of these multisegment fragments, scaled to what is expected at 20 degrees C, increased to 46, 80, and 229 ns (all +/-20%), respectively. When the ionic strength and the Debye shielding is low, the dynamics of these short fragments even at physiological temperature is nearly the same as for fully extended weakly bending rods with the same lengths and axial ratios. When the ionic strength is increased to 85 mM, the main relaxation time for each of these multisegment fragments is reduced 20-50% which suggests that at physiological salt and temperature conditions the links in 2-4-segment-long fragments exhibit significant thermally induced flexing. Provided that the recombinant spectrin fragments can serve as a model for native spectrin, this implies that, at physiological conditions, the overall conformational dynamics of a native spectrin protein containing 20-40 segments equals that of a flexible polymer.  相似文献   

12.
Higher-order structures of chromatin in solution.   总被引:4,自引:0,他引:4  
Neutron scatter studies have been made on gently prepared chicken erythrocyte chromatin over a range of ionic strength. At low ionic strength the mass per unit length of the '10 nm nucleofilament corresponds to one nucleosome per 8--12 nm and a DNA packing ratio of between 6 and 9. From the contrast dependence of the cross-section radius of gyration of the nucleofilament the following parameters have been obtained; RgDNA' the cross-section radius of gyration (Rg) when DNA dominates the scatter; RgP, the cross-section Rg when protein dominates the scatter; Rc, the cross-section Rg at infinite contrast and alpha, the constant which describes the dependence of the cross-section Rg on contrast variation. From our understanding of the structure of the core particle, various arrangement of core particles in the nucleofilament have been tested. In models consistent with the above parameters the core particles are arranged edge-to-edge or with the faces of the core particles inclined to within 20 degrees to the axis of the nucleofilament. With increase of ionic strength the transition to the second-order chromatin structure has been followed. This gave the interesting result that above 20 microM NaCL or 0.4 mM MgCL2 the cross-section Rg increases abruptly to about 9 nm with a packing ratio of 0.2 nucleosome/mn and with further increase of ionic strength the Rg increases to 9.5 nm while the packing ratio increases threefold to 0.6 nucleosome/nm. This suggests a family of supercoils of nucleosomes which contract with increasing ionic strength. In its most contracted form the diameter of the hydrated supercoil has been found from the radial distribution function to be 34 nm. Models for the arrangements of core particles in the 34-nm supercoil are discussed.  相似文献   

13.
Agar-gelatin complex coacervates are studied by small angle neutron scattering (SANS), rheology (in both flow and temperature scan modes) and differential scanning calorimetry (DSC) in order to probe the microscopic structure of this dense protein-polysaccharide-rich phase. DSC and isochronal temperature sweep (rheology) experiments yielded a characteristic temperature at approximately 35+/-2 degrees C. Rheology data revealed a second characteristic temperature at approximately 75+/-5 degrees C which was absent in DSC thermograms. In the flow mode, shear viscosity (eta) was found to scale with (Carreau model) applied shear rate (gamma ) as: eta(gamma ) approximately (gamma )(-k) with k=1.2+/-0.2 indicating non-Newtonian and shear-thinning features independent of ionic strength. The static structure factor S(q) deduced from SANS data in the low wave vector (0.018 A(-1)相似文献   

14.
Immune complex-induced C1 activation and fluid phase C1 autoactivation have been compared in order to elucidate the immune complex role in the C1 activation process. Kinetic analyses revealed that immune complex-bound C1 activates seven times faster than fluid phase C1 spontaneously activates. The rate of spontaneous C1 activation increased after decreasing the solution ionic strength. In fact at one-half physiologic ionic strength (i.e., 0.08 M), the kinetics of spontaneous C1 activation were indistinguishable from the kinetics of activation of immune complex-bound C1 at physiologic ionic strength. The enhanced fluid phase C1 activation at low ionic strength resulted neither from C1 nor C1q aggregation, nor from selective effects on the C1r2S2 subunit; however, at the reduced ionic strength, the C1 association constant (defined for C1q + C1r2S2 in equilibrium C1qr2S2) did increase to 2.3 X 10(8) M-1, which is equal to that for C1 bound to an immune complex at physiologic ionic strength. Therefore, C1 can spontaneously activate in the fluid phase as rapidly as C1 on an immune complex when the strength of interaction between C1q and C1r2S2 is the same in both systems. In conclusion, under physiologic conditions, C1q and C1r2S2 are two weakly interacting proteins. Immune complexes provide a site for the assembly of a stable C1 complex, in which C1q and C1r2S2 remain associated long enough for C1q to activate C1r2S2. Thus, immune complexes enhance the intrinsic C1 autoactivation process by strengthening the association of C1q with C1r2S2.  相似文献   

15.
A procedure for the isolation of the human complement (C) protein C9 is described. The procedure allowin. The purified protein has the electrophoretic mobility of an alpha-globulin, and is a single polypeptide chain with a m.w. of 71,000. No impurities were detected either on gel electrophoretic or immunochemical examination. C9 is a glycoprotein containing 7.8% carbohydrate, and in terms of residues per mole, 3.0 glucosamine, 17.6 neutral hexose, and 7.4 sialic acid. Its amino acid composition is typical of a globular serum protein. Upon automated Edman degradation of reduced and alkylated C9, no amino acid residues were released, suggesting a blocked N-terminus. The concentration of C9 in normal human serum is 58 +/- 8 microgram/ml. A high titer rabbit antiserum was produced and employed to immunochemically deplete serum of C9. The CH50 of the C9-depleted serum was identical to that of whole human serum; however, membrane fragments of erythrocytes lysed by C9-depleted serum lacked the typical ultrastructural C lesions, which constitute the dimeric membrane attack complex.  相似文献   

16.
In cartilage proteoglycan aggregates, link protein stabilizes the binding of proteoglycan monomers to hyaluronate by binding simultaneously to hyaluronate and to the G1 globular domain of proteoglycan monomer core protein. Studies reported here involving metal chelate affinity chromatography demonstrate that link protein is a metalloprotein that binds Zn2+, Ni2+, and Co2+. Zn2+ and Ni2+ decrease the solubility of link protein and result in its precipitation. However, link protein is readily soluble and functional in low ionic strength solvents from which divalent cations have been removed with Chelex 100. These observations make it possible to study the biochemical properties of link protein in low ionic strength, physiologic solvents. Studies were carried out to define the oligomeric state of link protein alone in physiologic solvents, and the transformation in oligomeric state that occurs when link protein binds hyaluronate. Sedimentation equilibrium studies demonstrate that in 0.15 M NaCl, 5 mM EDTA, 50 mM Tris, pH 7, link protein exists as a monomer-hexamer equilibrium controlled by a formation constant of 2 x 10(27) M-5, yielding a delta G' of -36 kcal/mol for the formation of the hexamer from six monomers. On binding hyaluronate oligosaccharides (HA10 or HA12), link protein dissociates to dimer. Link protein hexamer is rendered insoluble by Zn2+. Greater than 90% of the protein is precipitated by 2 mol of Zn2+/mol of link protein monomer. The binding of hyaluronate oligosaccharide by link protein strongly inhibits the precipitation of link protein by Zn2+. The link protein/hyaluronate oligosaccharide complex is completely soluble in the presence of 2 mol of Zn2+/mol of link protein. At higher molar ratios of Zn2+/link protein, the inhibitory effect of hyaluronate oligosaccharide on the precipitation of link protein is gradually overcome. Hyaluronate oligosaccharide is not dissociated from link protein by Zn2+. Hyaluronate remains bound to the link protein which is precipitated by Zn2+, or to the link protein which binds to Zn2(+)-charged iminodiacetate-Sepharose columns. Hyaluronate oligosaccharides and Zn2+ bind to different sites on link protein.  相似文献   

17.
In regenerating rat liver, an elevated protein kinase activity was detected which phosphorylated ribosomal protein S6 and histones. The properties of this enzyme were closely similar with those of protease-activated protein kinase C with Mr 45,000. During the study of the mechanism of proteolytic activation, type III protein kinase C (encoding alpha-sequence) was shown to be subjected to limited proteolysis by trypsin-like protease and converted to protein kinase M in ionic strength- and pH-dependent manner. This reaction was stimulated in the presence of Ca2+ and phospholipid under slightly higher ionic strength condition than physiological level (greater than 140 mM NaCl) and alkaline pH (7.5-8.0). These results suggest that activation of Na+/H+ exchanger in plasma membrane may trigger this type of proteolytic activation of protein kinase C. In addition to protein kinase M, another type of protease-activated kinase with Mr 80,000 was detected when limited proteolysis of protein kinase C was performed on inactive form of this enzyme (in the absence of either Ca2+ or phospholipid or both activators) under lower ionic strength condition. The molecular mass of this active enzyme was slightly smaller (approximately 200) than that of native protein kinase C. However, it is not clear at this time whether this small fragment was released from amino-terminal or carboxy-terminal domain to make protein kinase C partially active in the absence of Ca2+ and phospholipid. Although it has been proposed that proteolytic degradation of protein kinase C is involved in down regulation of this enzyme, the physiological significance of these two types of protease-activated forms of protein kinases in liver has remained obscure.  相似文献   

18.
CAP-23/NAP-22, a neuron-specific protein kinase C substrate, is Nalpha-myristoylated and interacts with calmodulin (CaM) in the presence of Ca2+ ions. Takasaki et al. (1999, J Biol Chem 274:11848-11853) have recently found that the myristoylated N-terminal nonapeptide of CAP-23/NAP-22 (mC/N9) binds to Ca2+ -bound CaM (Ca2+/CaM). In the present study, small-angle X-ray scattering was used to investigate structural changes of Ca2+/CaM induced by its binding to mC/N9 in solution. The binding of one mC/N9 molecule induced an insignificant structural change in Ca2+/CaM. The 1:1 complex appeared to retain the extended conformation much like that of Ca2+/CaM in isolation. However, it could be seen that the binding of two mC/N9 molecules induced a drastic structural change in Ca2+/CaM, followed by a slight structural change by the binding of more than two but less than four mC/N9 molecules. Under the saturated condition (the molar ratio of 1:4), the radius of gyration (Rg) for the Ca2+/CaM-mC/N9 complex was 19.8 +/- 0.3 A. This value was significantly smaller than that of Ca2+/CaM (21.9 +/- 0.3 A), which adopted a dumbbell structure and was conversely 2-3 A larger than those of the complexes of Ca2+/CaM with the nonmyristoylated target peptides of myosin light chain kinase or CaM kinase II, which adopted a compact globular structure. The pair distance distribution function had no shoulder peak at around 40 A, which was mainly due to the dumbbell structure. These results suggest that Ca2+/CaM interacts with Nalpha-myristoylated CAP-23/NAP-22 differently than it does with other nonmyristoylated target proteins. The N-terminal amino acid sequence alignment of CAP-23/NAP-22 and other myristoylated proteins suggests that the protein myristoylation plays important roles not only in the binding of CAP-23/NAP-22 to Ca2+/CaM, but also in the protein-protein interactions related to other myristoylated proteins.  相似文献   

19.
The intimate structure of the complexes located at the sites of DNA loops attachment to the nuclear skeleton was analysed. It is shown that: there are at least three components of the attachment site complex: DNA, protein, RNA; protein moiety consists of 7-8 species with Mr 70-17 kDa. Their association with DNA is resistant to ionic detergents, high salt and urea treatments. The DNA-protein complex is also resistant to the SDS-pronase-phenol deproteinisation procedure; the buoyant density of the complex is the same as DNA density. RNase digestion at low ionic strength reduces density of the complex while the same treatment at 0,4 M NaCl has no effect; DNA-protein complexes isolated with urea-high salt treatment are visualised as globular particles 25-35 nm in diameter with DNA loops attached. These particles were not observed after detergent treatment although protein composition of the complex remained the same.  相似文献   

20.
The bis(terpyridine)cobalt(II), Co(terpy)2(2+), reduction of cytochrome c peroxidase compound I, CcP-I, has been investigated using stopped-flow techniques as a function of ionic strength in pH 7.5 buffers at 25 degrees C. Co(terpy)2(2+) initially reduces the Trp191 radical site in CcP-I with an apparent second-order rate constant, k2, equal to 6.0+/-0.4x10(6) M(-1)s(-1) at 0.01 M ionic strength. A pseudo-first-order rate constant of 480 s(-1) was observed for the reduction of CcP-I by 79 microM Co(terpy)2(2+) at 0.01 M ionic strength. The one-electron reduction of CcP-I produces a second enzyme intermediate, CcP compound II (CcP-II), which contains an oxyferryl, Fe(IV), heme. Reduction of the Fe(IV) heme in CcP-II by Co(terpy)2(2+) shows saturation kinetics with a maximum observed rate constant, k3max, of 24+/-2 s(-1) at 0.01 M ionic strength. At low reductant concentrations, the apparent second-order rate constant for Co(terpy)2(2+) reduction of CcP-II, k3, is 1.2+/-0.5x10(6) M(-1) s-1. All three rate constants decrease with increasing ionic strength. At 0.10 M ionic strength, values of k2, k3, and k3max decrease to 6.0+/-0.8x10(5) M(-1) s(-1), 1.2+/-0.5x10(5) M(-1) s(-1), and 11+/-3 s(-1), respectively. Both the product, Co(terpy)2(3+), and ferricytochrome c inhibit the rate of Co(terpy)2(2+) reduction of CcP-I and CcP-II. Gel-filtration studies show that a minimum of two Co(terpy)2(3+) molecules bind to the native enzyme in low ionic strength buffers.  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