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 15 毫秒
1.
Binding of the 58 kDa monomer and 44 kDa alpha beta dimer forms of terminal deoxynucleotidyl transferase to double stranded DNA was demonstrated by gel retardation and tryptophan fluorescence quenching. The dissociation constants and cooperativity parameters were similar to those that have been determined for binding of these two forms of terminal transferase to single stranded DNA. However, the double stranded DNA binding site size of 10 nucleotides was half the size expected. The efficacy of blunt ended DNA as an initiator in the polymerization reaction catalyzed by terminal transferase was demonstrated by radiometric assays and product analyses on agarose gels. The initial reaction kinetics indicated that dGTP but not dATP was added efficiently to a blunt double stranded DNA 3' end. These results are correlated with current models for in vivo terminal transferase function.  相似文献   

2.
M A Griep  C S McHenry 《Biochemistry》1988,27(14):5210-5215
The beta subunit of Escherichia coli DNA polymerase III holoenzyme binds Mg2+. Reacting beta with fluoresceinmaleimide (FM) resulted in one label per beta monomer with full retention of activity. Titration of FM-beta with Mg2+ resulted in a saturable 11% fluorescence enhancement. Analysis indicated that there was one noncooperative magnesium binding site per beta monomer with a dissociation constant of 1.7 mM. Saturable fluorescence enhancement was also observed when titration was with Ca2+ or spermidine(3+) but not with the monovalent cations Na+ and K+. The Mg2+-induced fluorescence enhancement was specific for FM-beta and was not observed with FM-glutathione, dimethoxystilbenemaleimide-beta, or pyrenylmaleimide-beta. Gel filtration studies indicated that the beta dimer-monomer dissociation occurred at physiologically significant beta concentrations and that the presence of 10 mM Mg2+ shifted the dimer-monomer equilibrium to favor monomers. Both the gel-filtered dimers and the gel-filtered monomers were active in the replication assay. These and other results suggested that the fluorescence increase which accompanies beta dissociation is due to a relief from homoquenching of FM when the beta dimer dissociates into monomers.  相似文献   

3.
The alpha 3 beta 3 complex of ATP synthase obtained from a thermophilic bacterium PS3 was isolated and found to show the ATPase activity (Kagawa, Y., Ohta, S., and Otawara-Hamamoto, Y. (1989) FEBS Lett. 249, 67-69). The structure and the nucleotide binding effects of the alpha 3 beta 3 complex were investigated by means of small-angle x-ray scattering and high performance liquid chromatography. The scattering profile from the alpha 3 beta 3 complex was explained with a model in which the complex is made of an ellipsoid of revolution with the axes of 121.8, 121.8, and 72.0 A having an elliptical hollow cavity with the axes of 35.4, 35.4, and 72.0 A. By the addition of Mg.AT(D)P, significant changes in the scattering profile were observed, in which the radius of gyration decreased from 44 to 35 A. This change was found by gel filtration to be caused by the dissociation reaction from the alpha 3 beta 3 hexamer to the alpha beta dimer. The dissociation of the alpha 3 beta 3 complex was not induced by unhydrolyzable ATP analogue, nor by Pi, Mg2+, and Pi + Mg2+. The structure of the dimer was well explained by the triaxial ellipsoidal model with the axes of 105.2, 39.4, and 108.2 A. The dissociation into the dimer is considered to be related to the ATPase activity because the AT(D)P-induced dissociation is observed only in the presence of Mg2+ ions.  相似文献   

4.
Terminal deoxynucleotidyltransferase activity is absolutely dependent on the presence of a divalent cation in the reaction mixture. This requirement can be satisfied by either Mg2+, Co2+, or Mn2+. When Mg2+ is used, the reaction rate is inhibited by metal ligands, and this inhibition can be reversed by Zn2+. Reaction rates in Mg2+ are also stimulated by the addition of micromolar amounts of Zn2+. To examine the role of Zn2+ in terminal transferase catalysis we analyzed for Zn2+ in homogeneous recombinant human terminal transferase preparations and found that Zn2+ is not an intrinsic part of enzyme molecule. Analysis of Zn2+ binding to terminal transferase under equilibrium conditions shows about 0.3 g of atom of Zn2+/mol of enzyme, suggesting that Zn2+ forms an easily dissociable complex with the enzyme molecule. Kinetic analyses showed that the stimulatory effect of Zn2+ is observed in several buffer systems. Zn2+ increases the affinity of the enzyme for the initiator about 2-fold and decreases affinity for dATP more than 10-fold, resulting in an increase in the apparent Vmax of the reaction. Using a 3'-ended 2',3'-dideoxyoligonucleotide as an inhibitor demonstrates that the inhibitor has no effect on the reaction rate in the absence of Zn2+ but is competitive with respect to the initiator in the presence of Zn2+. These results suggest that Zn2+ is a positive effector for terminal transferase, interacting with oligonucleotide and enzyme near the initiator binding site. Binding of Zn2+ to the enzyme appears to induce conformational changes that greatly increase the Vmax of the reaction with a concomitant decrease in the affinity of the enzyme for dNTP.  相似文献   

5.
Radioiodinated human choriogonadotropin was affinity-cross-linked with a cleavable (nondisulfide) homobifunctional reagent to the hormone receptor on porcine granulosa cells and the solubilized sample was electrophoresed. Cross-linked samples revealed four additional bands of slower electrophoretic mobility in addition to the hormone alpha, beta, and alpha beta dimer bands. The four bands corresponded to masses of 68, 74, 102, and 136 kDa whereas the alpha beta dimer band corresponded to 50 kDa. Formation of the four bands requires the 125I-hormone to bind specifically to the receptor with subsequent cross-linking. Binding can be prevented by excess of native hormone but not by follitropin. A monofunctional analog of the cross-linking reagent failed to produce the four bands. They were also produced by cross-linking Triton X-100-solubilized hormone-receptor complexes. Reagent concentration-dependent cross-linking revealed that their formation was sequential; smaller complexes formed first and then larger ones. When gels of the cross-linked sample were treated with reagents that cleave covalent cross-links and then electrophoresed in a second dimension gel, 18-, 24-, 28-, and 34-kDa components were released, in addition to the alpha and beta subunits of the native hormone. Simultaneous peptide mapping of the cross-linked complexes in the gel matrix with Staphylococcus V8 protease or papain revealed progressive proteolysis to generate terminal fragments of 30 or 27 kDa, respectively. These fragments were unique to and commonly present in the 74-, 102-, and 136-kDa hormone-receptor complexes but were not produced by proteolysis of the cross-linked human choriogonadotropin (hCG) alpha beta dimer or the hCG alpha subunit. Apparently, the radioactively labeled segment(s) of the alpha subunit of 125I-hCG was cross-linked to the 24-kDa component. The results demonstrate the protein nature of the receptor and suggest that 125I-hCG was initially cross-linked to the 24-kDa component to generate the 74-kDa complex, then the 28- and 34-kDa components were sequentially cross-linked to the 24-kDa component in the 74-kDa complex to generate the 102- and 134-kDa complexes.  相似文献   

6.
Mg2+ interacts with the alpha subunits of guanine nucleotide-binding regulatory proteins (G proteins) in the presence of guanosine-5'-[gamma-thio]triphosphate (GTP-gamma S) to form a highly fluorescent complex from which nucleotide dissociates very slowly. The apparent Kd for interaction of G alpha X GTP gamma S with Mg2+ is approximately 5 nM, similar to the Km for G protein GTPase activity X G beta gamma increases the rate of dissociation of GTP gamma S from G alpha X GTP gamma S or G alpha X GTP gamma S X Mg2+ at low concentrations of Mg2+. When the concentration of Mg2+ exceeds 1 mM, G beta gamma dissociates from G beta gamma X G alpha X GTP gamma S X Mg2+. Compared with the dramatic effect of Mg2+ on binding of GTP gamma S to G alpha, the metal has relatively little effect on the binding of GDP. However, G beta gamma increases the affinity of G alpha for GDP by more than 100-fold. High concentrations of Mg2+ promote the dissociation of GDP from G beta gamma X G alpha X GDP, apparently without causing subunit dissociation. The steady-state rate of GTP hydrolysis is strictly correlated with the rate of dissociation of GDP from G alpha under all conditions examined. Thus, there are at least two sites for interaction of Mg2+ with G protein-nucleotide complexes. Furthermore, binding of G beta gamma and GTP gamma S to G alpha is negatively cooperative, while the binding interaction between G beta gamma and GDP is strongly positive.  相似文献   

7.
The catalytic site of Escherichia coli F1 was probed using a reactive ATP analogue, adenosine triphosphopyridoxal (AP3-PL). For complete loss of enzyme activity, about 1 mol of AP3-PL bound to 1 mol of F1 was estimated to be required in the presence or absence of Mg2+. About 70% of the label was bound to the alpha subunit and the rest to the beta subunit in the absence of Mg2+, and the alpha Lys-201 and beta Lys-155 residues, respectively, were the major target residues (Tagaya, M., Noumi, T., Nakano, K., Futai, M., and Fukui, T. (1988) FEBS Lett. 233, 347-351). Addition of Mg2+ decreased the AP3-PL concentration required for half-maximal inhibition, and predominant labeling of the beta subunit (beta Lys-155 and beta Lys-201) with the reagent. ATP and ADP were protective ligands in the presence and absence of Mg2+. The alpha subunit mutation (alpha Lys-201----Gln or alpha Lys-201 deletion) were active in oxidative phosphorylation. However, purified mutant F1s showed impaired low multi-site activity, although their uni-site catalyses were essentially normal. Thus alpha Lys-201 is not a catalytic residue, but may be important for catalytic cooperativity. Mutant F1s were inhibited less by AP3-PL in the absence of Mg2+, and consistent with this, modifications of their alpha subunits by AP3-PL were reduced. AP3-PL was more inhibitory to the mutant enzymes in the presence of Mg2+, and bound to the beta Lys-155 and beta Lys-201 residues of mutant F1 (alpha Lys-201----Gln). These results strongly suggest that alpha Lys-201, beta Lys-155, and beta Lys-201 are located close together near the gamma-phosphate group of ATP bound to the catalytic site, and that the two beta residues and the gamma-phosphate group become closer to each other in the presence of Mg2+.  相似文献   

8.
EF-hand peptides have been shown to bind calcium and dimerize to form an intact protein domain [Shaw, G.S., Hodges, R.S. & Sykes, B.D. (1990). Science, 249, 280-283]. A synthetic 33-residue EF-hand peptide with the sequence of carp parvalbumin CD site demonstrated a seven-fold increase in the apparent calcium dissociation constant with a eight-fold decrease in peptide concentration when fit to a single-site calcium-binding model. This observation is consistent with EF-hand dimerization. This paper describes a method to determine the dimerization dissociation constant and the calcium dissociation constants for both the monomer and dimer forms of this EF-hand peptide using circular dichroism techniques. By monitoring the increase in negative molar ellipticity at 222 nm with increasing peptide concentration under calcium-saturating conditions the dimerization dissociation constant for the synthetic parvalbumin CD site was determined to be 55.68+/-10.76 microM. Using the dimerization constant, the calcium dissociation constants for both the monomer and dimer forms of this peptide were determined by monitoring the change in ellipticity of peptide solutions on addition of increasing amounts of calcium. A fit of this data to a mathematical model that takes into account dimerization results in calcium dissociation constants of 421.3+/-21.56 and 47.06+/-6.72 microM for the monomer and dimer forms, respectively.  相似文献   

9.
The guanine nucleotide binding properties of rap1 protein purified from human neutrophils were examined using both the protein kinase A-phosphorylated and the non-phosphorylated forms of the protein. Binding of GTP[S] (guanosine 5'-[gamma-thio]triphosphate) or GDP was found to be slow in the presence of free Mg2+, but very rapid in the absence of Mg2+. The binding of guanine nucleotides was found to correlate with the loss of endogenous nucleotide from the rap1 protein, which was rapid in the absence of Mg2+. The relative affinities of GTP and GDP for the binding site on rap1 were modulated by the presence of Mg2+, with a preferential affinity (approx. 15-fold) for GTP observed only in the absence of this bivalent cation. The dissociation of GDP from rap1 was not affected by the G-protein beta/gamma-subunit complex. Phosphorylation of rap1 in vitro by protein kinase A did not modify any of the observed nucleotide-binding parameters. Furthermore, the ability of a cytosolic rap1 GTPase-activating protein to stimulate neutrophil rap1 GTP hydrolysis was not modified by phosphorylation. These data suggest that the activation of rap in vivo may be regulated by the release of endogenous GDP, but that phosphorylation by protein kinase A does not affect guanine nucleotide binding or hydrolysis.  相似文献   

10.
Glyoxalase I was purified to homogeneity from bovine brain using affinity chromatography on S-hexylglutathione-Sepharose 6B with a yield of 22%. The enzyme was a dimer (44,000 Daltons) composed of, apparently, identical subunits (22,000 Daltons), as shown by SDS electrophoresis, and contained one mole of Zn2+/monomer. The active site metal ion, Zn2+, was removed by dialysis against EDTA, but the activity of the apoenzyme obtained was not completely restored after addition of Co2+ and Zn2+ (<25%), while a recovery of 50% was obtained after addition of Mg2+. The enzyme was inhibited by S-bromobenzylglutathione and S-p-nitrobenzylglutathione with a Ki value of 21 microM and 32 microM, respectively. The highest dissociation constant observed for the brain enzyme with respect to that reported for human erythrocytes, or other mammalian forms of enzyme could be related to a tissue-specific dependence of the glyoxalase I activity.  相似文献   

11.
We have previously reported that rabbit skeletal muscle phosphorylase kinase is phosphorylated by glycogen synthase (casein) kinase-1 (CK-1) primarily on the beta subunit (beta = 1 mol of PO4; alpha = 0.2 mol of PO4) when the reaction was carried out in beta-glycerophosphate. The resultant enzyme activation was 16-fold (Singh, T. J., Akatsuka, A., and Huang, K.-P. (1982) J. Biol. Chem. 257, 13379-13384). In the present study we found that in Tris-Cl buffer CK-1 catalyzes the incorporation of greater than 2 mol of PO4/monomer into each of the alpha and beta subunits. Phosphorylase kinase activation resulting from the higher level of phosphorylation remained 16-fold. 32P-Labeled tryptic peptides from the alpha and beta subunits were analyzed by isoelectric focusing. Cyclic AMP-dependent protein kinase (A-kinase) phosphorylates a single major site in each of the alpha and beta subunits at 1.5 mM Mg2+. In addition to these two sites, A-kinase phosphorylates at least three other sites in the alpha subunit at 10 mM Mg2+. CK-1 also catalyzes the phosphorylation of multiple sites in both the alpha and beta subunits. Of the two major sites phosphorylated by CK-1 in the beta subunit, one of these sites is also recognized by A-kinase. At least three sites are phosphorylated by CK-1 in the alpha subunit. One of these sites is recognized by CK-1 only after a prior phosphorylation of phosphorylase kinase by A-kinase at a single site in each of the alpha and beta subunits at 1.5 mM Mg2+. The roles of the different phosphorylation sites in phosphorylase kinase activation are discussed.  相似文献   

12.
Tubulin dimer dissociation and proteolytic accessibility   总被引:2,自引:0,他引:2  
The alpha and beta subunits of the tubulin dimer each possess a distal C-terminal subtilisin cleavage site which, when cleaved, releases an acidic, small peptide. In addition, each possesses an internal site, cleaved by trypsin in alpha and chymotrypsin in beta, which connects the amino and carboxyl structural domains. A model of the dimer is presented which suggests that the beta C-terminal subtilisin site may be more accessible in the monomer than in the dimer. Kinetics of cleavage at this site on the dimer yield straight-line plots of log (undigested fraction) versus time, from which pseudo-first-order rate constants are obtained. Temperature effects on the rate constant are due to changes in the activity of subtilisin, not to temperature-induced unfolding around this site. The rate constant is proportional to the subtilisin/tubulin ratio, whether this is varied by changing the concentration of subtilisin or of tubulin. However, if the rate constant increases due to decreasing tubulin concentration, the extrapolated zero time intercept decreases. The decrease in zero time intercept is interpreted as being due to the appearance of a rapidly digested fraction upon dilution of tubulin. The increase observed in this fast fraction with dilution of tubulin is fully reversible upon reconcentration. It is suggested that this fast fraction represents monomeric beta-tubulin and the concentration dependence of this fast fraction indicates a dissociation constant of about 1.5 X 10(-7) M.  相似文献   

13.
The 90-kDa heat shock protein, HSP90, of the mouse has two isoforms, alpha and beta, which are electrophoretically separable. We have investigated the native forms of HSP90 molecules under physiological conditions and determined their isoform compositions. Analysis by sodium dodecyl sulfate-polyacrylamide gel electrophoresis showed that HSP90 purified from mouse lymphoma L5178Y cells consists of approximately 40% alpha and 60% beta isoforms. Analysis by nondenaturing polyacrylamide gel electrophoresis showed that the purified HSP90 exists predominantly as a dimer, but a considerable amount of monomer was also detected. Western blotting using polyclonal anti-mouse HSP90 antibodies revealed that the native forms of HSP90 in the crude L5178Y cell lysates are also dimer and monomer. The nondenaturing polyacrylamide gel electrophoresis resolved the dimeric forms into two separate bands that were identified as alpha/alpha and beta/beta homodimers by two methods: sodium dodecyl sulfate-polyacrylamide gel electrophoresis and peptide mapping. In addition, the results showed that the monomeric form consists mainly of the beta isoform. Both the alpha and beta isoforms were shown to bind equally to actin filaments.  相似文献   

14.
An alpha beta heterodimer of the F1-ATPase of Rhodospirillum rubrum was isolated by extraction of chromatophores with LiCl. Each alpha beta heterodimer contains one tightly bound ADP, which is released upon removal of medium Mg2+. The dimer can be reversibly dissociated by removal of Mg(2+)-ions. The alpha beta heterodimer restores both ATP-synthetic and -hydrolytic activities to LiCl-treated chromatophores, saturation being achieved at approximately 2 mmol alpha beta.mol BChl-1. The heterodimer itself hydrolyses Mg-ATP with an activity distinct from RF1, being unaffected by azide or sulphite ions. The Vmax and Km (ATP) for this Mg(2+)-dependent activity were 110 +/- 10 nmol.min-1.mg protein-1 and 100 +/- 30 microM, respectively. The Km did not differ significantly from that of RF1.  相似文献   

15.
Protein I from intestinal epithelium is biochemically and immunologically related to the fibroblast 36-kDa substrate of the Rous sarcoma virus-encoded tyrosine protein kinase (Gerke and Weber (1984) EMBO J. 3, 227-233). Protein I is a Ca2+-binding protein containing two copies each of a 36- and 10-kDa subunit. Denaturation/renaturation experiments show that the 36-kDa subunit is a monomer, whereas the 10-kDa subunit forms a dimer. Mixing of the subunits leads to reconstituted protein I. Physicochemical properties of protein I and its isolated subunits reveal a Ca2+-dependent conformational change in the 36-kDa subunit which involves the exposure of 1 or more tyrosine residues to a more aqueous environment. This change points to a Ca2+ binding constant of about 10(4) M-1 in the presence of 2 mM Mg2+ and induces the ability of protein I and the 36-kDa subunit to bind in vitro to F-actin and nonerythroid spectrin. The same high Ca2+ requirement has been reported for the in vitro tyrosine phosphorylation of a 35-kDa protein from A-431 carcinoma cells by the epidermal growth factor receptor kinase (Fava and Cohen (1984) J. Biol. Chem. 259, 2636-2645). Here we show that this 35-kDa substrate is biochemically and immunologically related to the 36-kDa subunit of protein I, which in turn corresponds to the substrate of the Rous sarcoma virus kinase. The protein of A-431 cells exists not only as a monomer but also as a dimer. The latter fraction contains a 10-kDa polypeptide immunologically related to the corresponding subunit of protein I. Given past results on the A-431 system, we speculate that the monomer rather than the dimer is the preferred in vitro substrate for the epidermal growth factor receptor kinase. Thus, the 10-kDa subunit, which induces dimerization of the phosphorylatable large subunit, may act as an inhibitor.  相似文献   

16.
J Pimmer  E Holler 《Biochemistry》1979,18(17):3714-3723
The association of phenylalanylptRNA and Mg2+ follows a biphasic concentration dependence as indicated by the active site directed fluorescent indicator 2-p-toluidinyl-naphthalene-6-sulfonate. The macroscopic dissociation constants are 0.16 +/- 0.03 and 4.1 +/- mM. The effect of Mg2+ on the association of enzyme and MgATP, on the synergistic binding of MgATP and L-phenylalaninol, and on the pre-steady-state synthesis and pyrophosphorolysis of the enzyme-phenylalanyladenylate complex in the absence and the presence of tRNA Phe has been measured by established equilibrium and stopped-flow techniques using 2-p-toluidinylnaphthalene-6-sulfonate. At 10 mM Mg2+, the association of enzyme and MgATP is biphasic with dissociation constants of 0.25 +/- 0.03 and 9.1 +/- 1.7 mM. At 2 mM Mg2+, a single dissociation constant of 5.0 +/- 0.5 mM is indicated. The coupling constant of the synergistic reaction is 15 at 1 mM Mg2+ and 290 at 10 mM Mg2+. The Hill constant of the sigmoidal dependence is 3.6. The strengthening of the synergism is believed to reflect a Mg2+-dependent coupling of the synergistic reactions at the two active sites of the enzyme, the coupling being negligible at 1 mM and maximal at 10 mM Mg2+. The pre-steady-state rate of adenylate synthesis is accelerated by the presence of Mg2+. The effect is to decrease the value of the Michaelis-Menten constant of MgATP. Another effect is to increase the rate constant when tRNA Phe is present. At subsaturating [MgATP], the [Mg2+] dependence of the observed rate constant is hyperbolical in the absence and sigmoidal (Hill constant, 3.5) in the presence of tRNA Phe. The rate of the pyrophosphorolysis is enhanced by a decrease of the Michaelis-Menten constant of MgPPi. The effects on the thermodynamics and kinetics parallel the occupancy of the low-affinity Mg2+-binding sites of the enzyme.  相似文献   

17.
The chromium(III) complex of ATP, an MgATP complex analogue, inactivates (Na+ + K+)-ATPase by forming a stable chromo-phosphointermediate. The rate constant k2 of inactivation at 37 degrees C of the beta, gamma-bidentate of CrATP is enhanced by Na+ (K0.5 = 1.08 mM), imidazole (K0.5 = 15 mM) and Mg2+ (K0.5 = 0.7 mM). These cations did not affect the dissociation constant of the enzyme-chromium-ATP complex. The inactive chromophosphoenzyme is reactivated slowly by high concentrations of Na+ at 37 degrees C. The half-maximal effect on the reactivation was reached at 40 mM NaCl, when the maximally observable reactivation was studied. However, 126 mM NaCl was necessary to see the half-maximal effect on the apparent reactivation velocity constant. K+ ions hindered the reactivation with a Ki of 70 microM. Formation of the chromophosphoenzyme led to a reduction of the Rb+ binding sites and of the capacity to occlude Rb+. The beta, gamma-bidentate of chromium(III)ATP (Kd = 8 microM) had a higher than the alpha, beta, gamma-tridentate of chromium(III)ATP (Kd = 44 microM) or the cobalt tetramine complex of ATP (Kd = 500 microM). The beta, gamma-bidentate of the chromium(III) complex of adenosine 5'-[beta, gamma-methylene]triphosphate also inactivated (Na+ + K+)ATPase. Although CrATP could not support Na+, K+ exchange in everted vesicles prepared from human red blood cells, it supported the Na+-Na+ and Rb+-Rb+ exchange. It is concluded that CrATP opens up Na+ and K+ channels by forming a relatively stable modified enzyme-CrATP complex. This stable complex is also formed in the presence of the chromium complex of adenosine 5'-[beta, gamma-methylene]triphosphate. Because the beta, gamma-bidentate of chromium ATP is recognized better than the alpha, beta, gamma-tridentate, it is concluded that the triphosphate site recognizes MgATP with a straight polyphosphate chain and that the Mg2+ resides between the beta- and the gamma-phosphorus. The enhancement of inactivation by Mg2+ and Na+ may be caused by conformational changes at the triphosphate site.  相似文献   

18.
Canonical glutathione (GSH) transferases are dimeric proteins with subunits composed of an N-terminal GSH binding region (domain 1) and a C-terminal helical region (domain 2). The stabilities of several GSH transferase dimers are dependent upon two groups of interactions between domains 1 and 2 of opposing subunits: a hydrophobic ball-and-socket motif and a buried charge cluster motif. In rGSTM1-1, these motifs involve residues F56 and R81, respectively. The structural basis for the effects of mutating F56 to different residues on dimer stability and function has been reported (Codreanu et al. (2005) Biochemistry 44, 10605-10612). Here, we show that the simultaneous disruption of both motifs in the F56S/R81A mutant causes complete dissociation of the dimer to a monomeric protein on the basis of gel filtration chromatography and multiple-angle laser light scattering. The fluorescence and far-UV CD properties of the double mutant as well as the kinetics of amide H/D exchange along the polypeptide backbone suggest that the monomer has a globular structure that is similar to a single subunit in the native protein. However, the mutant monomer has severely impaired catalytic activity, suggesting that the dimer interface is vital for efficient catalysis. Backbone amide H/D exchange kinetics in the F56S and F56S/R81A mutants indicate that a reorganization of the loop structure between helix alpha2 and strand beta3 near the active site is responsible for the decreased catalytic activity of the monomer. In addition, the junction between the alpha4 and alpha5 helices in F56S/R81R shows decreased H/D exchange, indicating another structural change that may affect catalysis. Although the native subunit interface is important for dimer stability, urea-induced unfolding of the F56S/R81A mutant suggests that the interface is not essential for the thermodynamic stability of individual subunits. The H/D exchange data reveal a possible molecular basis for the folding cooperativity observed between domains 1 and 2.  相似文献   

19.
The 8-kDa light chain of dynein (DLC8) is ubiquitously expressed in various cell types. Other than serving as a light chain of the dynein complexes, this highly conserved protein has been shown to bind a larger number of proteins with diverse biological functions. DLC8 forms a homodimer via three-dimensional domain swapping of an internal beta-strand (the beta2-strand) at neutral pH. The protein undergoes non-reversible dimer-to-monomer dissociation when the pH value of the protein solution decreases. The three-dimensional structure of the DLC8 monomer determined by NMR spectroscopy at pH 3.0 showed that the protein is well folded. The major conformational change accompanied by dimer dissociation is in the beta2-strand of the protein, which undergoes transition from a beta-strand to a nascent alpha-helix. The monomer form of DLC8 is not capable of binding to target proteins. Insertion of two flexible amino acid residues in the tight beta1/beta2-loop dramatically stabilized the monomer conformation of the protein. NMR studies showed that the mutation altered the conformation as well as the three-dimensional domain swapping-mediated assembly of the DLC8 dimer. The mutant DLC8 was unable to bind to its targets even at physiological pH. The three-dimensional structure of the mutant protein in its monomeric form provides the structural basis of the mutation-induced stabilization of the monomer conformation. Based on the experimental data, we conclude that the formation of the beta2-strand swapping-mediated dimer is mandatory for the structure and function of DLC8. We further note that the DLC8 dimer represents a novel mode of three-dimensional domain swapping.  相似文献   

20.
T A Ono  H Mino 《Biochemistry》1999,38(27):8778-8785
Binding of Mn2+ to manganese-depleted photosystem II and electron donation from the bound Mn2+ to an oxidized YZ tyrosine were studied under the same equilibrium conditions. Mn2+ associated with the depleted membranes in a nonsaturating manner when added alone, but only one Mn2+ ion per photosystem II (PS II) was bound to the membranes in the presence of other divalent cations including Ca2+ and Mg2+. Mn2+-dependent electron donation to photosystem II studied by monitoring the decay kinetics of chlorophyll fluorescence and the electron paramagnetic resonance (EPR) signal of an oxidized YZ tyrosine (YZ+) after a single-turnover flash indicated that the binding of only one Mn2+ ion to the manganese-depleted PS II is sufficient for the complete reduction of YZ+ induced by flash excitation. The results indicate that the manganese-depleted membranes have only one unique binding site, which has higher affinity and higher specificity for Mn2+ compared with Mg2+ and Ca2+, and that Mn2+ bound to this unique site can deliver an electron to YZ+ with high efficiency. The dissociation constant for Mn2+ of this site largely depended on pH, suggesting that a single amino acid residue with a pKa value around neutral pH is implicated in the binding of Mn2+. The results are discussed in relation to the photoactivation mechanism that forms the active manganese cluster.  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