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.
D Panda  S Roy  B Bhattacharyya 《Biochemistry》1992,31(40):9709-9716
Concentration-dependent dissociation of dimers of goat brain tubulin S and tubulin was studied by fluorescence anisotropy. Upon dilution, assembly-competent fluorescein 5'-maleimide labeled dimers of tubulin S and tubulin show a progressive decrease in fluorescence anisotropy. That this lowering of anisotropy results from the dissociation of tubulin S dimers into monomers was shown by dilution experiments with unlabeled homologous and heterologous proteins. A nonlinear least-squares fit of the data gave a dissociation constant of 7.1 x 10(-8) M for tubulin S compared to 7.2 x 10(-7) M for tubulin at 25 degrees C in 0.1 M PEM buffer, pH 7.0. van't Hoff plots of dimer-monomer dissociation of tubulin S and tubulin also show considerable differences in delta H and delta S. Effects of ionic strength and colchicine on the equilibrium constants are also substantially different for tubulin and tubulin S. The implications of these observations on the influence of C-terminal tails on tubulin structure are discussed.  相似文献   

2.
Brain tubulin has been conjugated with dichlorotriazinyl- aminofluorescein (DTAF) to form a visualizable complex for the study of tubulin dynamics in living cells. By using several assays we confirm the finding of Keith et al. (Keith, C. H., J. R. Feramisco, and M. Shelanski, 1981, J. Cell Biol., 88:234-240) that DTAF-tubulin polymerizes like control tubulin in vitro. The fluorescein moiety of the complex is readily bleached by the 488-nm line from an argon ion laser. When irradiations are performed over short times (less than 1 s) and in the presence of 2 mM glutathione, a mixture of DTAF-tubulin and control protein (as occurs after microinjection of the fluorescent conjugate into living cells) will retain full polymerization activity. Slow bleaching (approximately 5 min) or bleaching without glutathione promotes formation of covalent cross-links between neighboring polypeptides and kills the polymerization activity of DTAF-tubulin, including some molecules that are neither cross-linked nor bleached. Even under conditions that damage DTAF-tubulin, however, DTAF- microtubules are not destroyed by bleaching. They will continue to elongate by addition of DTAF-tubulin subunits to their free ends, and they neither bind nor exchange subunits along their lateral surfaces. These results suggest that DTAF-tubulin is a suitable analog for tubulin, both in studies of protein incorporation and for investigations of fluorescence redistribution after photobleaching.  相似文献   

3.
B Bhattacharyya  J Wolff 《Biochemistry》1976,15(11):2283-2288
The rate of binding of colchicine to tubulin to tubulin is enhanced by certain anions. Among the inorganic anions tested, only sulfate was effective. The organic anions include mostly dicarboxylic acids, among which tartrate was the most effective. This effect occurs onlt at low concentrations of colchicine (less than 0.6 X 10(-5) M). The rate increase dor sulfate and L-(+)-tartrate is ca. 2.5-fold at 1.0 mM and plateaus at a limiting value of ca. 4-fold at 100mM. The overall dissociation rate of the colchicine from the complex, which includes both the true rate of dissociation and the rate of irreversible denaturation of tubulin, is not influenced by 1.0 mM tartrate. The affinity constants for colchicine determined from the rate constants are 8.7 X 10(6) and 2.1 X 10(7) M-1 in the absence and the presence of 1.0 mM L-(+)-tartrate. The limiting value is 3.2 X 10(7) M-1. The affinity constant calculated from steady-state measurements is 3.2 X 10(6) M-1 with or without anions. The binding of other ligands like podophyllotoxin, vinblastine, and 1 -anilino-8-naphthalenesulfonate to tubulin is not affected by tartrate. No major conformational changes resulting from anion treatment could be detected by circular dichroism or intrinsic fluorescence. However, the ability of tubulin to polymerize is inhibited by L-(+)-tartrate at concentrations that increase the rate of colchicine binding. We conclude that anions must have a local effect at or near the binding site which enhances the binding rate of colchicine and which may be related to inhibition of polymerization.  相似文献   

4.
Guanidine x HCl (GdnHCl)-induced unfolding of tetrameric N(5)-(L-1-carboxyethyl)-L-ornithine synthase (CEOS; 141,300 M(r)) from Lactococcus lactis at pH 7.2 and 25 degrees C occurred in several phases. The enzyme was inactivated at approximately 1 M GdnHCl. A time-, temperature-, and concentration-dependent formation of soluble protein aggregates occurred at 0.5-1.5 M GdnHCl due to an increased exposure of apolar surfaces. A transition from tetramer to unfolded monomer was observed between 2 and 3.5 M GdnHCl (without observable dimer or trimer intermediates), as evidenced by tyrosyl and tryptophanyl fluorescence changes, sulfhydryl group exposure, loss of secondary structure, size-exclusion chromatography, and sedimentation equilibrium data. GdnHCl-induced dissociation and unfolding of tetrameric CEOS was concerted, and yields of reactivated CEOS by dilution from 5 M GdnHCl were improved when unfolding took place on ice rather than at 25 degrees C. Refolding and reconstitution of the enzyme were optimal at 相似文献   

5.
We have used fluorescence correlation spectroscopy to analyze the interaction of GTP-tubulin with rhodamine-labeled RB3, a neural protein of the stathmin family, and to determine the kinetic pathway of the association process. RB3 displayed slow association-dissociation kinetics with tubulin depending on the square of the tubulin concentration. The values of the apparent association and dissociation rate constants of the complex of two tubulin dimers and RB3 are determined to be (3.52+/-0.14)x10(-3) micro;M(-2)/s and (1.9+/-0.6)x10(-3) s(-1) respectively. The value of the equilibrium dissociation constant for the first tubulin-RB3 interaction is estimated to be >or=7 microM at 20 degrees C.  相似文献   

6.
The unfolding and dissociation of the tetrameric enzyme fructose-1,6-bisphosphatase from pig kidney by guanidine hydrochloride have been investigated at equilibrium by monitoring enzyme activity, ANS binding, intrinsic (tyrosine) protein fluorescence, exposure of thiol groups, fluorescence of extrinsic probes (AEDANS, MIANS), and size-exclusion chromatography. The unfolding is a multistate process involving as the first intermediate a catalytically inactive tetramer. The evidence that indicates the existence of this intermediate is as follows: (1) the loss of enzymatic activity and the concomitant increase of ANS binding, at low concentrations of Gdn.HCl (midpoint at 0.75 M), are both protein concentration independent, and (2) the enzyme remains in a tetrameric state at 0.9 M Gdn.HCl as shown by size-exclusion chromatography. At slightly higher Gdn.HCl concentrations the inactive tetramer dissociates to a compact dimer which is prone to aggregate. Further evidence for dissociation of tetramers to dimers and of dimers to monomers comes from the concentration dependence of AEDANS-labeled enzyme anisotropy data. Above 2.3 M Gdn.HCl the change of AEDANS anisotropy is concentration independent, indicative of monomer unfolding, which also is detected by a red shift of MIANS-labeled enzyme emission. At Gdn.HCl concentrations higher than 3.0 M, the protein elutes from the size-exclusion column as a single peak, with a retention volume smaller than that of the native protein, corresponding to the completely unfolded monomer. In the presence of its cofactor Mg(2+), the denaturated enzyme could be successfully reconstituted into the active enzyme with a yield of approximately 70-90%. Refolding kinetic data indicate that rapid refolding and reassociation of the monomers into a nativelike tetramer and reactivation of the tetramer are sequential events, the latter involving slow and small conformational rearrangements in the refolded enzyme.  相似文献   

7.
G R Parr  G G Hammes 《Biochemistry》1976,15(4):857-862
The kinetics of dissociation and reassembly of rabbit skeletal muscle phosphofructokinase has been studied using fluorescence, stopped-flow fluorescence and enzyme activity measurements. The dissociation of the fully active tetramer in 0.8 M guanidine hydrochloride (0.1 M potassium phosphate, pH 8.0) occurs in three kinetic phases as measured by changes in the protein fluorescence emission intensity: dissociation of tetramer to dimer with a relaxation time of a few milliseconds; dissociation of dimer to monomer with a relaxation time of a few seconds; and a conformational change of the monomer with a relaxation time of a few minutes. All three phases exhibit first-order kinetics; ATP (0.05 mM) retards the second step but does not influence the rate of the other two processes. The rate of the second process increases with decreasing temperature; this may be due to the involvement of hydrophobic interactions in the stabilization of the dimeric enzyme. A further unfolding of the monomer polypeptide chain occurs at higher guanidine concentrations, and the relaxation time associated with this process was found to be 83 ms in 2.5 M guanidine, 0.1 M potassium phosphate (pH 8.0) at 23 degrees C. The phosphofructokinase monomers were reassembled from 0.8 M guanidine chloride by 1:10 dilution of the guanidine hydrochloride concentration and yielded a protein with 70-94% of the original activity, depending on the protein concentration. The reactivation process follows second-order kinetics; ATP (5 mM) increases the rate of reactivation without altering the reaction order, while fructose 6-phosphate does not influence the rate of reaction. The rate-determining step is probably the association of monomers to form the dimer.  相似文献   

8.
We describe a system for extending stopped-flow analysis to the kinetics of ligand capture and release by cell surface receptors in living cells. While most mammalian cell lines cannot survive the shear forces associated with turbulent stopped-flow mixing, we determined that a murine hematopoietic precursor cell line, 32D, is capable of surviving rapid mixing using flow rates as great as 4.0 mL/s, allowing rapid processes to be quantitated with dead times as short as 10 ms. 32D cells do not express any endogenous epidermal growth factor (EGF) receptor or other ErbB family members and were used to establish monoclonal cell lines stably expressing the EGF receptor. Association of fluorescein-labeled H22Y-murine EGF (F-EGF) to receptor-expressing 32D cells was observed by measuring time-dependent changes in fluorescence anisotropy following rapid mixing. Dissociation of F-EGF from EGF-receptor-expressing 32D cells was measured both by chase experiments using unlabeled mEGF and by experiments in which equilibrium was perturbed by dilution. Comparison of these dissociation experiments showed that little, if any, ligand-induced dissociation occurs in the chase dissociation experiments. Data from a series of association and dissociation experiments, performed at various concentrations of F-EGF in the nanomolar range and at multiple cell densities, were simultaneously analyzed using global analysis techniques and fit to a two independent receptor-class model. Our analysis is consistent with the presence of two distinct receptor populations having association rate constants of k(on1) = 8.6 x 10(6) M(-1) s(-1) and k(on2) = 2.4 x 10(6) M(-1) s(-1) and dissociation rate constants of k(off1) = 0.17 x 10(-2) s(-1) and k(off2) = 0.21 x 10(-2) s(-1). The magnitudes of these parameters suggest that under physiological conditions, in which cells are transiently exposed to nanomolar concentrations of ligand, ligand capture and release may function as the first line of regulation of the EGF receptor-induced signal transduction cascade.  相似文献   

9.
Microtubule assembly in vivo was studied by hapten-mediated immunocytochemistry. Tubulin was derivatized with dichlorotriazinylaminofluorescein (DTAF) and microinjected into living, interphase mammalian cells. Sites of incorporation were determined at the level of individual microtubules by double-label immunofluorescence. The haptenized tubulin was localized by an anti-fluorescein antibody and a second antibody conjugated with fluorescein. Total microtubules were identified by anti-tubulin and a secondary antibody conjugated with rhodamine. Contrary to recent studies (Salmon, E. D., et al., 1984, J. Cell Biol., 99:2165-2174; Saxton, W. M., et al., 1984, J. Cell Biol., 99:2175-2186) which suggest that tubulin incorporates all along the length of microtubules in vivo, we found that microtubule assembly in interphase cells was in vivo, as in vitro, an end-mediated process. Microtubules that radiated out toward the cell periphery incorporated the DTAF-tubulin solely at their distal, that is, their plus ends. We also found that a proportion of the microtubules connected to the centrosomes incorporated the DTAF-tubulin along their entire length, which suggests that the centrosome can nucleate the formation of new microtubules.  相似文献   

10.
The catalytic properties of two ATPases which had been purified from bovine brain microtubules (Tominaga, S. & Kaziro, Y. (1983) J. Biochem. 93, 1085-1092) were studied. ATPase I, which had a molecular weight of 33,000, required the presence of 1.0 microM tubulin, 0.2 mM Mg2+, and 10 mM Ca2+ for maximal activity. The activation of ATPase I by tubulin was specific to the native form of tubulin, which could not be replaced by F-actin or tubulin denatured either by heat or more mildly by dialysis in the absence of glycerol. ATPase I was not specific to ATP, and GTP, and to a lesser extent, UTP and CTP were also hydrolyzed. Km for ATP of ATPase I was about 0.04 mM. ATPase I was inhibited by 5 mM Mg2+, 0.04 M K+, 10(-3) M vanadate, 10 mM N-ethylmaleimide, or 20% (v/v) glycerol. ATPase II, which was associated with membrane vesicles, required the presence of 0.2-2.0 mM Mg2+ and 20 mM KCl for activity. Tubulin stimulated the reaction of ATPase II only partially, and the addition of Ca2+ was rather inhibitory. ATPase II was specific to ATP with a Km value of 0.14 mM. It was inhibited by 1.6 mM N-ethylmaleimide and 20% (v/v) glycerol, but was not very sensitive to vanadate. Instead, ATPase II was inhibited by trifluoperazine, chlorpromazine, and nicardipin at 10(-3) M.  相似文献   

11.
Excimer-forming cysteines in tubulin are detected by the presence of excimer fluorescence in N-(1-pyrenyl)maleimide-labeled tubulin. The ratio of excimer/monomer fluorescence of labeled protein remained unchanged upon its dilution. These results indicating that both partner of each pair(s) of cysteine are located in the same subunit. The excimer fluorescence is insensitive to prior treatment of tubulin with either colchicine or GTP, indicating that pairs of cysteines protected by those drugs are not involved in excimer formation. This excimer fluorescence of N-(1-pyrenyl)maleimide-labeled tubulin disappeared upon treatment with SDS, guanidinium chloride (GdmCl) and urea. Studies with GdmCl induced unfolding of N-(1-pyrenyl)maleimide-labeled tubulin showed that the loss of excimer fluorescence precedes subunit dissociation. The loss of both colchicine-binding activity and the excimer fluorescence with increasing temperature indicates a major conformational change of the tubulin molecule at elevated temperatures.  相似文献   

12.
Flow cytometry was used to demonstrate the presence of beta-lactoglobulin (betaLG) receptors on living murine hybridoma MARK-3 cells using a fluorescein isothiocyanate-betaLG conjugate (FITC-betaLG: molar ratio of 5:1). A site occupation curve was produced using a shift in the mean channel fluorescence at various concentrations of FITC-betaLG. The binding of labelled ligand was concentration dependent and was inhibited by unlabelled betaLG. The on-rate constant was 3.2x10(2) M(-1) min(-1) and the off-rate constant was 0.002 min(-1). Scatchard plot analysis gave a dissociation constant (K(d)) of 44+/-21x10(-7) and 39+/-24x10(-5) M (n=3). Flow cytometry indicated that at least 15% of the FITC-betaLG were internalized for 5 min and that internalization was temperature- and time-dependent. The internalization was confirmed by 3-D fluorescence microscopy (CELLScan system).  相似文献   

13.
The finding that exchange of tubulin subunits between tubulin dimers (alpha-beta + alpha'beta' <--> alpha'beta + alphabeta') does not occur in the absence of protein cofactors and GTP hydrolysis conflicts with the assumption that pure tubulin dimer and monomer are in rapid equilibrium. This assumption underlies the many physical chemical measurements of the K(d) for dimer dissociation. To resolve this discrepancy we used surface plasmon resonance to determine the rate constant for dimer dissociation. The half-time for dissociation was approximately 9.6 h with tubulin-GTP, 2.4 h with tubulin-GDP, and 1.3 h in the absence of nucleotide. A Kd equal to 10(-11) M was calculated from the measured rate for dissociation and an estimated rate for association. Dimer dissociation was found to be reversible, and dimer formation does not require GTP hydrolysis or folding information from protein cofactors, because 0.2 microM tubulin-GDP incubated for 20 h was eluted as dimer when analyzed by size exclusion chromatography. Because 20 h corresponds to eight half-times for dissociation, only monomer would be present if dissociation were an irreversible reaction and if dimer formation required GTP or protein cofactors. Additional evidence for a 10(-11) M K(d) was obtained from gel exclusion chromatography studies of 0.02-2 nM tubulin-GDP. The slow dissociation of the tubulin dimer suggests that protein tubulin cofactors function to catalyze dimer dissociation, rather than dimer assembly. Assuming N-site-GTP dissociation is from monomer, our results agree with the 16-h half-time for N-site GTP in vitro and 33 h half-life for tubulin N-site-GTP in CHO cells.  相似文献   

14.
The effects of elevated pH and high salt concentrations on tubulin   总被引:1,自引:0,他引:1  
The effects of incubating phosphocellulose-purified bovine tubulin at 4 degrees C in nucleotide-free buffers at alkaline pH or at high concentrations of NaCl, KCl, (NH4)2SO4, or NH4Cl have been studied. At pH greater than or equal to 7.5 or at NaCl concentrations greater than or equal to 0.7 M, tubulin releases bound nucleotides irreversibly and loses, with apparent first-order kinetics, the ability to assemble into microtubules. In 0.1 M 1,4-piperazinediethanesulfonic acid buffer, pH 6.9, in the presence of 1.3 M NH4Cl, tubulin undergoes more rapid loss of capacity to assemble than it does in NaCl and KCl, but 1.3 M (NH4)2SO4 causes no detectable change in tubulin after 1-h incubation. Incubation at high pH or at high neutral salt concentrations also causes an apparently irreversible change in the ultraviolet difference spectrum and in the sedimentation velocity profile of tubulin. At elevated salt concentrations a decrease of approximately 10% in the molar ellipticity within the wavelength range 220-260 nm is observed. The changes that occur during 1-h exposure to pH 8.0 can be completely prevented by including 1 mM guanosine 5'-triphosphate (GTP) or 4 M glycerol in the buffer, but those which occur at pH 9.0 cannot be prevented by these additions. In 1 M NaCl when the ratio of bound guanine nucleotide to tubulin reaches approximately 1.0, tubulin loses the abilities to assemble into microtubules and to bind colchicine. The rate of loss of nucleotide in 2 M NaCl is decreased in the presence of 1 mM GTP, and tubulin is protected almost completely from 1 M NaCl-induced loss of GTP (and retains the ability to exchange [3H]GTP as well) in the presence of bound colchicine. Investigators who anticipate exposing tubulin to buffers of elevated pH or high concentrations of chaotropic salts should be extremely cautious in interpreting the resulting data unless they can demonstrate that irreversible alteration of the protein has not occurred.  相似文献   

15.
The rate of exchange of tubulin that is incorporated into spindle microtubules with dimeric tubulin in the cytoplasm has been measured in sea urchin eggs by studying fluorescence redistribution after photobleaching (FRAP). Dichlorotriazinyl amino fluorescein (DTAF) has been used to label bovine brain tubulin. DTAF-tubulin has been injected into fertilized eggs of Lytechinus variegatus and allowed to equilibrate with the endogenous tubulin pool. Fluorescent spindles formed at the same time that spindles were seen in control eggs, and the injected embryos proceeded through many cycles of division on schedule, suggesting that DTAF-tubulin is a good analogue of tubulin in vivo. A microbeam of argon laser light has been used to bleach parts of the fluorescent spindles, and FRAP has been recorded with a sensitive video camera. Laser bleaching did not affect spindle structure, as seen with polarization optics, nor spindle function, as seen by rate of progress through mitosis, even when one spindle was bleached several times in a single cell cycle. Video image analysis has been used to measure the rate of FRAP and to obtain a low resolution view of the fluorescence redistribution process. The half-time for spindle FRAP is approximately 19 s, even when an entire half-spindle is bleached. Complete exchange of tubulin in nonkinetochore spindle and astral microtubules appeared to occur within 60-80 s at steady state. This rate is too fast to be explained by a simple microtubule end-dependent exchange of tubulin. Efficient microtubule treadmilling would be fast enough, but with current techniques we saw no evidence for movement of the bleached spot during recovery, which we would expect on the basis of Margolis and Wilson's model (Nature (Lond.)., 1981, 293:705)-- fluorescence recovers uniformly. Microtubules may be depolymerizing and repolymerizing rapidly and asynchronously throughout the spindle and asters, but the FRAP data are most compatible with a rapid exchange of tubulin subunits all along the entire lengths of nonkinetochore spindle and astral microtubules.  相似文献   

16.
The pathway for the in vitro equilibrium unfolding of the tubulin heterodimer by guanidinium chloride (GdmCl) has been studied using several spectroscopic techniques, specifically circular dichroism (CD), two-photon Fluorescence Correlation Spectroscopy (FCS), and time-resolved fluorescence, including lifetime and dynamic polarization. The results show that tubulin unfolding is characterized by distinct processes that occur in different GdmCl concentration ranges. From 0 to 0.5 M GdmCl, a slight alteration of the tubulin heterodimer occurs, as evidenced by a small, but reproducible increase in the rotational correlation time of the protein and a sharp decrease in the secondary structure monitored by CD. In the range 0.5-1.5 M GdmCl, significant decreases in the steady-state anisotropy and average lifetime of the intrinsic tryptophan fluorescence occur, as well as a decrease in the rotational correlation time, from 48 to 26 nsec. In the same GdmCl range, the number of protein molecules (labeled with Alexa 488), as determined by two-photon FCS measurements, increases by a factor of two, indicating dissociation of the tubulin dimer into monomers. From 1.5 to 4 M GdmCl, these monomers unfold, as evidenced by the continual decrease in the tryptophan steady-state anisotropy, average lifetime, and rotational correlation time, concomitant with secondary structural changes. These results help to elucidate the unfolding pathway of the tubulin heterodimer and demonstrate the value of FCS measurements in studies on oligomeric protein systems.  相似文献   

17.
The dissociation constants for GTP and GDP with tubulin were determined to be equal to 1.1 ± 0.4 × 10?7 M and 1.5 ± .6 × 10?7 (4°), respectively. A lower limit for the dissociation constant for ATP was established as equal to 6 × 10?4 M. The equivalent binding of GTP and GDP is not readily consistent with a mechanism in which the role of GTP in microtubule assembly is to bind to the protein to induce a conformation which is able to polymerize. An ATP-induced polymerization of tubulin apparently involves a transphosphorylation reaction in which GTP is formed and mediates the assembly. For this reaction to occur with desalted tubulin trace amounts of GDP are required; in the reaction of 0.1 mM ATP with 22.0 μM tubulin, 0.1 μM GDP induces about 80% as much tubule formation as is seen with 0.1 mM GTP alone.  相似文献   

18.
A bioriented chromosome is tethered to opposite spindle poles during congression by bundles of kinetochore microtubules (kMts). At room temperature, kinetochore fibers are a dominant component of mitotic spindles of PtK2 cells. PtK2 cells at room temperature were injected with purified tubulin covalently bound to DTAF and congression movements of individual chromosomes were recorded in time lapse. Congression movements of bioriented chromosomes between the poles occur over distances of 4.5 microns or greater. DTAF-tubulin injection had no effect on either the velocity or extent of these movements. Other cells were lysed, fixed, and the location of DTAF-tubulin incorporation was detected from digitally processed images of indirect immunofluorescence of an antibody to DTAF. Microtubules were labeled with an anti-beta tubulin antibody. At 2-5 minutes after injection, concentrated DTAF-tubulin staining was seen in the kinetochore fibers proximal to the kinetochores; a low concentration of DTAF-tubulin staining occurred at various sites through the remaining length of the fibers toward the pole. Kinetochore fibers in the same cell displayed different lengths (0.2 to 4 microns) of concentrated DTAF-tubulin incorporation proximal to the kinetochore, as did sister kinetochore fibers. Ten minutes after injection, the lengths of DTAF-containing chromosomal fibers were greater than expected if incorporation resulted solely from the lengthening of kinetochore microtubules due to congression movements of the chromosomes. Besides incorporation as a result of chromosome movement, two other mechanisms might explain the length of the DTAF-containing segments: 1) a poleward flux of tubulin subunits (Mitchison, 1989) or 2) capture of DTAF-containing nonkinetochore microtubules.  相似文献   

19.
Intestinal fat digestion is carried out by the concerted action of pancreatic lipase and its protein cofactor colipase. Colipase is secreted from pancreas as a procolipase and is transformed into colipase by the trypsin cleavage of the Arg5-Gly6 bond during liberation of an N-terminal pentapeptide. The kinetic parameters for the lipase-colipase system compared to the lipase-procolipase system has been compared using trioctanoin and Intralipid as substrates. It was found that at pH 7.0 the Kmapp using Intralipid as substrate was the same for procolipase and colipase, 0.06 mM and 0.05 mM, respectively. At pH 8.0, however, the Kmapp were different-0.23 mM for procolipase and 0.08 mM for colipase. In a similar way the binding between colipase and lipase had a dissociation constant of 2.4 x 10(-6) M at pH 7.0, while for procolipase--lipase binding the dissociation constant was 4.1 x 10(-6) M with no significant difference. At pH 8.0 the binding between colipase and lipase was stronger, Kd being 2.0 x 10(-7) M, while weaker for procolipase and lipase, Kd being 1.0 x 10(-5) M. It is concluded that at the physiological pH value as is found in the intestine, the activation of procolipase to colipase has no influence on the hydrolysis of trioctanoin or Intralipid in the presence of bile salt.  相似文献   

20.
Cytoplasmic sn-glycerol-3-phosphate dehydrogenase, labelled covalently with fluorescein isothiocyanate, shows an enzyme-concentration-dependent fluorescence anisotropy. The anisotropy versus enzyme concentration curve is shifted towards higher concentrations when substrates are present. The comparison of the dissociation constants estimated from anisotropy measurements and derived from kinetic experiments suggests that the substrate-induced dissociation of the dimeric dehydrogenase is slow with respect to the enzymatic reaction catalyzed by either its monomeric or dimeric form. The fluorescence anisotropy of the fluorescent dye-labelled dehydrogenase increase with time upon addition of unlabelled fructose-1,6-bisphosphate aldolase approaching a limiting value. This fact indicates the binding of fructose-1,6-bisphosphate aldolase aldose aldolase to glycerolphosphate dehydrogenase. A model is proposed assuming simultaneous binding of tetrameric fructose-1,6-bisphosphate aldolase to monomeric and dimeric glycerolphosphate dehydrogenase with 1:1 stoichiometry. The dissociation constants, as parameters fitted to the experimental curves, were estimated as 0.2 microM and 1 microM for aldolase-dimeric-glycerolphosphate-dehydrogenase and aldolase-monomeric-glycerolphosphate-dehydrogenase complexes respectively.  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