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.
Barbar E  Kleinman B  Imhoff D  Li M  Hays TS  Hare M 《Biochemistry》2001,40(6):1596-1605
Cytoplasmic dynein is a multisubunit ATPase that transforms chemical energy into motion along microtubules. LC8, a 10 kDa light chain subunit of the dynein complex, is highly conserved with 94% sequence identity between Drosophila and human. The precise function of this protein is unknown, but its ubiquitous expression and conservation suggest a critical role in the function of the dynein motor complex. We have overexpressed LC8 from Drosophila melanogaster and characterized its dimerization and folding using analytical ultracentrifugation, size-exclusion chromatography, circular dichroism, and fluorescence spectroscopy. Sedimentation equilibrium measurements of LC8 at pH 7 reveal a reversible monomer-dimer equilibrium with a dissociation constant of 12 microM at 4 degrees C. At lower pH, LC8 dissociates to a monomer, with a transition midpoint at pH 4.8. Far-UV CD and fluorescence spectra demonstrate that pH-dissociated LC8 retains native secondary and tertiary structures, while the diminished near-UV CD signal shows loss of quaternary structure. The observation that dimeric LC8 dissociates at low pH can be explained by titration of a histidine pair in the dimer interface. Equilibrium denaturation experiments with a protein concentration range spanning almost 2 orders of magnitude indicate that unfolding of LC8 dimer is a two-stage process, in which global unfolding is preceded by dissociation to a folded monomer. The nativelike tertiary structure of the monomer suggests a role for the monomer-dimer equilibrium of LC8 in dynein function.  相似文献   

2.
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   

3.
4.
5.
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   

6.
Porcine heart citrate synthase, a dimeric protein of Mr = 100,000 composed of two identical subunits, is shown to undergo a monomer-dimer equilibrium. The extent of dimerization is found to be dependent on the concentration of citrate synthase, pH, ionic strength, and the specific buffer system employed. Oxaloacetate and citrate, substrates for the forward and reverse reaction catalyzed by citrate synthase, affect dimerization at concentrations of the protein which exists as monomer in their absence. The dissociation of citrate synthase dimers has been demonstrated utilizing the techniques of gel permeation chromatography, fluorescence polarization, fluorescence energy transfer, and heat denaturation. Earlier studies of citrate synthase quarternary structure found the protein to be nondissociable except under denaturing conditions or extensive modification; however, most former studies were performed at relatively high protein concentration, ionic strength, and pH, conditions which stabilize the dimer. In light of recent evidence derived from x-ray crystallographic studies showing amino acid residues from one subunit contributing to the citrate and CoA binding sites of the other, the dissociation into monomers would be expected to have profound effects on citrate synthase activity and regulation, as well as overall tricarboxylic acid cycle activity.  相似文献   

7.
The thermal stability of the methionine repressor protein from Escherichia coli (MetJ) has been examined over a wide range of pH (pH 3.5-10) and ionic strength conditions using differential scanning calorimetry. Under reducing conditions, the transitions are fully reversible, and thermograms are characteristic of the cooperative unfolding of a globular protein with a molecular weight corresponding to the MetJ dimer, indicating that no dissociation of this dimeric protein occurs before unfolding of the polypeptide chains under most conditions. In the absence of reducing agent, repeated scans in the calorimeter show only partial reversibility, though the thermodynamic parameters derived from the first scans are comparable to those obtained under fully reversible conditions. The protein is maximally stable (Tm 58.5 degrees C) at about pH 6, close to the estimated isoelectric point, and stability is enhanced by increasing ionic strength in the range I = 0.01-0.4 M. The average calorimetric transition enthalpy (delta Hm) for the dimer is 505 +/- 28 kJ mol-1 under physiological conditions (pH 7, I = 0.125, Tm = 53.2 degrees C) and shows a small temperature dependence which is consistent with an apparent denaturational heat capacity change (delta Cp) of about +8.9 kJ K-1 mol-1. The effects of both pH and ionic strength on the transition temperature and free energy of MetJ unfolding are inconsistent with any single amino acid contribution and are more likely the result of more general electrostatic interactions, possibly including significant contributions from electrostatic repulsion between the like-charged monomers which can be modeled by a Debye-Hückel screened potential.  相似文献   

8.
J S Lolkema  G T Robillard 《Biochemistry》1990,29(43):10120-10125
The original proposal of Saier stating that P-enolpyruvate-dependent mannitol phosphorylation is catalyzed by the monomeric form of the bacterial phosphotransferase enzyme IImtl, which would be the form predominantly existing in the phospholipid bilayer, whereas mannitol/mannitol-P exchange would depend on the transient formation of functional dimers, is refuted [Saier, M.H. (1980) J. Supramol. Struct. 14, 281-294]. The correct interpretation of the proportional relation between the rate of mannitol phosphorylation in the overall reaction and the enzyme concentration is that enzyme IImtl is dimeric under the conditions employed. Differences measured in the enzyme concentration dependency of the overall and exchange reactions were caused by different assay conditions. The dimer is favored over the monomer at high ionic strength and basic pH. Mg2+ ions bind specifically to enzyme IImtl, inducing dimerization. A complex formed by mixing inorganic phosphate, F-, and Mg2+ at sufficiently high concentrations inhibits enzyme IImtl, in part, by dissociation of the dimer. Enzyme IImtl was dimeric in 25 mM Tris, pH 7.6, and 5 mM Mg2+ over a large enzyme concentration range and under many different turnover conditions. The association/dissociation equilibrium was demonstrated in phosphate bufers, pH 6.3. The dimer was the most active form both in the overall and in the exchange reaction under the conditions assayed. The monomer was virtually inactive in mannitol/mannitol-P exchange but retained 25% of the activity in the overall reaction.  相似文献   

9.
A kinetic analysis of dimer dissociation, TATA DNA binding, and thermal inactivation of the yeast Saccharomyces cerevisiae and human TATA binding proteins (TBP) was conducted. We find that yeast TBP dimers, like human TBP dimers, are slow to dissociate in vitro (t(1/2) approximately 20 min). Mild mutations in the crystallographic dimer interface accelerate the rate of dimer dissociation, whereas severe mutations prevent dimerization. In the presence of excess TATA DNA, which measures the entire active TBP population, dimer dissociation represents the rate-limiting step in DNA binding. These findings provide a biochemical extension to genetic studies demonstrating that TBP dimerization prevents unregulated gene expression in yeast [Jackson-Fisher, A. J., Chitikila, C., Mitra, M., and Pugh, B. F. (1999) Mol. Cell 3, 717-727]. In the presence of vast excesses of TBP over TATA DNA, which measures only a very small fraction of the total TBP, the monomer population in a monomer/dimer equilibrium binds DNA rapidly, which is consistent with a simultaneous binding and bending of the DNA. Under conditions where other studies failed to detect dimers, yeast TBP's DNA binding activity was extremely labile in the absence of TATA DNA, even at temperatures as low as 0 degrees C. Kinetic analyses of TBP instability in the absence of DNA at 30 degrees C revealed that even under fairly stabilizing solution conditions, TBP's DNA binding activity rapidly dissipated with t(1/2) values ranging from 6 to 26 min. TBP's stability appeared to vary with the square root of the TBP concentration, suggesting that TBP dimerization helps prevent TBP inactivation.  相似文献   

10.
Dialysis kinetics measurements have been made to study the effect of ionic strength on the dimerization of lysozyme in acidic solutions that lead to the growth of tetragonal lysozyme crystals. Using glutaraldehyde cross-linked dimers of lysozyme, we have determined that both monomers and dimers can escape from 25,000 molecular weight cutoff dialysis membranes with velocity constants of 5.1 x 10(-7) and 1.0 x 10(-7) s(-1) for the monomer and dimer species, respectively. The flux from 25K MWCO membranes has been measured for lysozyme in pH 4.0 buffered solutions of 1, 3, 4, 5, and 7% NaCl over a wide range of protein concentrations. Assuming that dimerization is the first step in crystallization, a simple monomer to dimer equilibrium was used to model the flux rates. Dimerization constants calculated at low protein concentrations were 265, 750, 1212, and 7879 M(-1) for 3, 4, 5, and 7% NaCl, respectively. These values indicate that dimerization increases with the ionic strength of the solution suggesting that aggregation is moderated by electrostatic interactions. At high protein concentrations and high supersaturation, the dimerization model does not describe the data well. However, the Li model that uses a pathway of monomer <-> dimer <-> tetramer <-> octamer <-> 16-mer fits the measured flux data remarkably well suggesting the presence of higher order aggregates in crystallizing solutions.  相似文献   

11.
Spectrophotometric data have been determined for mesoferrihaem at several pH values and over a range of concentration covering four orders of magnitude. The data reveal a dimerization process according to the equation 2 monomer in equilibrium dimer + H+, analogous to earlier findings for deuteroferrihaem and protoferrihaem. The value of K (defined as K = [dimer] [H+]/[monomer]2) was found to be 6.92.10(-2). This is close to the value for deuteroferrihaem but much less than that for protoferrihaem. This is interpreted in terms of possible additional bonding between the delocalized electron systems in protoferrihaem dimers relative to those of mesoferrihaem and deuteroferrihaem. Rate constants for dimerization were determined by temperature-jump spectrophotometry. The pH dependence of the rate constants is explained in terms of two distinct pathways for the dimerization process. These involve either direct reaction between two undissociated monomer molecules or alternatively an initial acid dissociation of a monomer molecule followed by reaction between an undissociated and dissociated molecule.  相似文献   

12.
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   

13.
Cooperativity in the interactions among proteins subunits and DNA is crucial for DNA recognition. LexA repressor was originally thought to bind DNA as a monomer, with cooperativity leading to tighter binding of the second monomer. The main support for this model was a high value of the dissociation constant for the LexA dimer (micromolar range). Here we show that the protein is a dimer at nanomolar concentrations under different conditions. The reversible dissociation of LexA dimer was investigated by the effects of hydrostatic pressure or urea, using fluorescence emission and polarization to monitor the dissociation process. The dissociation constant lies in the picomolar range (lower than 20 pM). LexA monomers associate with an unusual large volume change (340 ml/mol), indicating the burial of a large surface area upon dimerization. Whereas nonspecific DNA has no stabilizing effect, specific DNA induces tightening of the dimer and a 750-fold decrease in the K(d). In contrast to the previous model, a tight dimer rather than a monomer is the functional repressor. Accordingly, the LexA dimer only loses its ability to recognize a specific DNA sequence by RecA-induced autoproteolysis. Our work provides insights into the linkage between protein-protein interactions, DNA recognition, and DNA repair.  相似文献   

14.
The polymerization of aryl sulfatase A (aryl sulfate sulfohydrolase, EC 3.1.6.1) has been studied by frontal gel chromatography on Sephadex G-200 and Bio-Gel A-5m under various conditions of pH, ionic strength, and temperature. The aryl sulfatase A molecule exists as a monomer and as a dimer at pH 7.5 and pH 4.5, respectively. The extent of dissociation is markedly pH-, protein concentration-, and ionic strength-dependent. Only a small effect of temperature was observed. The enthalpy change (ΔHo) for the dissociation was ?2.5 ± 1 kcal/mol at pH 5.5–5.6, and the entropy change for dissociation of the enzyme dimer to two monomeric units was ?47 cal mol?1 deg?1. Sulfate ion has little effect on the extent of dissociation of the enzyme at pH 5.6. The present studies suggest that the dissociation of rabbit liver aryl sulfatase A is regulated by the ionization of amino acid residues whose apparent pK is between pH 5 and 6. The driving force for the association of the subunits of the enzyme is primarily ionic and/or ionic/hydrogen bond formation. The small enthalpy change and the fact that dissociation is strongly favored by an increase in the ionic strength suggest that hydrophobic interactions play only a minor role in stabilizing the dimeric quaternary structure relative to the monomeric state. The monomeric form of the enzyme exhibits the anomalous kinetics often observed with sulfatase A but the dimer does not show anomalous kinetics. Since aryl sulfatase A is probably in the dimeric form in the lysosome, the anomalous kinetics of the enzyme are unlikely to be of physiological importance in the intact lysosome.  相似文献   

15.
D A Yphantis  T Arakawa 《Biochemistry》1987,26(17):5422-5427
Recombinant DNA derived human interferon gamma (IFN-gamma) from Escherichia coli was examined by equilibrium ultracentrifugation. Short-column equilibrium experiments at pH 6.9 in 0.1 M ammonium acetate buffer gave a z-average molecular weight of 33,500 +/- 1400 at infinite dilution, corresponding to 1.98 +/- 0.08 times the formula weight. Long- (2.6 mm) column experiments at pH 7.5 in 0.04 M imidazole buffer gave a molecular weight of 33,400 +/- 500. Under the latter conditions IFN-gamma behaves somewhat nonideally, with the departure from ideality accounted for by an effective (Donnan) charge of about 6+. No association of this dimer to form tetramer or higher polymers was observed, with the association constant for formation of tetramer from dimer K24 found to be less than 34 L mol-1. Similarly, no dissociation to monomers was observable, with the dissociation constant to monomer K21 being less than 5 X 10(-8) mol L-1. At pH 3.55 in 0.02 M buffer (acetate plus acetic acid), there was virtually complete dissociation of the dimer to monomer. Extreme nonideality was seen in this low ionic strength system, and the effective charge on the protein was estimated to be about 11+. The reduced molecular weight M(1 -upsilon rho) of the monomer was found to be about 4.09 +/- 0.20 kg mol-1; this corresponds to a molecular weight of 16,410 +/- 820, with the Scatchard definition of components. A small amount of a polymer with a molecular weight of about 0.5 X 10(6) was detected under these conditions.  相似文献   

16.
Kim DH  Jang DS  Nam GH  Yun S  Cho JH  Choi G  Lee HC  Choi KY 《Biochemistry》2000,39(42):13084-13092
Equilibrium and kinetic analyses have been carried out to elucidate the folding mechanism of homodimeric ketosteroid isomerase (KSI) from Comamonas testosteroni. The folding of KSI was reversible since the activity as well as the fluorescence and CD spectra was almost completely recovered after refolding. The equilibrium unfolding transitions monitored by fluorescence and CD measurements were almost coincident with each other, and the transition midpoint increased with increasing protein concentration. This suggests that the KSI folding follows a simple two-state mechanism consisting of native dimer and unfolded monomer without any thermodynamically stable intermediates. Sedimentation equilibrium analysis and size-exclusion chromatography of KSI at different urea concentrations supported the two-state model without any evidence of folded monomeric intermediates. Consistent with the two-state model, (1)H-(15)N HSQC spectra obtained for KSI in the unfolding transition region could be reproduced by a simple addition of the spectra of the native and the unfolded KSI. The KSI refolding kinetics as monitored by fluorescence intensity could be described as a fast first-order process followed by a second-order and a subsequent slow first-order processes with rate constants of 60 s(-)(1), 5.4 x 10(4) M(-)(1).s(-)(1), and 0.017 s(-)(1), respectively, at 0.62 M urea, suggesting that there may be a monomeric folding intermediate. After a burst phase that accounts for >83% of the total amplitude, the negative molar ellipticity at 225 nm increased slowly in a single phase at a rate comparable to that of the bimolecular intermediate step. The kinetics of activity recovery from the denatured state were markedly dependent upon the protein concentration, implying that the monomers are not fully active. Taken together, our results demonstrate that the dimerization induces KSI to fold into the complete structure and is crucial for maintaining the tertiary structure to perform efficient catalysis.  相似文献   

17.
The formation of monomer from several hemoglobins has been investigated by sedimentation equilibrium. The use of the split-beam photoelectric scanning absorption optical system has enabled observations to be made routinely down to 1 μg/ml. (6.2 × 10−8m-heme) with strict spectral control of the integrity of the hemoglobin molecule. The results show that the dissociation constant of dimer to monomer at neutral pH and moderate ionic strength is so small that monomer is present in reversible equilibrium with dimer only in fractions too small to be detectable. Any appreciable monomer formation is irreversible and accompanied by usually pronounced spectral changes. This irreversible monomer formation is probably a consequence of the presence of heavy-metal ions in solution and may be inhibited by 10−3m-EDTA. Hemoglobin ligands possessing chelating ability also inhibit monomer formation.  相似文献   

18.
The mechanism of herpesviral protease activation upon dimerization was studied using two independent spectroscopic assays augmented by directed mutagenesis. Spectroscopic changes, attributable to dimer interface conformational plasticity, were observed upon dimerization of Kaposi's sarcoma-associated herpesvirus protease (KSHV Pr). KSHV Pr's dissociation constant of 585 +/- 135 nM at 37 degrees C was measured by a concentration-dependent, 100-fold increase in specific activity to a value of 0.275 +/- 0.023 microM product min(-1) (microM enzyme)(-1). A 4 nm blue-shifted fluorescence emission spectrum and a 25% increase in ellipticity at 222 nm were detected by circular dichroism upon dimer association. This suggested enhanced hydrophobic packing within the dimer interface and/or core, as well as altered secondary structures. To better understand the structure-activity relationship between the monomer and the dimer, KSHV Pr molecules were engineered to remain monomeric via substitution of two separate residues within the dimer interface, L196 and M197. These mutants were proteolytically inactive while exhibiting the spectroscopic signature and thermal stability of wild type, dissociated monomers (T(M) = 75 degrees C). KSHV Pr conformational changes were found to be relevant in vivo, as the autoproteolytic inactivation of KSHV Pr at its dimer disruption site [Pray et al. (1999) J. Mol. Biol. 289, 197-203] was detected in viral particles from KSHV-infected cells. This characterization of structural plasticity suggests that the structure of the KSHV Pr monomer is stable and significantly different from its structure in the dimer. This structural uniqueness should be considered in the development of compounds targeting the dimer interface of KSHV Pr monomers.  相似文献   

19.
Apoptosis is an essential mechanism of cell death required for normal development and homeostasis of all multicellular organisms. Smac/DIABLO is a dimeric protein important in the control of apoptosis by removing the inhibitory activity of IAPs (inhibitor of apoptosis proteins). In vitro studies reveal that dimerization is required for its function. Here we investigate the structural and thermodynamic features of folding and dimerization of Smac/DIABLO. To disturb the folded, dimeric structure, we used high hydrostatic pressure, low and high temperatures, and chemical denaturing agents. Conformational changes were monitored using spectroscopic techniques such as fluorescence and circular dichroism (CD) as well as gel filtration chromatography. Our data show that Smac/DIABLO is very stable under pressures up to 3.1 kbar, even at subzero temperatures. A complete denaturation/dissociation process is obtained when we use high concentrations of urea, which affect its secondary structure as assessed by CD. The association of pressure and subdenaturing urea concentrations also results in complete denaturation/dissociation of the protein. Under these conditions, unfolding of the protein shows concentration dependence that is in accordance with the dimer-monomer dissociation equilibrium, confirming Smac/DIABLO dissociation. These results suggest that most of the treatments lead to a reversible disruption of the dimeric structure with a dissociation constant ( K d) of 34 x 10 (-21) M (34 zM). This tight dimer is biologically relevant, considering that monomeric mutants bind IAP with low affinity. The extremely high stability of the dimeric form of Smac/DIABLO also implies that once expressed in the cell the protein has a low probability of dissociation and, consequently, loss of function. In addition, the stability in the zeptomolar range is the highest so far measured for a dimeric protein. It also indicates that under most circumstances Smac/DIABLO does not exist as a monomer in the cell and suggests that the dimer-to-monomer equilibrium does not play a regulatory role in the Smac/DIABLO-IAP interaction.  相似文献   

20.
1. From the standpoint of monomer-dimer equilibrium of hog kidney D-amino acid oxidase [EC 1.4.3.3] and the interaction between the enzyme and small molecules, the effect of pH on the binding of p-aminobenzoate to the monomer and dimer of the enzyme was studied by kinetic methods and spectrophotometric titration. 2. The maximum binding number of p-aminobenzoate to the dimer is two molecules, and there is no interaction between the two active sites of the dimer (i.e., no cooperativity) over the range of pH from 6.5 to 10. 3. The affinity of the dimer for p-aminobenzoate is several times higher than that of the monomer at pH 6.5-10, and consequently p-aminobenzoate induces dimerization in the equilibrium state of D-amino acid oxidase. The interaction energy of two subunits of the dimer is stabilized by the binding of p-aminobenzoate by 1-2 kcal/mole over the pH range studied. 4. The binding sites of the quasi-substrate, p-aminobenzoate, in the dimer and the intersubunit binding site of the dimer are clearly different, because p-aminobenzoate induces dimerization of the enzyme. 5. The pK values of ionizing groups in the free monomer and the free dimer which participate in the binding of the competitive inhibitor, p-aminobenzoate, are approximately the same, 8.7, as determined from the pH dependence of the affinity of the inhibitor for the enzyme. Furthermore, no pK for the enzyme-inhibitor complex in the pH range 6.5-10 was observed. 6. There is no interaction between the two ionizing groups of the dimer during protonation-deprotonation, because a theoretical equation involving no cooperativity between the two ionizing groups in the dimer explains the results well.  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