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 62 毫秒
1.
Electron transfer within complexes of cytochrome c (Cc) and cytochrome c peroxidase (CcP) was studied to determine whether the reactions are gated by fluctuations in configuration. Electron transfer in the physiological complex of yeast Cc (yCc) and CcP was studied using the Ru-39-Cc derivative, in which the H39C/C102T variant of yeast iso-1-cytochrome c is labeled at the single cysteine residue on the back surface with trisbipyridylruthenium(II). Laser excitation of the 1:1 Ru-39-Cc-CcP compound I complex at low ionic strength results in rapid electron transfer from RuII to heme c FeIII, followed by electron transfer from heme c FeII to the Trp-191 indolyl radical cation with a rate constant keta of 2 x 10(6) s-1 at 20 degrees C. keta is not changed by increasing the viscosity up to 40 cP with glycerol and is independent of temperature. These results suggest that this reaction is not gated by fluctuations in the configuration of the complex, but may represent the elementary electron transfer step. The value of keta is consistent with the efficient pathway for electron transfer in the crystalline yCc-CcP complex, which has a distance of 16 A between the edge of heme c and the Trp-191 indole [Pelletier, H., and Kraut, J. (1992) Science 258, 1748-1755]. Electron transfer in the complex of horse Cc (hCc) and CcP was examined using Ru-27-Cc, in which hCc is labeled with trisbipyridylruthenium(II) at Lys-27. Laser excitation of the Ru-27-Cc-CcP complex results in electron transfer from RuII to heme c FeII with a rate constant k1 of 2.3 x 10(7) s-1, followed by oxidation of the Trp-191 indole to a radical cation by RuIII with a rate constant k3 of 7 x 10(6) s-1. The cycle is completed by electron transfer from heme c FeII to the Trp-191 radical cation with a rate constant k4 of 6.1 x 10(4) s-1. The rate constant k4 decreases to 3.4 x 10(3) s-1 as the viscosity is increased to 84 cP, but the rate constants k1 and k3 remain the same. The results are consistent with a gating mechanism in which the Ru-27-Cc-CcP complex undergoes fluctuations between a major state A with the configuration of the hCc-CcP crystalline complex and a minor state B with the configuration of the yCc-CcP complex. The hCc-CcP complex, state A, has an inefficient pathway for electron transfer from heme c to the Trp-191 indolyl radical cation with a distance of 20.5 A and a predicted value of 5 x 10(2) s-1 for k4A. The observed rate constant k4 is thus gated by the rate constant ka for conversion of state A to state B, where the rate of electron transfer k4B is expected to be 2 x 10(6) s-1. The temperature dependence of k4 provides activation parameters that are consistent with the proposed gating mechanism. These studies provide evidence that configurational gating does not control electron transfer in the physiological yCc-CcP complex, but is required in the nonphysiological hCc-CcP complex.  相似文献   

2.
During iron acquisition by the cell, complete homodimeric transferrin receptor 1 in an unknown state (R1) binds iron-loaded human serum apotransferrin in an unknown state (T) and allows its internalization in the cytoplasm. T also forms complexes with metals other than iron. Are these metals incorporated by the iron acquisition pathway and how can other proteins interact with R1? We report here a four-step mechanism for cobalt(III) transfer from CoNtaCO(3)(2-) to T and analyze the interaction of cobalt-loaded transferrin with R1. The first step in cobalt uptake by T is a fast transfer of Co(3+) and CO(3)(2-) from CoNtaCO(3)(2-) to the metal-binding site in the C-lobe of T: direct rate constant, k(1)=(1.1+/-0.1) x 10(6) M(-1) s(-1); reverse rate constant, k(-1)=(1.9+/-0.6) x 10(6) M(-1) s(-1); and equilibrium constant, K=1.7+/-0.7. This step is followed by a proton-assisted conformational change of the C-lobe: direct rate constant, k(2)=(3+/-0.3) x 10(6) M(-1) s(-1); reverse rate constant, k(-2)=(1.6+/-0.3) x 10(-2) s(-1); and equilibrium constant, K(2a)=5.3+/-1.5 nM. The two final steps are slow changes in the conformation of the protein (0.5 h and 72 h), which allow it to achieve its final thermodynamic state and also to acquire second cobalt. The cobalt-saturated transferrin in an unknown state (TCo(2)) interacts with R1 in two different steps. The first is an ultra-fast interaction of the C-lobe of TCo(2) with the helical domain of R1: direct rate constant, k(3)=(4.4+/-0.6)x10(10) M(-1) s(-1); reverse rate constant, k(-3)=(3.6+/-0.6) x 10(4) s(-1); and dissociation constant, K(1d)=0.82+/-0.25 muM. The second is a very slow interaction of the N-lobe of TCo(2) with the protease-like domain of R1. This increases the stability of the protein-protein adduct by 30-fold with an average overall dissociation constant K(d)=25+/-10 nM. The main trigger in the R1-mediated iron acquisition is the ultra-fast interaction of the metal-loaded C-lobe of T with R1. This step is much faster than endocytosis, which in turn is much faster than the interaction of the N-lobe of T with the protease-like domain. This can explain why other metal-loaded transferrins or a protein such as HFE-with a lower affinity for R1 than iron-saturated transferrin but with, however, similar or higher affinities for the helical domain than the C-lobe-competes with iron-saturated transferrin in an unknown state towards interaction with R1.  相似文献   

3.
Kinetic interactions of beta-lactam antibiotics such as penicillin-G and cefotaxime with normal, penicillin-susceptible PBP2x from Streptococcus pneumoniae and a penicillin-resistant PBP2x (PBP2x(R)) from a resistant clinical isolate (CS109) of the bacterium have been extensively characterized using electrospray mass spectrometry coupled with a fast reaction (quench flow) technique. Kinetic evidence for a two-step acylation of PBP2x by penicillin-G has been demonstrated, and the dissociation constant, K(d) of 0.9 mm, and the acylation rate constant, k(2) of 180 s(-1), have been determined for the first time. The millimolar range K(d) implies that the beta-lactam fits to the active site pocket of the penicillin-sensitive PBP rather poorly, whereas the extremely fast k(2) value indicates that this step contributes most of the binding affinity of the beta-lactam. The values of K(d) (4 mm) and k(2) (0.56 s(-1)) were also determined for PBP2x(R). The combined value of k(2)/K(d), known as overall binding efficiency, for PBP2x(R) (137 m(-1) s(-1)) was over 1000-fold slower than that for PBP2x (200,000 m(-1) s(-1)), indicating that a major part is played by the acylation steps in penicillin resistance. Most of the decreased binding efficiency of PBP2x(R) comes from the decreased ( approximately 300-fold) k(2). Kinetic studies of cefotaxime acylation of the two PBP2x proteins confirmed all of the above findings. Deacylation rate constants (k(3)) for the third step of the interactions were determined to be 8 x 10(-6) s(-1) for penicilloyl-PBP2x and 5.7 x 10(-4) s(-1) for penicilloyl-PBP2x(R), corresponding to over 70-fold increase of the deacylation rate for the resistant PBP2x(R). Similarly, over 80-fold enhancement of the deacylation rate was found for cefotaxime-PBP2x(R) complex (k(3) = 3 x 10(-4) s(-1)) as compared with that of cefotaxime-PBP2x complex (3.5 x 10(-6) s(-1)). This is the first time that such a significant increase of k(3) values was found for a beta-lactam-resistant penicillin-binding protein. These data indicate that the deacylation step also plays a role, which is much more important than previously thought, in PBP2x(R) resistance to beta-lactams.  相似文献   

4.
The kinetics and thermodynamics of Bi(III) exchange between bismuth mononitrilotriacetate (BiL) and human serum transferrin as well as those of the interaction between bismuth-loaded transferrin and transferrin receptor 1 (TFR) were investigated at pH 7.4-8.9. Bismuth is rapidly exchanged between BiL and the C-site of human serum apotransferrin in interaction with bicarbonate to yield an intermediate complex with an effective equilibrium constant K(1) of 6 +/- 4, a direct second-order rate constant k(1) of (2.45 +/- 0.20) x 10(5) M(-1) s(-1), and a reverse second-order rate constant k(-1) of (1.5 +/- 0.5) x 10(6) M(-1) s(-1). The intermediate complex loses a single proton with a proton dissociation constant K(1a) of 2.4 +/- 1 nM to yield a first kinetic product. This product then undergoes a modification in its conformation followed by two proton losses with a first-order rate constant k(2) = 25 +/- 1.5 s(-1) to produce a second kinetic intermediate, which in turn undergoes a last modification in the conformation to yield the bismuth-saturated transferrin in its final state. This last process rate-controls Bi(III) uptake by the N-site of the protein and is independent of the experimental parameters with a constant reciprocal relaxation time tau(3)(-1) of (3 +/- 1) x 10(-2) s(-1). The mechanism of bismuth uptake differs from that of iron and probably does not involve the same transition in conformation from open to closed upon iron uptake. The interaction of bismuth-loaded transferrin with TFR occurs in a single very fast kinetic step with a dissociation constant K(d) of 4 +/- 0.4 microM, a second-order rate constant k(d) of (2.2 +/- 1.5) x 10(8) M(-1) s(-1), and a first-order rate constant k(-d) of 900 +/- 400 s(-1). This mechanism is different from that observed with the ferric holotransferrin and implies that the interaction between TFR and bismuth-loaded transferrin probably takes place on the helical domain of the receptor which is specific for the C-site of transferrin and HFE. The relevance of bismuth incorporation by the transferrin receptor-mediated iron acquisition pathway is discussed.  相似文献   

5.
Iron release from ovotransferrin in acidic media (3 < pH < 6) occurs in at least six kinetic steps. The first is a very fast (相似文献   

6.
The reaction between antiplasmin (A) and a low-molecular-weight form of plasmin (P) proceeds in at least two steps: a fast reversible second-order reaction followed by a slower irreversible first-order transition, and may be represented by: P +A k1 in equilibrium k-1 PA k2 leads to PA'. The low-Mr plasmin, which is obtained by limited elastase digestion, is composed of an intact B chain and a small A chain lacking the lysine-binding sites. The k1 of the reaction is (6.5 +/- 0.5) x 10(5) M-1 s-1 which is 30--60 times smaller than that for normal plasmin and antiplasmin. The dissociation constant of the first step is 1.9 x 10(-9) M which is 10 times higher than for normal plasmin and antiplasmin. The rate constant of the second step is (4.2 +/- 0.2) x 10(-3) s-1 for both normal and low-Mr plasmin. Low Mr plasmin which has substrate bound to its active site does not react or reacts only very slowly with antiplasmin. The reaction rate, however, is only slightly influenced by 6-aminohexanoic acid in concentrations up to 1 mM which decrease the reaction rate of normal plasmin approximately 50-fold. The findings further indicate that the lysine-binding site(s) of plasmin are of great importance for the rate of its reaction with antiplasmin.  相似文献   

7.
Wu SP  Cowan JA 《Biochemistry》2003,42(19):5784-5791
ISA type proteins mediate cluster transfer to apoprotein targets. Rate constants have been determined for cluster transfer from Schizosaccharomyces pombe ISA to apo Fd. Substitution of the cysteine residues of ISA produced derivative proteins (C72A, C136A, and C138A) that were found to be at least as active in cluster transfer reactions as the native form at 25 degrees C (k(2) approximately 170 M(-1) min(-1) for native, k(2) approximately 169 M(-1) min(-1) for C72A, k(2) approximately 206 M(-1) min(-1) for C136A, and k(2) approximately 242 M(-1) min(-1) for C138A), although the yield of cluster transfer was found to be lower as a consequence of the enhanced lability of clusters in the derivative proteins. Minor variations in rate constant for the ISA Cys derivatives do not reflect any change in the affinity of binding to the apo Fd since k(2) was found to be independent of the concentration of apo Fd over the range of 1-25 microM. The pH dependence of cluster transfer rates was found to be similar for native and C136A ISA, with an observed pK(a) of 7.8 determined from the pH profiles for cluster transfer activity of each protein. The temperature dependence of the rate constant defining the cluster transfer reaction for the wild type versus this C136A ISA derivative is distinct (DeltaH* approximately 6.3 kcal mol(-1) and DeltaS* approximately -27.3 cal K(-1) mol(-1) for native and DeltaH* approximately 2.7 kcal mol(-1) and DeltaS* approximately -38.9 cal K(-1) mol(-1) for C136A ISA). Instability of the protein-bound cluster precluded a comparison with data from pH and temperature dependencies for the two other Cys derivatives. Experiments to determine the dependence of reaction rate constants on viscosity indicate cluster transfer is rate-limiting. A comparison of cross-species rate constants for cluster transfer to apo Fd targets from Homo sapiens and S. pombe demonstrated that the identity of the Fd is less critical for promoting cluster transfer from Sp ISA (at 25 degrees C, k(2) approximately 170 M(-1) min(-1) for Sp Fd and k(2) approximately 169 M(-1) min(-1) for Hs Fd). This contrasts with an earlier observation for ISU-mediated cluster assembly [Wu, S., et al. (2002) Biochemistry 41, 8876-8885], where the rates differed for Hs and Sp target Fd's, suggesting distinct binding sites for binding of holo ISA and ISU to apo Fd.  相似文献   

8.
A sensitive competitive method was developed for assessing the reactivity of compounds toward glutathione and toward thiols in general. The method employs the reaction of the fluorogenic reagent fluorescein-5-maleimide (FM) with glutathione (GSH) to generate a large increase in fluorescence emission. When the reaction is measured in the presence of a compound that competes with FM toward GSH, the rate constant for fluorescent product formation increases while the total amount of product formed at the end of the reaction decreases. These changes in the presence of a series of competitor concentrations allow one to calculate the rate constant of the reaction of the competitor with GSH. At 23 degrees C, pH 7.40 in PBS buffer the second-order rate constant of the FM-GSH reaction is k2 = (1.67 +/- 0.32) x 10(4) M(-1) x s(-1). Two GSH-reactive compounds were evaluated: the second-order rate constant for the reaction of PNU-27707 with GSH under our experimental conditions is k(i) = 5660 +/- 266 M(-1) x s(-1), while that of PNU-37802 is k(i) = 21,200 +/- 1600 M(-1) x s(-1). The method is easily adaptable to a high-throughput screening format.  相似文献   

9.
Octadecylrhodamine B chloride (R18) and ganglioside GD1a (virus receptor) were incorporated into small unilamellar liposomes [Hoekstra et al. (1984) Biochemistry 23, 5675-5681]. Upon interaction of these liposomes with PR8 influenza viruses without prebinding, two types of dequenching were observed at 37 degrees C, both second-order processes: a fast reaction at pH 5.3, 2k = 17.53 x 10(-3) (Q.s)-1, and a slow reaction at pH 7.4, 2k = 0.335 x 10(-3) (Q.s)-1. The maximal level of dequenching was the same for both. Upon prebinding of liposomes to PR8 viruses (30 min, 0 degrees C, pH 7.4) at high concentrations, a very fast dequenching occurred when the prebinding mixture was diluted into prewarmed (37 degrees C) 10 mM PBS, pH 5.3. For the initial phase, a first-order rate constant of 0.5 s-1 could be extrapolated. After a quick drop in velocity during the first 30 s, the reaction was kinetically indistinguishable from the one found without prebinding. A second-order process with 2k = 16.52 x 10(-3) (Q.s)-1 became rate-limiting. The fast reactions at pH 5.3 can be abolished by inactivation or removal of the virus hemagglutinin. We conclude that the reaction at pH 5.3 reflects the hemagglutinin-dependent fusion process known to occur between influenza viruses and partner membranes at low pH; however, second-order kinetics indicate that specific binding rather than fusion is the rate-limiting step. For the slow dequenching, which is not affected by prebinding, the rate constant is 20 times lower than for the fast reaction, and the process is independent of viral hemagglutinin.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
The kinetics and thermodynamics of Al(III) exchange between aluminum citrate (AlL) and human serum transferrin were investigated in the 7.2-8.9 pH range. The C-site of human serum apotransferrin in interaction with bicarbonate removes Al(III) from Al citrate with an exchange equilibrium constant K1 = (2.0 +/- 0.6) x 10(-2); a direct second-order rate constant k1 = 45 +/- 3 M(-1) x s(-1); and a reverse second-order rate constant k(-1) = (2.3 +/- 0.5) x 10(3) M(-1) x s(-1). The newly formed aluminum-protein complex loses a single proton with proton dissociation constant K1a = (15 +/- 3) nM to yield a first kinetic intermediate. This intermediate then undergoes a modification in its conformation followed by two proton losses; first-order rate constant k2 = (4.20 +/- 0.02) x 10(-2) s(-1) to produce a second kinetic intermediate, which in turn undergoes a last slow modification in the conformation to yield the aluminum-loaded transferrin in its final state. This last process rate-controls Al(III) uptake by the N-site of the protein and is independent of the experimental parameters with a constant reciprocal relaxation time tau3(-1) = (6 +/- 1) x 10(-5) x s(-1). The affinities involved in aluminum uptake by serum transferrins are about 10 orders of magnitude lower than those involved in the uptake of iron. The interactions of iron-loaded transferrins with transferrin receptor 1 occur with average dissociation constants of 3 +/- 1 and 5 +/- 1 nM for the only C-site iron-loaded and of 6.0 +/- 0.6 and 7 +/- 0.5 nM for the iron-saturated ST in the absence or presence of CHAPS, respectively. No interaction is detected between receptor 1 and aluminum-saturated or mixed C-site iron-loaded/N-site aluminum-loaded transferrin under the same conditions. The fact that aluminum can be solubilized by serum transferrin in biological fluids does not necessarily imply that its transfer from the blood stream to cytoplasm follows the receptor-mediated pathway of iron transport by transferrins.  相似文献   

11.
Manganese superoxide dismutase (Mn-SOD), a critical mitochondrial antioxidant enzyme, becomes inactivated and nitrated in vitro and potentially in vivo by peroxynitrite. Since peroxynitrite readily reacts with transition metal centers, we assessed the role of the manganese ion in the reaction between peroxynitrite and Mn-SOD. Peroxynitrite reacts with human recombinant and Escherichia coli Mn-SOD with a second order rate constant of 1.0 +/- 0.2 x 10(5) and 1.4 +/- 0.2 x 10(5) m(-)1 s(-)1 at pH 7.47 and 37 degrees C, respectively. The E. coli apoenzyme, obtained by removing the manganese ion from the active site, presents a rate constant <10(4) m(-)1 s(-)1 for the reaction with peroxynitrite, whereas that of the manganese-reconstituted apoenzyme (apo/Mn) was comparable to that of the holoenzyme. Peroxynitrite-dependent nitration of 4-hydroxyphenylacetic acid was increased 21% by Mn-SOD. The apo/Mn also promoted nitration, but the apo and the zinc-substituted apoenzyme (apo/Zn) enzymes did not. The extent of tyrosine nitration in the enzyme was also affected by the presence and nature (i.e. manganese or zinc) of the metal center in the active site. For comparative purposes, we also studied the reaction of peroxynitrite with low molecular weight complexes of manganese and zinc with tetrakis-(4-benzoic acid) porphyrin (tbap). Mn(tbap) reacts with peroxynitrite with a rate constant of 6.8 +/- 0.1 x 10(4) m(-)1 s(-)1 and maximally increases nitration yields by 350%. Zn(tbap), on the other hand, affords protection against nitration. Our results indicate that the manganese ion in Mn-SOD plays an important role in the decomposition kinetics of peroxynitrite and in peroxynitrite-dependent nitration of self and remote tyrosine residues.  相似文献   

12.
Iron uptake by transferrin from triacetohydroxamatoFe(III) (Fe(AHA)3) in the presence of bicarbonate has been investigated between pH 7 and 8.2. The protein transits from the opened apo- to the closed holoform by several steps with the accumulation of at least three kinetic intermediates. All these steps are accompanied by proton losses, probably occurring from the protein ligands and the side-chains involved in the interdomain H-bonding nets. The minor bihydroxamatoFe(III) species Fe(AHA)2 exchanges its iron with the C-site of apotransferrin in interaction with bicarbonate without detectable formation of any intermediate protein-iron-ligand mixed complex; direct second-order rate constant k1=4.15(+/-0.05)x10(7) M(-1) s(-1). The kinetic product loses a single proton and undergoes a modification in its conformation followed by the loss of two or three protons; first-order rate constant k2=3.25(+/-0.15) s(-1). This induces a new modification in the conformation; first-order rate constant k3=5.90(+/-0.30)x10(-2) s(-1). This new modification in conformation rate controls iron uptake by the N-site of the protein and is followed by a single proton loss; K3a=6.80 nM. Finally, the holoprotein or the monoferric transferrin in its thermodynamic equilibrated state is produced by a last modification in the conformation occurring in about 4000 seconds. But for the Fe(AHA)3 dissociation and the involvement of Fe(AHA)2 in the first step of iron uptake, this mechanism is identical to that reported for iron uptake from FeNAc3. This implies that the exchange of iron between a chelate and serum-transferrin occurs by a single general mechanism. The nature of the iron-providing chelate is only important for the first kinetic step of the exchange, which can be slowed to such an extent that it rate limits the exchange of iron.  相似文献   

13.
The reaction of myeloperoxidase compound I (MPO-I) with chloride ion is widely assumed to produce the bacterial killing agent after phagocytosis. Two values of the rate constant for this important reaction have been published previously: 4.7 x 106 M-1.s-1 measured at 25 degrees C [Marquez, L.A. and Dunford, H.B. (1995) J. Biol. Chem. 270, 30434-30440], and 2.5 x 104 M-1.s-1 at 15 degrees C [Furtmüller, P.G., Burner, U. & Obinger, C. (1998) Biochemistry 37, 17923-17930]. The present paper is the result of a collaboration of the two groups to resolve the discrepancy in the rate constants. It was found that the rate constant for the reaction of compound I, generated from myeloperoxidase (MPO) and excess hydrogen peroxide with chloride, decreased with increasing chloride concentration. The rate constant published in 1995 was measured over a lower chloride concentration range; the 1998 rate constant at a higher range. Therefore the observed conversion of compound I to native enzyme in the presence of hydrogen peroxide and chloride ion cannot be attributed solely to the single elementary reaction MPO-I + Cl- --> MPO + HOCl. The simplest mechanism for the overall reaction which fit the experimental data is the following: MPO+H2O2 ⇄k-1k1 MPO-I+H2O MPO-I+Cl- ⇄k-2k2 MPO-I-Cl- MPO-I-Cl- -->k3 MPO+HOCl where MPO-I-Cl- is a chlorinating intermediate. We can now say that the 1995 rate constant is k2 and the corresponding reaction is rate-controlling at low [Cl-]. At high [Cl-], the reaction with rate constant k3 is rate controlling. The 1998 rate constant for high [Cl-] is a composite rate constant, approximated by k2k3/k-2. Values of k1 and k-1 are known from the literature. Results of this study yielded k2 = 2.2 x 106 M-1.s-1, k-2 = 1.9 x 105 s-1 and k3 = 5.2 x 104 s-1. Essentially identical results were obtained using human myeloperoxidase and beef spleen myeloperoxidase.  相似文献   

14.
The interaction of DNA and human serum albumin (HSA) in the presence of anticancer drug carboplatin was studied with piezoelectric quartz crystal impedance (PQCI) and electrochemistry techniques. In the PQCI analysis, the correlative parameters including the frequency (f0), the motional resistance (R(m)), and the static capacitance (C0) in the experiment were obtained and discussed in detail. Additionally, the kinetics parameters of the cross-linking process were investigated and a response kinetics model was deduced. The values of association rate constant k1, dissociation rate constant k(-1) and the reaction equilibrium constant K were estimated to be 1.895 mg/ml(-1) s(-1), 4.7 x 10(-5) s(-1) and 4.03 x 10(4) (mg/ml)-1, respectively. Furthermore, cyclic voltammetry (CV) and electrochemical AC impedance techniques were employed to testify the cross-linking process.  相似文献   

15.
Low-density lipoproteins (LDL) play a key role in the delivery of photosensitizers to tumor cells in photodynamic therapy. The interaction of deuteroporphyrin, an amphiphilic porphyrin, with LDL is examined at equilibrium and the kinetics of association/dissociation are determined by stopped-flow. Changes in apoprotein and porphyrin fluorescence suggest two classes of bound porphyrins. The first class, characterized by tryptophan fluorescence quenching, involves four well-defined sites. The affinity constant per site is 8.75 x 10(7) M(-1) (cumulative affinity 3.5 x 10(8) M(-1)). The second class corresponds to the incorporation of up to 50 molecules into the outer lipidic layer of LDL with an affinity constant of 2 x 10(8) M(-1). Stopped-flow experiments involving direct LDL porphyrin mixing or porphyrin transfer from preloaded LDL to albumin provide kinetic characterization of the two classes. The rate constants for dissociation of the first and second classes are 5.8 and 15 s(-1); the association rate constants are 5 x 10(8) M(-1) s(-1) per site and 3 x 10(9) M(-1) s(-1), respectively. Both fluorescence and kinetic analysis indicate that the first class involves regions at the boundary between lipids and the apoprotein. The kinetics of porphyrin-LDL interactions indicates that changes in the distribution of photosensitizers among various carriers could be very sensitive to the specific tumor microenvironment.  相似文献   

16.
Reactivation of tetrameric porcine skeletal muscle lactic dehydrogenase after dissociation and extensive unfolding of the monomers by 6 M guanidine hydrochloride (Gdn . HCl) is characterized by sigmoidal kinetics, indicating a complex mechanism involving rate-limiting folding and association steps. For analysis of the association reactions, chemical cross-linking with glutaraldehyde may be used [Hermann, R., Jaenicke, R., & Rudolph, R. (1981) Biochemistry 20, 2195-2201]. The data clearly show that the formation of a dimeric intermediate is determined by a first-order folding reaction of the monomers with k1 = (8.0 +/- 0.1) x 10(-4) s-1. The rate constant of the association of dimers to tetramers which represents the second rate-limiting step on the pathway of reconstitution after guanidine denaturation, was then determined by reactivation and cross-linking experiments after dissociation in 0.1 M H3PO4 containing 1 M Na2SO4. The rate constant for the dimer association (which is the only rate-limiting step after acid dissociation) was k2 = (3.0 +/- 0.5) x 10(4) M-1 s-1. On the basis of the given two rate constants, the complete reassociation pattern of porcine lactic dehydrogenase after dissociation and denaturation in 6 M Gdn . HCl can be described by the kinetic model (formula: see text).  相似文献   

17.
18.
The first two reactions in the phosphotransfer sequence of bacterial phosphoenolpyruvate:glycose phosphotransferase systems are the autophosphorylation of Enzyme I by phosphoenolpyruvate followed by the transfer of the phospho group to the low-molecular weight protein, HPr. Transient state kinetic methods were used to estimate the second-order rate constants for both phosphotransfer reactions. These measurements support previous conclusions that only the dimer of Enzyme I, EI2, is autophosphorylated, and that the rate of formation of dimer is slow compared to the rate of its phosphorylation. The rate constants of the two autophosphorylation reactions of EI2 by PEP are 6.6 x 10(6) M(-1) s(-1), and differ from one another by a factor of less than 3. The rate constant for the transfer reaction between phospho-EI2 and HPr is unusually large for a covalent reaction between two proteins (220 x 10(6) M(-1) s(-1)), while the constant for the reverse reaction is 4.2 x 10(6) M(-1) s(-1). Using the previously reported equilibrium constant for the autophosphorylation reaction, 1.5, the overall equilibrium constant for phosphotransfer from PEP to HPr is 80, somewhat higher than that previously reported. The results also show that EI2 can phosphorylate multiple molecules of HPr without dissociating to a monomer (EI), and that EI can accept a phospho group from phospho-HPr. These results are directly applicable to predicting the rates of phosphoenolpyruvate phosphotransferase system sugar uptake in whole cells.  相似文献   

19.
All-trans-retinoyl chloromethane is a potent irreversible inactivator of bovine opsin in retinal rod outer segments. The inactivation appears to be due to specific modification of the apoprotein at the 11-cis-retinaldehyde binding site. The reaction follows pseudo first order kinetics at 37 degrees C. A moderate dissociation constant for the initial reversible complex of 6.1 mM could be derived with a first order rate constant of 1.8x10(-1) min-1 for the conversion to the irreversible inactivated form. Native rhodopsin or rhodopsin regenerated from opsin by addition of 11-cis-retinaldehyde is completely protected against inactivation by trans-retinoyl chloromethane. All-trans-retinaldehyde does not provide this protection for the irreversible inactivation.  相似文献   

20.
The time course of electron transfer in vitro between soluble domains of the Rieske iron-sulfur protein (ISP) and cytochrome f subunits of the cytochrome b(6)f complex of oxygenic photosynthesis was measured by stopped-flow mixing. The domains were derived from Chlamydomonas reinhardtii and expressed in Escherichia coli. The expressed 142-residue soluble ISP apoprotein was reconstituted with the [2Fe-2S] cluster. The second-order rate constant, k(2)((ISP-f)) = 1.5 x 10(6) m(-1) s(-1), for ISP to cytochrome f electron transfer was <10(-2) of the rate constant at low ionic strength, k(2)((f-PC))(> 200 x 10(6) m(-1) s(-1)), for the reduction of plastocyanin by cytochrome f, and approximately 1/30 of k(2)((f-PC)) at the ionic strength estimated for the thylakoid interior. In contrast to k(2)((f-PC)), k(2)((ISP-f)) was independent of pH and ionic strength, implying no significant role of electrostatic interactions. Effective pK values of 6.2 and 8.3, respectively, of oxidized and reduced ISP were derived from the pH dependence of the amplitude of cytochrome f reduction. The first-order rate constant, k(1)((ISP-f)), predicted from k(2)((ISP-f)) is approximately 10 and approximately 150 times smaller than the millisecond and microsecond phases of cytochrome f reduction observed in vivo. It is proposed that in the absence of electrostatic guidance, a productive docking geometry for fast electron transfer is imposed by the guided trajectory of the ISP extrinsic domain. The requirement of a specific electrically neutral docking configuration for ISP electron transfer is consistent with structure data for the related cytochrome bc(1) complex.  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