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.
We have studied the effect of GDP and its analog guanyl-5'-yl thiophosphate (GDP beta S) on the interaction between rhodopsin and transducin (Gt). Stabilization of the light-induced active intermediate, metarhodopsin II (MII), by bound Gt (extra-MII effect) monitored the catalytic interaction between the proteins. Extra-MII can be completely abolished by GDP, with a half-suppression at 10 microM under the conditions (4 degrees C, pH 8, 7.5 nM photoactivated rhodopsin). The effect of GDP did not depend on divalent cations, in contrast to GTP-induced dissociation of the complex. The GDP analog GDP beta S did not affect extra-MII although it binds to the MII-Gt complex with only three times lower affinity (reversal of the GDP effect by GDP beta S). However, GDP beta S enhanced considerably the efficiency of synthetic rhodopsin peptide competition against the formation of extra-MII. GDP and GDP beta S slow the Gt activation rate (monitored by kinetic light scattering), with the same relative efficiencies. We therefore assume that GDP, GDP beta S, and GTP bind at the same site. We discuss a generalized induced fit mechanism, where MII induces opening of the Gt nucleotide site and release of GDP which in turn is obligatory to establish the MII-stabilizing rhodopsin-Gt three-loop interaction (K?nig, B., Arendt, A., McDowell, J.H., Kahlert, M., Hargrave, P.A., and Hofmann, K.P. (1989) Proc. Natl. Acad. Sci. U.S.A. 86, 6878-6882). The GDP beta S/GDP difference is discussed in terms of bound GDP disturbing the interaction with two and GDP beta S with only one of the rhodopsin binding sites. Mechanistically, our results indicate a critical role of the beta-phosphate interaction with the nucleotide binding site in the GDP-induced transformation of Gt.  相似文献   

2.
S K Gibson  J H Parkes  P A Liebman 《Biochemistry》1999,38(34):11103-11114
Phosphorylation reduces the lifetime and activity of activated G protein-coupled receptors, yet paradoxically shifts the metarhodopsin I-II (MI-MII) equilibrium (K(eq)) of light-activated rhodopsin toward MII, the conformation that activates G protein. In this report, we show that phosphorylation increases the apparent pK for MII formation in proportion to phosphorylation stoichiometry. Decreasing ionic strength enhances this effect. Gouy-Chapman theory shows that the change in pK is quantitatively explained by the membrane surface potential, which becomes more negative with increasing phosphorylation stoichiometry and decreasing ionic strength. This lowers the membrane surface pH compared to the bulk pH, increasing K(eq) and the rate of MII formation (k(1)) while decreasing the back rate constant (k(-)(1)) of the MI-MII relaxation. MII formation has been observed to depend on bulk pH with a fractional stoichiometry of 0.6-0.7 H(+)/MII. We find that the apparent fractional H(+) dependence is an artifact of altering the membrane surface charge during a titration, resulting in a fractional change in membrane surface pH compared to bulk pH. Gouy-Chapman calculations of membrane pH at various phosphorylation levels and ionic strengths suggest MII formation behavior consistent with titration of a single H(+) binding site with 1:1 stoichiometry and an intrinsic pK of 6.3 at 0.5 degrees C. We show evidence that suggests this same site has an intrinsic pK of 5.0 prior to light activation and its protonation before activation greatly enhances the rate of MII formation.  相似文献   

3.
Cytosolic and nuclear estrogen receptor forms of chicken oviduct have been studied by (1) measuring hormone dissociation kinetics and by (2) sucrose density gradient analysis on high salt gradients. Estradiol dissociates from the receptor in chicken oviduct cytosol at 22 degrees C following a two-phase exponential process. The fraction of receptor with a fast dissociation rate (k = 120 X 10(-3) min-1) decreases as a function of the pre-incubation at 22 degrees C; after prolonged pre-incubation only the slowly dissociating (k = 12.3 X 10(-3) min-1) form remains. Dissociation of moxestrol, a synthetic estrogen with a higher affinity, from the cytosol receptor at 30 degrees C is similar, showing a transition of a fast dissociating form (k = 120 X 10(-3) min-1) to a slowly dissociating form (k = 7.6 X 10(-3) min-1) as a result of pre-incubation at 30 degrees C. A concomitant temperature-dependent shift of the estrogen receptor from a 4.8 S to a 6.1 S form was observed with moxestrol but not with estradiol as a ligand. Sodium molybdate (20 mM) and NaSCN (400 mM) inhibit the temperature-dependent increase in sedimentation coefficient, but molybdate allows the formation of a receptor form which shows intermediary dissociation kinetics. Estrogen receptor, precipitated with ammonium sulfate (0-35%) shows monophasic dissociation kinetics of estradiol (k = 39.5 X 10(-3) min-1) and for moxestrol (k = 10.8 X 10(-3) min-1), suggesting full receptor activation only with moxestrol as a ligand. Moxestrol-receptor complexes obtained by ammonium sulfate precipitation sediment at 0 degree C at 4.8 S. Only after subsequent incubation at 30 degrees C a shift from 4.8 S to 5.9 S is observed, suggesting that the formation of the slowly dissociating form of the receptor may precede the formation of a stable transformed receptor complex. The nuclear estrogen receptor with estradiol as a ligand shows biphasic dissociation kinetics at 22 degrees C (k = 70 X 10(-3) min-1; k = 14.0 X 10(-3) min-1). The ratio of both components (1:1) does not change after preincubation of the nuclear receptor extract at 22 degrees C. Moxestrol dissociates from the nuclear receptor at 30 degrees C monophasically with a slow rate (k = 6.1 X 10(-3) min-1), suggesting that it is extracted as an activated hormone-receptor complex.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

4.
Temperature dependence of photoinhibition and photoprotective mechanisms (10-35 degrees C) was investigated for Chenopodium album leaves grown at 25 degrees C under 500 micro mol quanta m(-2) s(-1). The fraction of active photosystem II (PSII) was determined after photoinhibitory treatment at different temperatures in the presence and absence of lincomycin, an inhibitor of chloroplast-encoded protein synthesis. In the absence of lincomycin, leaves were more tolerant to photoinhibition at high (25-35 degrees C) than at low (11-15 degrees C) temperatures. In the presence of lincomycin, the variation in the tolerance to photoinactivation became relatively small. The rate constant of photoinactivation (k(pi)) was stable at 25-35 degrees C and increased by 50% with temperature decrease from 25 to 11 degrees C. The rate constant of recovery of inactivated PSII (k(rec)) was more sensitive to temperature; it was very low at 11 degrees C and increased by an order of magnitude at 35 degrees C. We conclude that the recovery of photoinactivated PSII plays an essential role in photoprotection at 11-35 degrees C. Partitioning of light energy to various photoprotective mechanisms was further analyzed to reveal the factor responsible for k(pi). The fraction of energy utilized in photochemistry was lower at lower temperatures. Although the fraction of heat dissipation increased with decreasing temperatures, the excess energy that is neither utilized by photochemistry nor dissipated by heat dissipation was found to be greater at lower temperatures. The k(pi) value was strongly correlated with the excess energy, suggesting that the excess energy determines the rate of photoinactivation.  相似文献   

5.
High affinity agonist binding to G protein-coupled receptors depends on the formation of a ternary complex between agonist, receptor, and G protein. This process is too slow to be accounted for by a simple diffusion-controlled mechanism. We have tested if the interaction between activated receptor and G protein is rate-limiting by fusing the coding sequence of the human A(1)-adenosine receptor to that of Galpha(i-1) (A(1)/Galpha(i-1)) and of Galpha(o) (A(1)/Galpha(o)). Fusion proteins of the expected molecular mass were detected following transfection of HEK293 cells. Ternary complex formation was monitored by determining the kinetics for binding of the high affinity agonist (-)-N(6)-3[(125)I](iodo-4-hydroxyphenylisopropyl)adenosine; these were similar in the wild-type receptor and the fusion proteins over the temperature range of 10 to 30 degrees C. Agonist dissociation may be limited by the stability of the ternary complex. This assumption was tested by creating fusion proteins in which the Cys(351) of Galpha(i-1) was replaced with glycine (A(1)/Galpha(i-1)C351G) or isoleucine (A(1)/Galpha(i-1)C351I) to lower the affinity of the receptor for the G protein. In these mutated fusion proteins, the dissociation rate of the ternary complex was accelerated; in contrast, the rate of the forward reaction was not affected. We therefore conclude that (i) receptor activation per se rather than its interaction with the G protein is rate-limiting in ternary complex formation; (ii) the stability of the ternary complex is determined by the dissociation rate of the G protein. These features provide for a kinetic proofreading mechanism that sustains the fidelity of receptor-G protein coupling.  相似文献   

6.
The iron-regulated surface determinants (Isd) of Staphylococcus aureus, including surface proteins IsdA, IsdB, IsdC, and IsdH and ATP-binding cassette transporter IsdDEF, constitute the machinery for acquiring heme as a preferred iron source. Here we report hemin transfer from hemin-containing IsdA (holo-IsdA) to hemin-free IsdC (apo-IsdC). The reaction has an equilibrium constant of 10 +/- 5 at 22 degrees C in favor of holo-IsdC formation. During the reaction, holo-IsdA binds to apo-IsdC and then transfers the cofactor to apo-IsdC with a rate constant of 54.3 +/- 1.8 s(-1) at 25 degrees C. The transfer rate is >70,000 times greater than the rate of simple hemin dissociation from holo-IsdA into solvent (k transfer = 54.3 s(-1) versus k -hemin = 0.00076 s(-1)). The standard free energy change, Delta G 0, is -27 kJ/mol for the formation of the holo-IsdA-apo-IsdC complex. IsdC has a higher affinity for hemin than IsdA. These results indicate that the IsdA-to-IsdC hemin transfer is through the activated holo-IsdA-apo-IsdC complex and is driven by the higher affinity of apo-IsdC for the cofactor. These findings demonstrate for the first time in the Isd system that heme transfer is rapid, direct, and affinity-driven from IsdA to IsdC. These results also provide the first example of heme transfer from one surface protein to another surface protein in Gram-positive bacteria and, perhaps most importantly, indicate that the mechanism of activated heme transfer, which we previously demonstrated between the streptococcal proteins Shp and HtsA, may apply in general to all bacterial heme transport systems.  相似文献   

7.
Rheumatoid factors (RFs) are autoantibodies, which recognize antigens on a constant region of immunoglobulin G (IgG). Among various RF classes, RF of the IgG class (IgGRF) forms immune complexes in rheumatoid joints and is implicated in the pathogenesis of rheumatoid arthritis (RA). To characterize the formation of IgGRF immune complexes, in the present study, IgGRF was isolated from sera of RA patients, and its interaction with immobilized IgG was analyzed and compared to that of IgMRF or IgARF by means of surface plasmon resonance. On gel filtration, the IgGRF was eluted as a single peak corresponding to IgG, excluding the possible formation of self-associating IgGRF complexes in solution. Sensorgrams of the interaction of IgGRF with immobilized IgG revealed that it clearly bound to the IgG at 6 degrees C, but not at 30 degrees C. The degree of interaction decreased inversely with an increase in temperature, suggesting that IgGRF is much more reactive at lower temperatures. In contrast, the interaction of IgARF and IgMRF with IgG at 6 degrees C was similar to that at 30 degrees C. The association rate constant (k(a)) of IgGRF decreased with an increase in temperature, while those of IgARF and IgMRF were similar under various thermal conditions. The dissociation rate constant (k(d)) of IgGRF was greatly reduced at 25 degrees C, but those of IgARF and IgMRF slightly increased with an increase in temperature. These results suggested that the mode of interaction of IgGRF with IgG differed from in the cases of IgMRF and IgARF. The kinetic properties of the IgGRF-IgG interaction may facilitate elucidation of the IgGRF immune complex formation in rheumatoid joints.  相似文献   

8.
The chromium moiety of gamma,beta-bidentate CrATP slowly accepts a ligand from the sarcoplasmic reticulum Ca-ATPase to form an exchange inert coordination complex (k + 1 = 0.083 min-1; k - 2 = 0.003 min-1, 37 degrees C, 100 microM CaCl2). The stability of the Cr3+ coordinate bonds allowed the complex to be isolated by filtration techniques at neutral pH without acid precipitation. We found 4-5 nmol of [gamma-32P]CrATP to bind to 1 mg of sarcoplasmic reticulum protein with the subsequent occlusion of 7-8 nmol of 45Ca2+. At 37 degrees C, the CrATP.ATPase complex could be formed in the absence of Ca2+, although the reaction was 2-3 times slower than in the presence of Ca2+. Inhibition by Pi, by orthovanadate, and by fluorescein 5'-isothiocyanate verified that the bound CrATP was at the catalytic site. The site of CrATP attachment was found to be on the A tryptic fragment, possibly on the A2 subfragment. It was determined that Ca2+ binding to high affinity sites on the enzyme controls the rate by which the Cr3+ moiety accepts the ligand from the enzyme. The rate of change in the EPR spectrum of iodoacetamide spin-labeled ATPase was shown to follow the rate of ligand acceptance, rather than the binding of Ca2+ and substrate per se. This particular change has been attributed to the formation of an activated complex that is immediately precursory to phosphorylation and indicates here that this complex cannot be properly formed until the metal has been chelated by the enzyme. It is concluded that control over metal chelation (Cr3+ here, Mg2+ in the normal mechanism) is one means by which Ca2+ activates the enzyme.  相似文献   

9.
A water-soluble yellow protein from E. halophila was previously shown to be photoactive (Meyer, T. E., E. Yakali, M. A. Cusanovich, and G. Tollin. 1987. Biochemistry. 26:418-423). Pulsed laser excitation in the protein visible absorption band (maximum at 445 nm) causes a rapid bleach of color (k = 7.5 x 10(3) s-1) followed by a slower dark recovery (k = 2.6 s-1). This is analogous to the photocycle of sensory rhodopsin II from Halobacterium (which also has k = 2.6 s-1 for recovery). We have now determined the quantum yield of the photobleaching process to be 0.64, which is comparable with that of bacteriorhodopsin (0.25), and is thus large enough to be biologically significant. Although the photoreactions of yellow protein were previously shown to be relatively insensitive to pH, ionic strength and the osmoregulator betaine, the present experiments demonstrate that temperature, glycerol, sucrose, and various alcohol-water mixtures strongly influence the kinetics of photobleaching and recovery. The effect of temperature follows normal Arrhenius behavior for the bleach reaction (Ea = 15.5 kcal/mol). The rate constant for the recovery reaction increases with temperature between 5 degrees C and 35 degrees C, but decreases above 35 degrees C indicating alternate conformations with differing kinetics. There is an order of magnitude decrease in the rate constant for photobleaching in both glycerol and sucrose solutions that can be correlated with the changes in viscosity. We conclude from this that the protein undergoes a conformational change as a consequence of the photoinduced bleach. Recovery kinetics are affected by glycerol and sucrose to a much smaller extent and in a more complicated manner.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
This study assessed the impact of various cryoprotectant (CPA) exposures on nuclear and cytoplasmic maturation in the immature cat oocyte as a prerequisite to formulating a successful cryopreservation protocol. In experiment 1, immature oocytes were exposed to 0, 0.75, 1.5, or 3.0 M of 1,2-propanediol (PrOH) or 1,2-ethanediol (EG) at room temperature (25 degrees C) or 0 degrees C for 30 min. After CPA removal and in vitro maturation, percentage of oocytes reaching metaphase II (MII) was reduced after exposure to 3.0 M PrOH at 0 degrees C or 3.0 M EG at both temperatures. All CPA exposures increased MII spindle abnormalities compared to control, except 1.5 M PrOH at 25 degrees C. In experiments 2 and 3, immature oocytes were exposed to CPA conditions yielding optimal nuclear maturation that either had caused spindle damage (0.75 M PrOH, 1.5 M EG, and 3.0 M PrOH at 25 degrees C) or not (1.5 M PrOH at 25 degrees C). After maturation and insemination in vitro, oocytes were cultured for 7 days to assess treatment influence on developmental competence. CPA exposure did not affect fertilization, but the high incidence of MII spindle abnormalities resulted in a low percentage of cleaved embryos. Blastocyst formation and quality were influenced by both CPA types (EG was more detrimental than PrOH) and concentration (3.0 M was more detrimental than 1.5 M). Overall, cat oocytes appear to be highly sensitive to CPA except after exposure to 1.5 M PrOH at 25 degrees C, a treatment that still allowed approximately 60% of the oocytes to reach MII and approximately 20% to form blastocysts.  相似文献   

11.
Purified mitochondrial malate dehydrogenase isoenzyme (m-MDH) of Toxocara canis muscle presented maximum activity at 48 degrees C. A clear change in slope of the Arrhenius plot was observed. The energy of activation calculated for the catalytic process showed values of 3.2 kcal/mol and 10.5 kcal/mol. Thermal inactivation of m-MDH showed that it is more thermolabile than the s-isoenzyme. The inactivation of the enzyme by heat could be reduced at least in part by the addition of 0.1 mM NADH. The heat denaturation showed to be a first-order process. The rate constant (k) was calculated as being of the order of 5.28 X 10(-4) s-1 at 40 degrees C. The activation energy for the heat inactivation process was 16.45 kcal/mol between 30 degrees C and 40 degrees C and 13.79 kcal/mol between 40 degrees C and 48 degrees C.  相似文献   

12.
The thermal activation barrier of guanosine triphosphate dependent dissociation of the light-induced rhodopsin-G-protein complex has been determined using a spectroscopic technique (enhanced formation of metarhodopsin II). The dissociation rate has been measured in the range - 2 degrees C less than or equal to t less than or equal to 12 degrees C. The Arrhenius plot yields apparent activation energies: 166 +/- 10 kJmol-1 with 5'-guanylylimidodiphosphate (GMPPNP) and 175 +/- 15 kJmol-1 with GTP. The rhodopsin-G-protein dissociation rate is linearly related to the concentration of GMPPNP in the measurable range (less than or equal to 200 microM). The data show that, at low temperature (1 degree C), the rate limiting step of G-protein activation is the bimolecular reaction between the protein and the nucleotide. This also seems to hold true for more physiological conditions as suggested by extrapolation and comparison with nucleotide exchange rates in the literature. The high activation barrier of the nucleotide exchange reaction is explained in terms of rapid endothermic preequilibrium between an inactive and an exchanging state of the rhodopsin-G-protein complex.  相似文献   

13.
The kinetics and energetics of the binding between barley alpha-amylase/subtilisin inhibitor (BASI) or BASI mutants and barley alpha-amylase 2 (AMY2) were determined using surface plasmon resonance and isothermal titration calorimetry (ITC). Binding kinetics were in accordance with a 1:1 binding model. At pH 5.5, [Ca(2+)] = 5 mM, and 25 degrees C, the k(on) and k(off) values were 8.3 x 10(+4) M(-1) s(-1) and 26.0 x 10(-4) s(-1), respectively, corresponding to a K(D) of 31 nM. K(D) was dependent on pH, and while k(off) decreased 16-fold upon increasing pH from 5.5 to 8.0, k(on) was barely affected. The crystal structure of AMY2-BASI shows a fully hydrated Ca(2+) at the protein interface, and at pH 6.5 increase of [Ca(2+)] in the 2 microM to 5 mM range raised the affinity 30-fold mainly due to reduced k(off). The K(D) was weakly temperature-dependent in the interval from 5 to 35 degrees C as k(on) and k(off) were only increasing 4- and 12-fold, respectively. A small salt dependence of k(on) and k(off) suggested a minor role for global electrostatic forces in the binding and dissociation steps. Substitution of a positively charged side chain in the mutant K140L within the AMY2 inhibitory site of BASI accordingly did not change k(on), whereas k(off) increased 13-fold. ITC showed that the formation of the AMY2-BASI complex is characterized by a large exothermic heat (Delta H = -69 +/- 7 kJ mol(-1)), a K(D) of 25 nM (27 degrees C, pH 5.5), and an unfavorable change in entropy (-T Delta S = 26 +/- 7 kJ mol(-1)). Calculations based on the thermodynamic data indicated minimal structural changes during complex formation.  相似文献   

14.
The role of membrane composition in modulating the rate of G protein-receptor complex formation was examined using rhodopsin and transducin (G(t)) as a model system. Metarhodopsin II (MII) and MII-G(t) complex formation rates were measured, in the absence of GTP, via flash photolysis for rhodopsin reconstituted in 1-stearoyl-2-oleoyl-sn-glycero-3-phosphocholine (18:0,18:1PC) and 1-stearoyl-2-docosahexaenoyl-sn-glycero-3-phosphocholine (18:0,22:6PC) bilayers, with and without 30 mol% cholesterol. Variation in bilayer lipid composition altered the lifetime of MII-G(t) formation to a greater extent than the lifetime of MII. MII-G(t) formation was fastest in 18:0,22:6PC and slowest in 18:0,18:1PC/30 mol% cholesterol. At 37 degrees C and a G(t) to photolyzed rhodopsin ratio of 1:1 in 18:0,22:6PC bilayers, MII-G(t) formed with a lifetime of 0.6 +/- 0.06 ms, which was not significantly different from the lifetime for MII formation. Incorporation of 30 mol% cholesterol slowed the rate of MII-G(t) complex formation by about 400% in 18:0,18:1PC, but by less than 25% in 18:0,22:6PC bilayers. In 18:0,22:6PC, with or without cholesterol, MII-G(t) formed rapidly after MII formed. In contrast, cholesterol in 18:0,18:1PC induced a considerable lag time in MII-G(t) formation after MII formed. These results demonstrate that membrane composition is a critical factor in determining the temporal response of a G protein-coupled signaling system.  相似文献   

15.
A previous study showed that with hypertonic sucrose treatment, a projection is formed in mouse metaphase II (MII) oocytes in proximity to the spindle and chromosomes, where a polarized cortical domain is located. However, little is known about the mechanisms involved in this process. Here, we designed a series of experiments to test the hypothesis that hypertonicity is the induction factor for the formation of projections in mouse MII oocytes. Our hypothesis was supported by the following evidence: 1) different concentrations of sucrose affected the formation and shape of projections, whereas serum or basic media had little effect; 2) other hypertonic sugar solutions could also induce projection formation; and 3) projections could also be induced by hypertonic NaCl solution. We then tested the hypothesis that the cytoskeleton was involved in the formation of hypertonicity-induced projections. This was investigated by culturing MII- and germinal vesicle-stage mouse oocytes in the presence or absence of cytoskeletal inhibitors, including cytochalasin B (disruption of actin filaments), nocodazole (disruption of microtubules), and taxol (polymerization of tubulin molecules). We found that none of the cytoskeletal inhibitors alone could prevent hypertonicity-induced projection formation, whereas the combination of cytochalasin B with nocodazole or with taxol blocked the formation of these projections in most matured oocytes. When immature oocytes were incubated in cytochalasin B, but not in nocodazole or taxol, the formation of an actin-rich domain and the peripheral positioning of the spindle were blocked during maturation; hence, no projections were formed, even after hypertonic sucrose treatment. Based on these observations, we propose that three components are necessary for projection formation: 1) a polarized cortical patch (e.g., an actin-rich domain), 2) rigid submembrane structures (e.g., a spindle and/or chromosomes), and 3) solid connections between the above. Any disturbance of one of these factors will affect the hypertonicity-induced projection formation. Hypertonicity-induced projection in mouse oocytes thus provides an experimental model for studies regarding cell polarity and the interaction between membrane and submembrane components.  相似文献   

16.
A method was developed to purify a 30-kDa protein from jelly fig (Ficus awkeotsang) pericarp, including preparation of jelly curd from achenes, extraction of proteins from the curd, and isolation of the 30-kDa protein by anion-exchanger and gel filtration. Chitinase activity was detected in the purified 30-kDa protein by activity staining in both non-denaturing gel electrophoresis and SDS-PAGE. Isoelectrofocusing showed that the isoelectric point of the 30-kDa protein was lower than pH 3.5. The K(m), k(cat), optimal pH and temperature of this putative chitinase were determined to be 0.076 mM, 0.089 s(-1), pH 4, and 60 degrees C, respectively. The purified 30-kDa protein was thermostable (retaining activity up to 65 degrees C for several hours) and could be stored at 4 degrees C for a year without apparent loss of chitinase activity. Antifungal activity of this putative chitinase was measured in terms of inhibition of Colletotrichum gloeosporioides spore germination.  相似文献   

17.
We have isolated a chaperonin from the hyperthermophilic archaeon Sulfolobus solfataricus based on its ability to inhibit the spontaneous refolding at 50 degrees C of dimeric S. solfataricus malic enzyme. The chaperonin, a 920-kDa oligomer of 57-kDa subunits, displays a potassium-dependent ATPase activity with an optimum temperature at 80 degrees C. S. solfataricus chaperonin promotes correct refoldings of several guanidine hydrochloride-denatured enzymes from thermophilic and mesophilic sources. At a molar ratio of chaperonin oligomer to single polypeptide chain of 1:1, S. solfataricus chaperonin completely inhibits spontaneous refoldings and suppresses aggregation upon dilution of the denaturant; refoldings resume upon ATP hydrolysis, with yields of active molecules and rates of folding notably higher than in spontaneous processes. S. solfataricus chaperonin prevents the irreversible inactivations at 90 degrees C of several thermophilic enzymes by the binding of the denaturation intermediate; the time-courses of inactivations are unaffected and most activity is regained upon hydrolysis of ATP. S. solfataricus chaperonin completely prevents the formation of aggregates during thermal inactivation of chicken egg white lysozyme at 70 degrees C, without affecting the rate of activity loss; ATP hydrolysis results in the recovery of most lytic activity. Tryptophan fluorescence measurements provide evidence that S. solfataricus chaperonin undergoes a dramatic conformational rearrangement in the presence of ATP/Mg, and that the hydrolysis of ATP is not required for the conformational change. The ATP/Mg-induced conformation of the chaperonin is fully unable to bind the protein substrates, probably due to disappearance or modification of the substrate binding sites. This is the first archaeal chaperonin whose involvement in protein folding has been demonstrated.  相似文献   

18.
We studied the nature of the interaction of delta-hexachlorocyclohexane (delta-HCCH), a pesticide having a stereoisomeric structure similar to inositol, with red blood cells. Cell survival data, measured as percent of hemoglobin released by delta-HCCH, show that the cell lysis increases with post exposure time. delta-HCCH at 55-60 micrograms/ml causes about 70% cell lysis after 24 h of exposure. The nature of interaction of delta-HCCH with membrane components was evaluated by studying the thermotropic transitions and protein structure of ghosts using Raman spectroscopy. Control ghosts show transitions with onset/completion temperatures 30 degrees C/38 degrees C (high temperature transition) and 3 degrees C/10 degrees C (middle temperature transition) when monitored by the I2935/I2850 ratio. The interaction of delta-HCCH drastically broadens the high temperature transition and shifts it to the temperature range of 10-29 degrees C. The plots of (I2880-90/I2850) vs. temperature show two transitions for control ghosts, one extending from -10 degrees C to 3 degrees C (lower temperature transition) and the other from about 7 degrees C to about 15 degrees C (middle temperature transition). Ghosts lysed with delta-HCCH shows only a single and a very broad transition in the range of about -3 degrees C to about 15 degrees C. These changes in the thermal transition properties suggest that delta-HCCH alters lipid and lipid-protein phases of erythrocyte membranes. The comparison of Raman spectra in the amide I and III regions of erythrocyte ghosts and purified band 3 with several amidated compounds reveals that cytoskeleton proteins contain highly amidated residues (probably glutamine and asparagine). The interaction of delta-HCCH with erythrocytes drastically alters the environment of these amidated residues indicating the involvement of cytoskeleton proteins. We conclude that the interaction of delta-HCCH with red blood cells disrupt membrane structure and change the environment of cytoskeleton proteins that could cause cell lysis.  相似文献   

19.
R Koren  G G Hammes 《Biochemistry》1976,15(5):1165-1171
Kinetic studies have been carried out of the monomer-dimer interaction of insulin, beta-lactoglobulin, and alpha-chymotrypsin using stopped-flow and temperature-jump techniques. The pH indicators bromothymol blue, bromophenol blue, and phenol red were used to monitor pH changes associated with the monomer-dimer interaction. In all three cases a kinetic process was observed which could be attributed to a simple monomer-dimer equilibrium, and association (k1) and dissociation (k-1) rate constants were determined. The results obtained are as follows: for insulin at 23 degrees C, pH 6.8, 0.125 M KNO3, k1 = 1.14 X 10(8) M-1 s-1, k-1 - 1.48 X 10(4)s(-1); for beta-lactoglobulin AB at 35 degrees C, pH 3.7, 0.025 M KNO3, d1 = 4.7 X 10(4) M-1 s-1, k-1 = 2.1 s-1; for alpha-chymotrypsin at 25 degreesC, pH 4.3, 0.05 M KNO3 k1 - 3.7 X 10(3) M-1 s-1, k-1 - 0.68 s-1. The kinetic behavior of the separated beta-lactoglobulin A and B was similar to that of the mixture. In the case of chymotrypsin, bromophenol blue was found to activate the enzyme catalyzed hydrolysis of p-nitrophenyl acetate, and a rate process was observed with the temperature jump which could be attributed to a conformational change of the indicator-protein complex. The association rate constant for dimer formation of insulin approaches the value expected for a diffusion-controlled process, while the values obtained for the other two proteins are below those expected for a diffusion-controlled reaction unless unusally large steric and electrostatic effects are present.  相似文献   

20.
S G Rhee  P B Chock 《Biochemistry》1976,15(8):1755-1760
The kinetics of protein fluorescence change exhibited by ADP or orthophosphate addition to the Mg2+-or Mn2+-activated unadenylylated glutamine synthetase from Escherichia coli were studied. The kinetic patterns of these reactions are incompatible with a simple bimolecular binding process and a mechanism which required protein isomerization prior to substrate binding. They are consistent with a mechanism in which direct substrate binding is followed by a substrate-induced conformational change step, ES in equilibrium ES. At pH 7.0 and 15 degrees C, the association constants for the direct binding (K1) of ADP to MnE1.0 and of Pi to MnE1.0ADP are 3.9 X 10(4) and 2.28 X 10(2) M(-1), respectively. The association constant for the direct binding of ADP to MnE1.0Pi is 2.3 X 10(4) M(-1) at pH 7.0 and 19 degrees C. The deltaG degrees for the substrate-induced conformational step are -3.5 and -1.3 kcal mol(-1) due to ADP binding to MnE1.0Pi and MnE1.0, respectively, and -1.4 kcal mol(-1) due to Pi binding to MnE1.0ADP. Rate constants, k2 and k(-2), for the isomerization step are: 90 and 9.5 s(-1) for ADP binding to MnE1.0, 440 and 0.36 s(-1) for ADP binding to MnE1.0Pi, and 216 and 1.8 s(-1) for Pi binding to MnE1.0ADP. Due to low substrate affinity, the association constant for direct Pi binding to MnE1.0 was roughly estimated to be 230 M(-1) and k2 = 750 s(-1), k(-2) = 250 s(-1). At 9 degrees C and pH 7.0, the estimated association constants for the direct ADP binding to MgE1.0 and MgE1.0 Pi are 1.8 X 10(4) and 1.6 X 10(4) M(-1), respectively; and the rate constants for the isomerization step associated with the corresponding reaction are k2 = 550 s(-1), k(-2) = 500 s(-1), and k2 = 210 s(-1), k(-2) = 100 s(-1). From the kinetic analysis it is evident that the inability of Mn2+ to support biosynthetic activity of the unadenylylated enzyme is due to the slow rate of ADP release from the MnE1.0PiADP complex. In contrast the large k(-2) obtained for ADP release from the MgE1.0ADP or MgE1.0PiADP complex indicates that this step is not rate limiting in the biosynthesis of glutamine since the k catalysis obtained under the same conditions is 7.2 s(-1).  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