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.
M Zolkiewski  A Ginsburg 《Biochemistry》1992,31(48):11991-12000
Dodecameric glutamine synthetase (GS) from Escherichia coli undergoes reversible, thermally induced partial unfolding without subunit dissociation. A single endotherm for Mn.GS (+/- active-site ligands) in the presence of 1 mM free Mn2+ and 100 mM KCl at pH 7 is observed by differential scanning calorimetry (DSC). Previous deconvolutions of DSC data for Mn.GS showed only two two-state transitions (with similar tm values; 51.6 +/- 2 degrees C), and indicated that cooperative interactions link partial unfolding reactions of all subunits within the Mn.enzyme dodecamer [Ginsburg, A., & Zolkiewski, M. (1991) Biochemistry 30, 9421]. A net uptake of 8.0 equiv of H+ by Mn.GS occurs during partial unfolding, as determined in the present DSC experiments conducted with four buffers having different heats of protonation at 50 degrees C. These data gave a value of 176 +/- 12 kcal (mol of dodecamer)-1 for delta Hcal corrected for buffer protonation. L-Glutamine and L-Met-(SR)-sulfoximine stabilize the Mn.GS dodecamer through the free energies of ligand binding, and these were shown to be partially and totally released, respectively, from the 12 active sites at high temperature. Ligand effects on Tm values from DSC were similar to those from spectral measurements of Trp and Tyr exposures in two subunit domains. Effects of varying [ADP] on DSC profiles of Mn.GS were complex; Tm is increased by low [ADP] and decreased by > 100 microM free ADP. This is due to the exposure of an additional low-affinity ADP binding site per GS subunit at high temperature with log K1' = 4.3 and log K2' = 3.6 at 60 degrees C relative to log K' = 5.5 for ADP binding at 30 degrees C, as determined by isothermal calorimetric and fluorescence titrations. Moreover, delta Hcal at > 27% saturation with ADP (corrected for ADP binding/dissociation) is approximately 80-100 kcal/mol more than in the absence of ligands. Changes in domain interactions could result from ADP bridging subunit contacts in the dodecamer. Each of the active-site ligands investigated here produces different effects on DSC profiles without uncoupling the extremely cooperative, partial unfolding reactions in the Mn.GS dodecamer.  相似文献   

2.
Active-site ligand interactions with dodecameric glutamine synthetase from Escherichia coli have been studied by calorimetry and fluorometry using the nonhydrolyzable ATP analogue 5'-adenylyl imidodiphosphate (AMP-PNP), L-glutamate, L-Met-(S)-sulfoximine, and the transition-state analogue L-Met-(S)-sulfoximine phosphate. Measurements were made with the unadenylylated enzyme at pH 7.1 in the presence of 100 mM KCl and 1.0 mM MnCl2, under which conditions the two catalytically essential metal ion sites per subunit are occupied and the stoichiometry of active-site ligand binding is equal to 1.0 equiv/subunit. Thermodynamic linkage functions indicate that there is strong synergism between the binding of AMP-PNP and L-Met-(S)-sulfoximine (delta delta G' = -6.4 kJ/mol). In contrast, there is a small antagonistic effect between the binding of AMP-PNP and L-glutamate (delta delta G' = +1.4 kJ/mol). Proton effects were negligible (less than or equal to 0.2 equiv of H+ release or uptake/mol) for the different binding reactions. The binding of AMP-PNP (or ATP) to the enzyme is entropically controlled at 303 K with delta H = +5.4 kJ/mol and delta S = +150 J/(K.mol). At 303 K, the binding of L-glutamate (delta H = -22.2 kJ/mol) or L-Met-(S)-sulfoximine [delta H = -45.6 kJ/mol with delta Cp approximately equal to -670 +/- 420 J/(K.mol)] to the AMP-PNP.Mn.enzyme complex is enthalpically controlled with opposing delta S values of -29 or -46 J/(K.mol), respectively. The overall enthalpy change is negative and the overall entropy change is positive for the simultaneous binding of AMP-PNP and L-glutamate or of AMP-PNP and L-Met-(S)-sulfoximine to the enzyme. For the binding of the transition-state analogue L-Met-(S)-sulfoximine phosphate (which inactivates the enzyme by blocking active sites), both enthalpic and entropic contributions also are favorable at 303 K [delta G' approximately equal to -109 and delta H = -54.8 kJ/mol of subunit and delta S approximately equal to +180 J/(K.mol)].  相似文献   

3.
A Ginsburg  M Zolkiewski 《Biochemistry》1991,30(39):9421-9429
Partial unfolding of dodecameric glutamine synthetase (GS) from Escherichia coli has been studied by differential scanning calorimetry (DSC). A single endotherm (tm = 51.6 +/- 0.1 degrees C and delta Hcal = 211 +/- 4 kcal/mol of enzyme) was observed in DSC experiments with Mn.GS in the presence of 1.0 mM free Mn2+ and 100 mM KCl at pH 7. The dodecameric structure of Mn.GS was retained throughout heating cycles, and thermal transitions were reversible as shown by rescans [with 6-18 mg of GS (Mr 622,000) from 15 to 68 degrees C at 20-60 degrees C/h] and by greater than 93% recovery of activity. A cooperative ratio delta Hcal/delta HvH of 1.6 +/- 0.1 and deconvolution analysis show two cooperative units (two-state transitions): t1 = 50.4 and t2 = 51.7 degrees C; the ratio of the relative sizes of thermally labile domains is approximately 1:2 as judged by delta H2/delta H1 approximately equal to 2. However, the thermally induced overall enthalpy change (0.34 cal/g) for GS dodecamer is only 5-10% of that for thermal unfolding of small globular proteins at 50 degrees C. The t1 and t2 values from deconvolutions of DSC data agree with t0.5 values previously calculated from spectral measurements of temperature-induced exposures of approximately 0.7 of 2 Trp and approximately 2 of 17 Tyr per subunit, respectively [Shrake et al. (1989) Biochemistry 28, 6281-6294], over a 14 degrees C temperature range using both stabilizing and destabilizing conditions for Mn.GS.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
Urea-induced dissociation and unfolding of manganese.glutamine synthetase (Mn.GS) have been studied at 37 degrees C (pH 7) by spectroscopic and calorimetric methods. In 0 to approximately 2 M urea, Mn.GS retains its dodecameric structure and full catalytic activity. Mn.GS is dissociated into subunits in 6 M urea, as evidenced by a 12-fold decrease in 90 degrees light scattering and a monomer molecular weight of 51,800 in sedimentation equilibrium studies. The light scattering decrease in 4 M urea parallels the time course of Trp exposure but occurs more rapidly than changes in secondary structure and Tyr exposure. Early and late kinetic steps appear to involve predominantly disruption of intra-ring and inter-ring subunit contacts, respectively, in the layered hexagonal structure of Mn.GS. The enthalpies for transferring Mn.GS into urea solutions have been measured by titration calorimetry. After correcting for the enthalpy of binding urea to the protein, the enthalpy of dissociation and unfolding of Mn.GS is 14 +/- 4 cal/g. A net proton uptake of approximately 50 H+/dodecamer accompanies unfolding reactions. The calorimetric data are consistent with urea binding to multiple, independent sites in Mn.GS and the number of binding sites increasing approximately 9-fold during the protein unfolding.  相似文献   

5.
Stopped-flow fluorimetric studies at 37 degrees C have shown that ADP, at optimal concentrations, can evoke Ca2+ or Mn2+ influx in fura-2-loaded human platelets without measurable delay. In contrast, the release of Ca2+ from intracellular stores is delayed in onset by about 200 ms. By working at a lower temperature, 17 degrees C, we have now shown that the rise in cytosolic calcium concentration ([Ca2+]i) evoked by ADP in the presence of external Ca2+ is biphasic. The use of Mn2+ as a tracer for bivalent-cation entry indicates that both phases of the ADP-evoked response are associated with influx. The fast phase of the ADP-evoked rise in [Ca2+]i, which occurs without measurable delay at both 17 degrees C and 37 degrees C, is consistent with Ca2+ entry mediated by receptor-operated channels in the plasma membrane. The delayed phase, indicated by Mn2+ quench, is coincident with the discharge of the intracellular Ca2+ stores. Forskolin did not inhibit the fast phases of ADP-evoked rise in [Ca2+]i or Mn2+ quench, but completely abolished ADP-evoked discharge of the intracellular stores, the delayed phase of the rise in [Ca2+]i observed in the presence of external Ca2+ and the second phase of Mn2+ quench. The timing of the delayed event appears to be modulated by [Ca2+]i: the delayed phase of Mn2+ quench coincides with discharge of the intracellular stores in the absence of added Ca2+, but with the second phase of the ADP-evoked rise in [Ca2+]i in the presence of extracellular Ca2+. Similarly, blockade of the early phase of Ca2+ entry by SK&F 96365 further delays the second phase. It is suggested that a pathway for Ca2+ entry which is regulated by the intracellular Ca2+ store exists in platelets. This pathway operates alongside, and appears to be modulated by the activity of other routes for Ca2+ entry into the cytosol.  相似文献   

6.
Reaction of the purified Ca2+-ATPase of sarcoplasmic reticulum at 0 degrees C at low [gamma-32P]ATP (0.1 to 0.67 microM) and enzyme (0.025 to 0.24 microM) concentration in the presence of 0.11 to 30 mM Ca2+ without added Mg2+ has resulted in the formation of phosphorylated intermediate (EP:maximal level of EP = 0.45 mol/mol of enzyme) at a very slow rate. Under these conditions, the reaction steps in which EP decomposition takes place are completely prevented. This has permitted us to study the EP formation reaction and its reversal specifically, with a considerably improved time resolution. An apparent rate constant of EP formation (Vf) increases in parallel with the concentration of Ca . ATP, but not with those of Mg . ATP, or of protonated or fully ionized free ATP. This suggests that Ca . ATP is the substrate under these conditions. If Co2+ or Mn2+ are in excess over the other ions during the reaction, Vf varies in parallel with [Co . ATP] or [Mn . ATP]. Thus, it appears that either Ca2+, Co2+, or Mn2+ can be complexed with ATP to form the effective substrate. An apparent rate constant of the back reaction of EP initiated by addition of ADP to EP (Vr) increases in proportion to [ADP] or [H . ADP], but is inhibited by increasing concentrations of the ADP complex with Ca2+ or Mg2+, indicating that free ADP or protonated ADP, or both, are actual substrates for the back reaction of EP. These results suggest a new type of site to which the metal moiety of metal . ATP complex remains bound after the release of ADP from the enzyme. An acid-stable phosphorylated intermediate (EP) produced in the presence of high Ca2+ concentrations (e.g. 0.11 mM) without added Mg2+ does not decompose spontaneously, and the major portion (approximately 90%) of this EP (EPD+) reacts with ADP to form ATP (ADP-sensitive). Upon chelating Ca2+ with ethylene glycol bis(beta-amino-ethyl ether)N,N,N',N'-tetraacetic acid (EGTA), EPD+ is converted to another form of EP (EPD-), which is unreactive with ADP (or ADP-insensitive). Addition of Mg2+, after initiation of the reaction leading to EPD- by EGTA, results in rapid production of Pi from a portion of EPD- with KMg approximately equal to 3.3 x 10(3) M-1. The fraction of EPD- that is Mg2+-sensitive (EPD-,M+) increases with reaction time at a much slower rate than the Mg2+-insensitive portion of EPD- (EPD-,M-). These results suggest that the enzyme reaction involves the sequential formation of at least three forms of acid-stable EP, viz. in the order of formation, EPD+, EPD-,M-, and EPD-,M+. The equilibrium between EPD+ and EPD-,M- is shifted by higher [K+] and [Ca2+] towards EPD+.  相似文献   

7.
Introduction of specific structural probes into substrate binding sites of Escherichia coli glutamine synthetase is now possible. Various analogues of ATP substituted with an amino or sulfhydryl moiety at the 6- or 8-position of the purine ring have been found to substitute for ATP in the autoinactivation reaction of the manganese enzyme with L-Met-(S)-sulfoximine at pH 7. Dissociation of enzyme complexes containing an ADP analogue, L-Met-(S)-sulfoximine phosphate, and 2 equiv of Mn2+ is negligible at neutral pH. Prior to binding of the mercapto nucleotides to active sites, 6-mercaptopurine ribonucleoside triphosphate (6-S-ATP) and 8-mercaptoadenosine 5'-triphosphate (8-S-ATP) also have been further modified with fluorescent and chromogenic probes for energy-transfer measurements [Maurizi, M. R., Kasprzyk, P. G., & Ginsburg, A. (1986) Biochemistry (following paper in this issue)] or with electron-dense markers for electron microscopic and X-ray crystallographic structural analyses. Binding 6-S-ATP or 8-S-ATP to enzyme active sites at pH 7.1 produced red shifts of approximately 6 nm in nucleotide spectra characteristic for transfer of these nucleotide analogues into more acidic and hydrophobic environments. The spectrum of 6-S-ADP at active sites was more red-shifted than that of 6-S-AMP attached to adenylylation sites. The thiol group at the 6- or 8-position of the purine ring of the bound nucleotides was accessible for reactions with alkylating or mercurial reagents. Alkylation or mercaptide formation produced large blue shifts in the spectrum of enzyme-bound 6-S-ADP or 8-S-ADP at active sites or of 6-S-AMP covalently bound at adenylylation sites. At least one of two tryptophanyl residues in each subunit is very near the nucleotide binding site, as evidenced by changes in tryptophanyl residue fluorescence on binding ATP, mercaptonucleotides, or other ATP analogues.  相似文献   

8.
J W Shriver  U Kamath 《Biochemistry》1990,29(10):2556-2564
The thermal unfolding of rabbit skeletal heavy meromyosin (HMM), myosin subfragment 1, and subfragment 2 has been studied by differential scanning calorimetry (DSC). Two distinct endotherms are observed in the DSC scan of heavy meromyosin. The first endotherm, with a Tm of 41 degrees C at pH 7.9 in 0.1 M KCl, is assigned to the unfolding of the subfragment 2 domain of HMM based on scans of isolated subfragment 2. The unfolding of the subfragment 2 domain is reversible both in the isolated form and in HMM. The unfolding of subfragment 2 in HMM can be fit as a single two-state transition with a delta Hvh and delta Hcal of 161 kcal/mol, indicating that subfragment 2 exists as a single domain in HMM. The unfolding of subfragment 2 is characterized by an extraordinarily large delta Cp of approximately 30,000 cal/(deg.mol). In the presence of nucleotides, the high-temperature HMM endotherm with a Tm of 48 degrees C shifts to higher temperature, indicating that this peak corresponds to the unfolding of the subfragment 1 domain. This assignment has been confirmed by comparison with isolated subfragment 1. The stabilizing effect of AMPPNP was significantly greater than that of ADP. The vanadate-trapped ADP species was slightly more stable than M.AMPPNP with a Tm at 58 degrees C. The unfolding of subfragment 1, both in the isolated form and in HMM, was irreversible. Only a single endotherm was noted in the DSC scans of the subfragment 1 domain of HMM and in freshly prepared subfragment 1 complexes.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
Glutamine synthetase (GS) catalyzes the ATP-dependent condensation of ammonia and glutamate to yield glutamine, ADP, and inorganic phosphate in the presence of divalent cations. Bacterial GS is an enzyme of 12 identical subunits, arranged in two rings of 6, with the active site between each pair of subunits in a ring. In earlier work, we have reported the locations within the funnel-shaped active site of the substrates glutamate and ATP and of the two divalent cations, but the site for ammonia (or ammonium) has remained elusive. Here we report the discovery by X-ray crystallography of a binding site on GS for monovalent cations, Tl+ and Cs+, which is probably the binding site for the substrate ammonium ion. Fourier difference maps show the following. (1) Tl+ and Cs+ bind at essentially the same site, with ligands being Glu 212, Tyr 179, Asp 50', Ser 53' of the adjacent subunit, and the substrate glutamate. From its position adjacent to the substrate glutamate and the cofactor ADP, we propose that this monovalent cation site is the substrate ammonium ion binding site. This proposal is supported by enzyme kinetics. Our kinetic measurements show that Tl+, Cs+, and NH4+ are competitive inhibitors to NH2OH in the gamma-glutamyl transfer reaction. (2) GS is a trimetallic enzyme containing two divalent cation sites (n1, n2) and one monovalent cation site per subunit. These three closely spaced ions are all at the active site: the distance between n1 and n2 is 6 A, between n1 and Tl+ is 4 A, and between n2 and Tl+ is 7 A. Glu 212 and the substrate glutamate are bridging ligands for the n1 ion and Tl+. (3) The presence of a monovalent cation in this site may enhance the structural stability of GS, because of its effect of balancing the negative charges of the substrate glutamate and its ligands and because of strengthening the "side-to-side" intersubunit interaction through the cation-protein bonding. (4) The presence of the cofactor ADP increases the Tl+ binding to GS because ADP binding induces movement of Asp 50' toward this monovalent cation site, essentially forming the site. This observation supports a two-step mechanism with ordered substrate binding: ATP first binds to GS, then Glu binds and attacks ATP to form gamma-glutamyl phosphate and ADP, which complete the ammonium binding site. The third substrate, an ammonium ion, then binds to GS, and then loses a proton to form the more active species ammonia, which attacks the gamma-glutamyl phosphate to yield Gln. (5) Because the products (Glu or Gln) of the reactions catalyzed by GS are determined by the molecule (water or ammonium) attacking the intermediate gamma-glutamyl phosphate, this negatively charged ammonium binding pocket has been designed naturally for high affinity of ammonium to GS, permitting glutamine synthesis to proceed in aqueous solution.  相似文献   

10.
A plasminogen activator with arginine ester hydrolysis activity (ABUSV-PA) has been identified and purified to homogeneity from Chinese Agkistrodon blomhoffii Ussurensis snake venom. ABUSV-PA, a monomeric protein with molecular mass of 27815.2 Da, was purified 180-fold with 0.02% recovery for protein and 3.6% recovery for esterase activity. ABUSV-PA reacts optimally with its substrate N(alpha)-tosyl-l-arginine-methyl ester (TAME) at approximately pH 7.5 and at 51 degrees C. Measurement from inductively coupled plasma-atomic emission spectroscopy (ICP-AES) reveals that ABUSV-PA is a Zn(2+)-containing protein with a stoichiometry of 1:1 [Zn(2+)]:[ABUSV-PA]. Analyses of esterase hydrolysis and UV absorption and CD spectra indicate that Zn(2+) plays an important role in maintaining the structural integrity rather than the esterase activity of ABUSV-PA. Divalent metal ions, including Ca(2+), Mg(2+), Cu(2+), Ni(2+), Mn(2+), and Co(2+), increase the TAME hydrolysis activity of ABUSV-PA. A red-shift of the emission wavelengths of the synchronous fluorescence of ABUSV-PA, compared to those of free Tyr and Trp, indicates a conformation where the Tyr and Trp residues are in exposed hydrophilic environments. The presence of zinc increases the hydrophobicity of the conformational environments surrounding the Trp residues of ABUSV-PA and affects the secondary structure of ABUSV-PA, as proved by UV absorption and CD spectroscopy.  相似文献   

11.
The relationship between Ca2+ influx (delta [Ca2+]i) and the formation of inositol 1,4,5-trisphosphate (IP3) was investigated in human platelets stimulated by various agonists. Both delta [Ca2+]i and IP3 were increased in proportion to the amount of the agonists (thrombin, ADP, PAF, STA2), the receptors of which were demonstrated in platelets, and were correlated with each other. However, the ratio of delta [Ca2+]i to IP3 was significantly varied among agonists. Furthermore, in thrombin stimulated platelets, IP3 was small at low temperature (20 degrees C) compared with that at high temperature (37 degrees C) in spite of the similar delta [Ca2+]i. Thus, Ca2+ influx in human platelets seems to be regulated directly through the receptor operated mechanism and IP3 may not be involved in it.  相似文献   

12.
The observed equilibrium constants (Kobs) of the creatine kinase (EC 2.7.3.2), myokinase (EC 2.7.4.3), glucose-6-phosphatase (EC 3.1.3.9), and fructose-1,6-diphosphatase (EC 3.1.3.11) reactions have been determined at 38 degrees C, pH 7.0, ionic strength 0.25, and varying free magnesium concentrations. The equilibrium constant (KCK) for the creatine kinase reaction defined as: KCK = [sigma ATP] [sigma creatine] divided by ([sigma ADP] [sigma creatine-P] [H+]) was measured at 0.25 ionic strength and 38 degrees C and was shown to vary with free [Mg2+]. The value was found to be 3.78 x 10(8) M-1 at free [Mg2+] = 0 and 1.66 x 10(9) M-1 at free [Mg2+] = 10(-3) M. Therefore, at pH 7.0, the value of Kobs, defined as Kobs = KCK[H+] = [sigma ATP] [sigma creatine] divided by ([sigma ADP] [sigma creatine-P] was 37.8 at free [Mg2+] = 0 and 166 at free [Mg2+] = 10(-3) M. The Kobs value for the myokinase reaction, 2 sigma ADP equilibrium sigma AMP + sigma ATP, was found to vary with free [Mg2+], being 0.391 at free [Mg2+] = 0 and 1.05 at free [Mg2+] = 10(-3) M. Taking the standard state of water to have activity equal to 1, the Kobs of glucose-6-P hydrolysis, sigma glucose-6-P + H2O equilibrium sigma glucose + sigma Pi, was found not to vary with free [Mg2+], being 110 M at both free [Mg2+] = 0 and free [Mg2+] = 10(-3) M. The Kobs of fructose-1,6-P2 hydrolysis, sigma fructose-1,6-P2 equilibrium sigma fructose-6-P + sigma Pi, was found to vary with free [Mg2+], being 272 M at free [Mg2+] = 0 and 174 M at free [Mg2+] = 0.89 x 10(-3) M.  相似文献   

13.
Direct measurements of phosphorylation of the Ca2+ ATPase of the sarcoplasmic reticulum (SR) have shown that the lifetime of the first phosphorylated intermediate in the Ca2+ transport cycle, E1 approximately P, increases with decreasing [Mg2+] (Dupont, Y. 1980. Eur. J. Biochem. 109:231-238). Previous x-ray diffraction work (Pascolini, D., and J.K. Blasie. 1988. Biophys. J. 54:669-678) under high [Mg2+] conditions (25 mM) indicated that changes in the profile structure of the SR membrane could be responsible for the low-temperature transient trapping of E1 approximately P that occurs at temperatures below 2-3 degrees C, the upper characteristic temperature th for lipid lateral phase separation in the membrane. We now present results of our study of the Ca2+ uptake kinetics and of the structure of the SR membrane at low [Mg2+] (less than or equal to 100 microM). Our results show a slowing in the kinetics of both phases of the Ca2+ uptake process and an increase in the duration of the plateau of the fast phase before the onset of the slow phase, indicating an increase in the lifetime (transient trapping) of E1 approximately P. Calcium uptake kinetics at low [Mg2+] and moderately low temperature (approximately 0 degree C) are similar to those observed at much lower temperatures (approximately -10 degrees C) at high [Mg2+]. The temperature-induced structural changes that we observed at low [Mg2+] are much more pronounced than those found to occur at higher [Mg2+]. Also, at the lower [Mg2+] the upper characteristic temperature th for lipid lateral phase separation was found to be higher, at approximately 8-10 degrees C. Our studies indicate that both temperature and [Mg2+] affect the structure and the functionality (as measured by changes in the kinetics of Ca2+ uptake) of the SR membrane. Membrane lipid phase behavior and changes in the Ca2+ ATPase profile structure seem to be related, and we have found that structural changes are responsible for the slowing of the kinetics of the fast phase of Ca2+ uptake, and could also mediate the effect that [Mg2+] has on E1 approximately P lifetime.  相似文献   

14.
Phosphorylation of red cell membranes at ambient temperatures with micromolar [32P]ATP in the presence of Na ions produced phosphoenzyme that was dephosphorylated rapidly upon the addition of ADP or K ions. However, as first observed by Blostein (1968, J. Biol. Chem., 243:1957), the phosphoenzyme formed at 0 degrees C under otherwise identical conditions was insensitive to the addition of K ions but was dephosphorylated rapidly by ADP. This suggested that the conformational transition from ADP-sensitive, K-insensitive Na pump phosphoenzyme (E1 approximately P) to K-sensitive, ADP-insensitive phosphoenzyme (E2P) is blocked at 0 degrees C. Since the ATP:ADP exchange reaction is a partial reaction of the overall enzyme cycle dependent upon the steady state level of E1 approximately P that is regulated by [Na], we examined the effects of temperature on the curve relating [Na] to ouabain-sensitive ATP:ADP exchange. The characteristic triphasic curve seen at higher temperatures when [Na] was between 0.5 and 100 mM was not obtained at 0 degrees C. Simple saturation was observed instead with a K0.5 for Na of approximately 1 mM. The effect of increasing temperature on the ATP:ADP exchange at fixed (150 mM) Na was compared with the effect of increasing temperature on (Na + K)-ATPase activity of the same membrane preparation. It was observed that (a) at 0 degrees C, there was significant ouabain-sensitive ATP:ADP exchange activity, (b) at 0 degrees C, ouabain-sensitive (Na + K)-ATPase activity was virtually absent, and (c) in the temperature range 5-37 degrees C, there was an approximately 300-fold increase in (Na + K)-ATPase activity with only a 9-fold increase in the ATP:ADP exchange. These observations are in keeping with the suggestion that the E1 approximately P----E2P transition of the Na pump in human red cell membranes is blocked at 0 degrees C. Previous work has shown that the inhibitory effect of Na ions and the low-affinity stimulation by Na of the rate of ATP:ADP exchange occur at the extracellular surface of the Na pump. The absence of both of these effects at 0 degrees C, where E1 approximately P is maximal, supports the idea that external Na acts through sites on the E2P form of the phosphoenzyme.  相似文献   

15.
Probes for fluorescence energy transfer measurements were introduced into active sites of dodecameric glutamine synthetase from Escherichia coli by substituting appropriate ATP analogues for ATP in the autoinactivation reaction of this enzyme with L-Met-(S)-sulfoximine and Mn2+ [Maurizi, M. R., & Ginsburg, A. (1986) Biochemistry (preceding paper in this issue)]. Two fluorescent donors, 8-mercapto-ATP alkylated with either 5-[[[(iodoacetyl)amino]ethyl]amino]naphthalene-1-sulfonic acid (AEDANS-ATP) or 1,N6-etheno-2-aza-ATP (aza-epsilon-ATP), and two acceptors, 6-mercaptopurine ribonucleotide triphosphate or 8-mercapto-ATP alkylated with the chromophore 4'-[[4-(dimethylamino)-phenyl]azo]-2-iodoacetanilide (6-Y- or 8-Y-ATP), were used. Fluorescence emissions of enzyme derivatives with 1 or 2 equiv of fluorescent donor per dodecamer and either an acceptor (Y) or ADP at the remaining active sites were compared at pH 7.0. The results, together with the known geometry of the enzyme, indicate that active site probes in the dodecamer are widely separated and that energy transfer occurs from a single donor to two or three acceptors on adjacent subunits. The calculated distance between equidistant active site probes on heterologously bonded subunits within the same hexagonal ring is 56-60 A. Probes on isologously bonded subunits are no closer than 60 A and may be as far apart as 78 A. Thus, active sites are away from the 6-fold axis of symmetry toward the outer edges of the dodecamer and are located greater than or equal to 30 A from the plane separating the hexagonal rings. During Zn2+-induced stacking of the same enzyme derivatives along the 6-fold axes of symmetry, additional quenches of fluorescent probes were dependent on the presence of acceptors on separate dodecamers. The Zn2+-induced face to face aggregation of dodecamers in the presence of 46 microM ZnCl2 and 9 mM MgCl2 at pH 7.0 had an Arrhenius activation energy of 22.3 +/- 0.2 kcal/mol and a second-order rate constant at 25 degrees C of approximately 10(5) M-1 s-1 at early stages. Time-dependent fluorescence quenches correlated well with the degree of linear polymer formation and reached maximum values of 47-70% quench when the average n-mer was six dodecamers. After correction for unquenched polymer ends, a fluorescent donor and an acceptor probe in layered dodecamers were estimated to be approximately 36 A apart--an average value if there is some twisting of single strands.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

16.
We have studied the uptake of Ca2+ and its redistribution between the cytoplasm and the intracellular stores in Ehrlich-ascites-tumour cells and rat thymocytes previously depleted of Ca2+ by incubation in Ca2(+)-free medium. Measurements included changes of the cytoplasmic Ca2+ concentration ([Ca2+]i), uptake of 45Ca2+ and uptake of Mn2+, a Ca2+ surrogate for Ca2+ channels. Refilling of the Ca2+ stores in thymocytes was very fast (half-filling time: 4 s at 37 degrees C) and very sensitive to temperature (10 times slower at 20 degrees C). It was always preceded by increase of [Ca2+]i. In the Ehrlich cell, both refilling and increase of [Ca2+]i were about one order of magnitude slower. The increase of [Ca2+]i and the refilling of the intracellular stores were both almost completely blocked by Ni2+ in thymocytes, but only partially in the Ehrlich cell. The rates of 45Ca2+ and Mn2+ uptake varied consistently with temperature and the kind of cell. These results suggest that the intracellular stores are refilled by Ca2+ taken up from the cytoplasm. We also find that filling of the Ca2+ stores decreases by about 90% the rate of Mn2+ uptake in thymocytes. This is direct evidence of modulation of the plasma-membrane Ca2+ entry by the degree of filling of the intracellular stores. This modulation occurs in the absence of agonists, suggesting some kind of signalling between the intracellular stores and the Ca2+ entry pathways of the plasma membrane.  相似文献   

17.
The side chain of Gln143, a conserved residue in manganese superoxide dismutase (MnSOD), forms a hydrogen bond with the manganese-bound solvent and is critical in maintaining catalytic activity. The side chains of Tyr34 and Trp123 form hydrogen bonds with the carboxamide of Gln143. We have replaced Tyr34 and Trp123 with Phe in single and double mutants of human MnSOD and measured their catalytic activity by stopped-flow spectrophotometry and pulse radiolysis. The replacements of these side chains inhibited steps in the catalysis as much as 50-fold; in addition, they altered the gating between catalysis and formation of a peroxide complex to yield a more product-inhibited enzyme. The replacement of both Tyr34 and Trp123 in a double mutant showed that these two residues interact cooperatively in maintaining catalytic activity. The crystal structure of Y34F/W123F human MnSOD at 1.95 A resolution suggests that this effect is not related to a conformational change in the side chain of Gln143, which does not change orientation in Y34F/W123F, but rather to more subtle electronic effects due to the loss of hydrogen bonding to the carboxamide side chain of Gln143. Wild-type MnSOD containing Trp123 and Tyr34 has approximately the same thermal stability compared with mutants containing Phe at these positions, suggesting the hydrogen bonds formed by these residues have functional rather than structural roles.  相似文献   

18.
Bundles of 10-100 fibers were dissected from the extensor digitorum longus muscle of mouse, mounted in an apparatus for optical recording, and stretched to long sarcomere length (> or = 3.6 microns). One fiber within the bundle was microinjected with furaptra, a fluorescent indicator that responds rapidly to changes in myoplasmic free [Ca2+] (delta [Ca2+]). Twitches and brief tetani were initiated by external stimulation. At myoplasmic furaptra concentrations of approximately 0.1 mM, the indicator's fluorescence signal during fiber activity (delta F/F) was well resolved. delta F/F was converted to delta [Ca2+] under the assumption that furaptra's myoplasmic dissociation constant for Ca2+ is 98 microM at 16 degrees C and 109 microM at 28 degrees C. At 16 degrees C, the peak amplitude of delta [Ca2+] during a twitch was 17.8 +/- 0.4 microM (+/-SEM; n = 8) and the half-width of delta [Ca2+] was 4.6 +/- 0.3 ms. At 28 degrees C, the peak and half-width values were 22.1 +/- 1.8 microM and 2.0 +/- 0.1 ms, respectively (n = 4). During a brief high-frequency tetanus, individual peaks of delta [Ca2+] were also well resolved and reached approximately the same amplitude that resulted from a single shock; the initial decays of delta [Ca2+] from peak slowed substantially during the tetanus. For a single twitch at 16 degrees C, the amplitude of delta [Ca2+] in fast-twitch fibers of mouse is not significantly different from that recently measured in fast- twitch fibers of frog (16.5 +/- 0.9 microM; Zhao, M., S. Hollingworth, and S.M. Baylor. 1996. Biophys. J. 70:896-916); in contrast, the half- width of delta [Ca2+] is surprisingly brief in mouse fibers, only about half that measured in frog (9.6 +/- 0.6 ms). The estimated peak rate at which Ca2+ is released from the sarcoplasmic reticulum in response to an action potential is also similar in mouse and frog, 140-150 microM/ms (16 degrees C).  相似文献   

19.
This study examines the activation of divalent cation entry into rat parotid gland acinar cells by using Mn2+ as a Ca2+ surrogate cation. Following muscarinic-cholinergic stimulation of dispersed parotid acini with carbachol (10 microM), the onset of internal Ca2+ release (cytosolic [Ca2+], [Ca2+]i, increase) and the stimulation of Mn2+ entry (increase in fura2 quenching) are not simultaneously detected. [Ca2+]i elevation, due to intracellular release, is detected almost immediately following carbachol addition and peak [Ca2+]i increase occurs at 6.0 +/- 0.8 sec. However, there is an interval (apparent lag) between carbachol addition and the detection of stimulated Mn2+ entry. This apparent lag is decreased from 26 +/- 3.1 sec to 9.2 +/- 1.5 sec when external Mn2+ ([Mn2+]0) is increased from 12.5 to 500 microM. It is not decreased further with increase in [Mn2+]0 from 500 microM to 1 mM (9.8 +/- 2.1 sec), although both intracellular free Mn2+ and [Mn2+-fura2]/[fura2] increase. Thus, at [Mn2+]0 < 500 microM, the observed lag time is partially due to a limitation in the magnitude of Mn2+ entry. Furthermore, neither peak [Ca2+]i nor the time required to reach peak [Ca2+]i is significantly altered by [Mn2+]0 (12.5 microM to 1 mM). At every [Mn2+]0 tested (i.e., 12.5 microM-1 mM), the apparent lag is significantly greater than the time required to reach peak [Ca2+]i. However, when carbachol stimulation of the [Ca2+]i increase is attenuated by loading the acini with the Ca2+ chelator, 2-bis(O-aminophenoxy)ethane-N,N,N',N'-tetraacetate (BAPTA), there is no detectable lag in carbachol stimulation of Mn2+ entry (with 1 mM [Mn2+]0). Importantly, in BAPTA-loaded acini, carbachol stimulates Mn2+ entry via depletion of the internal Ca2+ pool and not via direct activation of other divalent cation entry mechanisms. Based on these results, we suggest that the apparent lag in the detection of carbachol stimulation of Mn2+ entry into parotid acinar cells is due to a retardation of Mn2+ entry by the initial increase in [Ca2+]i, due to internal release, which most likely occurs proximate to the site of divalent cation entry.  相似文献   

20.
Cytoplasmic free Ca2+ concentration, [Ca2+]i, was estimated in single rabbit blood platelets by digital imaging microscopy with the use of the specific Ca(2+)-indicator dye Fura-2. Uneven distribution and low level of [Ca2+]i was found in the resting platelet even in the presence of extracellular 1 mM Ca2+. Thrombin at 1 unit/ml immediately caused a transient increase in [Ca2+]i, which was followed by a secondary and sustained increase in [Ca2+]i. The distribution of increased levels of [Ca2+]i was also shown to be uneven within the cell. The presence of 1 mM EGTA in the medium only slightly decreased the initial rise in [Ca2+]i, but completely inhibited the latter phase, a sustained rise in [Ca2+]i. This result shows that the initial rise of [Ca2+]i might not be caused by Ca2+ influx, but might be induced by mobilization of Ca2+ from intracellular Ca2+ storage sites. This speculation is further supported by the fact that the elevated [Ca2+]i induced by thrombin immediately decreased to the base line value when 3 mM EGTA was applied. Thus, thrombin induced elevation of [Ca2+]i is suggested to consist of two different processes, namely the mobilization of Ca2+ from the intracellular storage sites and the successive Ca2+ influx through the receptor activated Ca2+ channels. Stimulation with ADP also caused a rapid elevation of platelet [Ca2+]i, but this effect of ADP was different form that of thrombin. Thus, the ADP induced rise in [Ca2+]i was accompanied by oscillation and was inhibited by extracellular EGTA. Our present experiment is the first report that clearly and directly reveals the differences between the effects of thrombin and ADP on [Ca2+]i of platelets.  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