首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Patrick J. O’Malley 《BBA》2002,1553(3):212-217
Hybrid density functional calculations are used to model tyrosine oxidation during electron transfer reactions of photosystem II. The predicted frequency values for the 7a and δCOH modes of the reduced form and the 7a mode of the oxidised radical form are in excellent agreement with experimental data obtained for Mn and Ca depleted systems by Hienerwadel et al. [Biochemistry 36 (1997) 15447] and Berthomieu et al. [Biochemistry 37 (1998) 10547]. The calculations confirm that the two tyrosines YD and YZ are protonated in the reduced form. On oxidation the larger 7a frequency value observed experimentally for YZ can be best explained by a greater localisation of the protonic charge released on formation of this tyrosyl free radical.  相似文献   

2.
Sjöholm J  Styring S  Havelius KG  Ho FM 《Biochemistry》2012,51(10):2054-2064
Cryogenic illumination of Photosystem II (PSII) can lead to the trapping of the metastable radical Y(Z)(?), the radical form of the redox-active tyrosine residue D1-Tyr161 (known as Y(Z)). Magnetic interaction between this radical and the CaMn(4) cluster of PSII gives rise to so-called split electron paramagnetic resonance (EPR) signals with characteristics that are dependent on the S state. We report here the observation and characterization of a split EPR signal that can be directly induced from PSII centers in the S(2) state through visible light illumination at 10 K. We further show that the induction of this split signal takes place via a Mn-centered mechanism, in the same way as when using near-infrared light illumination [Koulougliotis, D., et al. (2003) Biochemistry 42, 3045-3053]. On the basis of interpretations of these results, and in combination with literature data for other split signals induced under a variety of conditions (temperature and light quality), we propose a unified model for the mechanisms of split signal induction across the four S states (S(0), S(1), S(2), and S(3)). At the heart of this model is the stability or instability of the Y(Z)(?)(D1-His190)(+) pair that would be formed during cryogenic oxidation of Y(Z). Furthermore, the model is closely related to the sequence of transfers of protons and electrons from the CaMn(4) cluster during the S cycle and further demonstrates the utility of the split signals in probing the immediate environment of the oxygen-evolving center in PSII.  相似文献   

3.
Incubation of Mn-depleted PSII membranes [PSII(-Mn)] with Fe(II) is accompanied by the blocking of Y(Z)(*) at the high-affinity Mn-binding site to exogenous electron donors [Semin et al. (2002) Biochemistry 41, 5854-5864] and a shift of the pK(app) of the hydrogen bond partner for Y(Z) (base B) from 7.1 to 6.1 [Semin, B. K., and Seibert, M. (2004) Biochemistry 43, 6772-6782]. Here we calculate activation energies (E(a)) for Y(Z)(*) reduction in PSII(-Mn) and Fe-blocked PSII(-Mn) samples [PSII(-Mn, +Fe)] from temperature dependencies of the rate constants of the fast and slow components of the flash-probe fluorescence decay kinetics. At pH < pK(app) (e.g., 5.5), the decays are fit with one (fast) component in both types of samples, and E(a) is equal to 42.2 +/- 2.9 kJ/mol in PSII(-Mn) and 46.4 +/- 3.3 kJ/mol in PSII(-Mn, +Fe) membranes. At pH > pK(app), the decay kinetics exhibit an additional slow component in PSII(-Mn, +Fe) membranes (E(a) = 36.1 +/- 7.5 kJ/mol), which is much lower than the E(a) of the corresponding component observed for Y(Z)(*) reduction in PSII(-Mn) samples (48.1 +/- 1.7 kJ/mol). We suggest that the above difference results from the formation of a strong low barrier hydrogen bond (LBHB) between Y(Z) and base B in PSII(-Mn, +Fe) samples. To confirm this, Fe-blocking was performed in D(2)O to insert D(+), which has an energetic barrier distinct from H(+), into the LBHB. Measurement of the pH effects on the rates of Y(Z)(*) reduction in PSII(-Mn, +Fe) samples blocked in D(2)O shows a shift of the pK(app) from 6.1 to 7.6, and an increase in the E(a) of the slow component. This approach was also used to measure the stability of the Y(Z)(*) EPR signal at various temperatures in both kinds of membranes. In PSII(-Mn) membranes, the freeze-trapped Y(Z)(*) radical is stable below 190 K, but half of the Y(Z)(*) EPR signal disappears after a 1-min incubation when the sample is warmed to 253 K. In PSII(-Mn, +Fe) samples, the trapped Y(Z)(*) radical is unstable at a much lower temperature (77 K). However, the insertion of D(+) into the hydrogen bond between Y(Z) and base B during the blocking process increases the temperature stability of the Y(Z)(*) EPR signal at 77 K. Again, these results indicate that Fe-blocking involves Y(Z) in the formation of a LBHB, which in turn is consistent with the suggested existence of a LBHB between Y(Z) and base B in intact PSII membranes [Zhang, C., and Styring, S. (2003) Biochemistry 42, 8066-8076].  相似文献   

4.
The Y(Z)-tyrosine radical was trapped by freezing immediately after illumination in Ca(2+)-depleted Photosystem II (PS II) membranes and the pH-dependent characteristics of the radical were investigated using CW-EPR and pulsed ENDOR. The spectrum of the Y*(Z) radical trapped in the Y*(Z)S(1) state at pH 5.5 was cation-like as reported in Mn-depleted PS II (H. Mino et al., Spectrochim. Acta A 53 (1997) 1465-1483). By illuminating the PS II-retaining S(2) state, the Y*(Z) radical and a broad doublet signal formed in the g approximately 2 region were trapped concomitantly. The spectrum of the trapped Y*(Z) radical in the Y*(Z)S(2) state was cation-like at pH 5.5 but the pulsed ENDOR measurements reveals the involvement of the neutral Y*(Z) radical in the doublet signal. At pH 7.0, the resulting Y*(Z) signal was the mixture of the cation-like and neutral radical spectra, and considerably different from the neutral radical found in Mn-depleted PS II. pH-Dependent changes in the properties of the Y*(Z) radical are discussed in relation to the redox events occurring in Ca(2+)-depleted PS II.  相似文献   

5.
We have previously shown that introduction of an engineered Met160 residue in ascorbate peroxidase (S160M variant) leads to the formation of a covalent link between Met160 and the heme vinyl group [Metcalfe, C. L., et al. (2004) J. Am. Chem. Soc. 126, 16242-16248]. In this work, we have used electronic spectroscopy, HPLC, and mass spectrometry to show that the introduction of a tyrosine residue at the same position (S160Y variant) leads, similarly, to the formation of a heme-tyrosine covalent link in an autocatalytic reaction that also leads to formation of a second covalent link from the heme to Trp41 [Pipirou, Z., et al. (2007) Biochemistry 46, 2174-2180]. Stopped-flow and EPR data implicate the involvement of a tyrosyl radical in the reaction mechanism. The results indicate that the heme can support the formation of different types of covalent links under appropriate conditions. The generality of this idea is discussed in the context of other heme enzymes.  相似文献   

6.
The crystal structure of the cyano-met form of Mt-trHbO revealed two unusual distal residues Y(CD1) and W(G8) forming a hydrogen-bond network with the heme-bound ligand [Milani, M., et al. (2003) Proc. Natl. Acad. Sci. U.S.A. 100, 5766-5771]. W(G8) is an invariant residue in group II and group III trHbs and has no counterpart in other globins. A previous study reported that changing Y(CD1) for a Phe causes a significant increase in the O2 combination rate, but almost no change in the O2 dissociation rate [Ouellet, H., et al. (2003) Biochemistry 42, 5764-5774]. Here we investigated the role of the W(G8) in ligand binding by using resonance Raman spectroscopy, stopped-flow spectrophotometry, and X-ray crystallography. For this purpose, W(G8) was changed, by site-directed mutagenesis, to a Phe in both the wild-type protein and the mutant Y(CD1)F to create the single mutant W(G8)F and the double mutant Y(CD1)F/W(G8)F, respectively. Resonance Raman results suggest that W(G8) interacts with the heme-bound O2 and CO, as evidenced by the increase of the Fe-O2 stretching mode from 559 to 564 cm-1 and by the lower frequency of the Fe-CO stretching modes (514 and 497 cm-1) compared to that of the wild-type protein. Mutation of W(G8) to Phe indicates that this residue controls ligand binding, as evidenced by a dramatic increase of the combination rates of both O2 and CO. Also, the rate of O2 dissociation showed a 90-1000-fold increase in the W(G8)F and Y(CD1)F/W(G8)F mutants, that is in sharp contrast with the values obtained for the other distal mutants Y(B10)F and Y(CD1)F [Ouellet, H., et al. (2003) Biochemistry 42, 5764-5774]. Taken together, these data indicate a pivotal role for the W(G8) residue in O2 binding and stabilization.  相似文献   

7.
M Baumgarten  J S Philo  G C Dismukes 《Biochemistry》1990,29(48):10814-10822
New evidence on the chloride requirement for photosynthetic O2 evolution has indicated that Cl- facilitates oxidation of the manganese cluster by the photosystem II (PSII) Tyr-Z+ radical. Illumination above 250 K of spinach PSII centers which are inhibited in O2 evolution by either Cl- depletion or F- substitution produces a new EPR signal which has magnetic characteristics similar to one recently discovered in samples inhibited by depletion of Ca2+ only [Boussac et al. (1989) Biochemistry 28, 8984; Sivaraja et al. (1989) Biochemistry 28, 9459]. The physiological roles of Cl- and Ca2+ in water oxidation are thus linked. The characteristics include a nearly isotropic g = 2.00 +/- 0.005, a symmetric line shape with line width = 16 +/- 2 mT, almost stoichiometric spin concentration relative to Try-D+ = 0.6 +/- 0.3 spin/PSII, very rapid spin relaxation at all temperatures measured down to 6 K, and an undetectable change in magnetic susceptibility upon formation (less than 1 mu B2). The signal appears to originate from a spin doublet (radical) in magnetic dipolar contact with a transition-metal ion, most probably a photooxidized protein residue within 10 A of the Mn cluster (Mn-proximal radical). It is distinct from the three other protein-bound radical-type electron donors found in the PSII reaction center: Tyr-D+, Tyr-Z+, and C+. This signal photoaccumulates to a stable level under continuous illumination at 270 K and decays only after illumination stops.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
We have probed the electrostatics of P680(+) reduction in oxygenic photosynthesis using histidine-tagged and histidine-tagged Y(D)-less Photosystem II cores. We make two main observations: (i) that His-tagged Chlamydomonas cores show kinetics which are essentially identical to those of Photosystem II enriched thylakoid membranes from spinach; (ii) that the microsecond kinetics, previously shown to be proton/hydrogen transfer limited [Schilstra et al. (1998) Biochemistry 37, 3974-3981], are significantly different in Y(D)-less Chlamydomonas particles when compared with both the His-tagged Chlamydomonas particles and the spinach membranes. The oscillatory nature of the kinetics in both Chlamydomonas samples is normal, indicating that S-state cycling is unaffected by either the histidine-tagging or the replacement of tyrosine D with phenylalanine. We propose that the effects on the proton-coupled electron transfers of P680(+) reduction in the absence of Y(D) are likely to be due to pK shifts of residues in a hydrogen-bonded network of amino acids in the vicinity of Y(Z). Tyrosine D is 35 A from Y(Z) and yet has a significant influence on proton-coupled electron transfer events in the vicinity of Y(Z). This finding emphasizes the delicacy of the proton balance that Photosystem II has to achieve during the water splitting process.  相似文献   

9.
G H Noren  B A Barry 《Biochemistry》1992,31(13):3335-3342
The currently accepted model for the location of the redox-active tyrosines, D and Z, in photosystem II suggests that they are symmetrically located on the D1 and D2 polypeptides, which are believed to form the heterodimer core of the reaction center. Z, the electron conduit from the manganese catalytic site to the primary chlorophyll donor, has been identified with tyrosine-161 of D1. The YF161D1 mutant of Synechocystis 6803 [Debus, R. J., Barry, B. A., Sithole, I., Babcock, G. T., & McIntosh, L. (1988b) Biochemistry 27, 9071-9074; Metz, J. G., Nixon, P. J., Rogner, M., Brudvig, G. W., & Diner, B. A. (1989) Biochemistry 28, 6960-6969], in which this tyrosine has been changed to a phenylalanine, should have no light-induced EPR (electron paramagnetic resonance) signal from a tyrosine radical. This negative result has indeed been obtained in analysis of one of two independently constructed mutants through the use of a non-oxygen-evolving core preparation (Metz et al., 1989). Here, we present an analysis of a YF161D1 mutant through the use of a photosystem II purification procedure that gives oxygen-evolving particles from wild-type Synechocystis cultures. In our mutant preparation, a light-induced EPR signal from a photosystem II radical is observed under conditions in which, in a wild-type preparation, we can accumulate an EPR signal from Z+. This EPR signal has a different lineshape from that of the Z+ tyrosine radical, and spin quantitation shows that this radical can be produced in up to 60% of the mutant reaction centers. The EPR lineshape of this radical suggests that photosystem II reaction centers of the YF161D1 mutant contain a redox-active amino acid.  相似文献   

10.
The Helical structures of d(C-G-C-A-m5C-G-T-G-m5C-G), d(m5C-G-C-A-m5C-G-T-G-C-G) and d(C-2aminoA-C-G-T-G) were studied in aqueous solution at various salt concentrations and temperatures by 1H-NMR spectroscopy. In 0.1 M NaCl solution only the B form was evidenced for these DNA fragments whereas in 4 M NaCl both B and Z forms, in slow exchange on the NMR time scale, were observed. Under these conditions the Z form accounted for less than 60% of the decamer conformation; conversely d(C-G)3 hexamers containing methylated cytidines were predominantly in the Z form (greater than 90%) [Tran-Dinh et al. (1984) Biochemistry 23, 1362; Cavaillès et al. (1984) J. Biomol. Struct. Dyn. 1, 1347-1371]. On the other hand, d(C-2aminoA-C-G-T-G) in which the d(2aminoA) X dT base pair forms three hydrogen bonds, was found to adopt the Z conformation in 4M NaCl solution which was not the case for d(C-A-C-G-T-G) (unpublished results). The present study shows that the B in equilibrium Z transition in solution is highly sequence-dependent and that correlation exists between the stability of the duplexes (essentially governed by the number of hydrogen bonds between complementary bases) and their ability to adopt the Z conformation.  相似文献   

11.
A role for redox-active tyrosines has been demonstrated in many important biological processes, including water oxidation carried out by photosystem II (PSII) of oxygenic photosynthesis. The rates of tyrosine oxidation and reduction and the Tyr/Tyr reduction potential are undoubtedly controlled by the immediate environment of the tyrosine, with the coupling of electron and proton transfer, a critical component of the kinetic and redox behavior. It has been demonstrated by Faller et al. that the rate of oxidation of tyrosine D (Tyr(D)) at room temperature and the extent of Tyr(D) oxidation at cryogenic temperatures, following flash excitation, dramatically increase as a function of pH with a pK(a) of approximately 7.6 [Faller et al. 2001 Proc. Natl. Acad. Sci. USA 98, 14368-14373; Faller et al. 2001 Biochemistry 41, 12914-12920]. In this work, we investigated, using FTIR difference spectroscopy, the mechanistic reasons behind this large pH dependence. These studies were carried out on Mn-depleted PSII core complexes isolated from Synechocystis sp. PCC 6803, WT unlabeled and labeled with (13)C(6)-, or (13)C(1)(4)-labeled tyrosine, as well as on the D2-Gln164Glu mutant. The main conclusions of this work are that the pH-induced changes involve the reduced Tyr(D) state and not the oxidized Tyr(D)() state and that Tyr(D) does not exist in the tyrosinate form between pH 6 and 10. We can also exclude a change in the protonation state of D2-His189 as being responsible for the large pH dependence of Tyr(D) oxidation. Indeed, our data are consistent with D2-His189 being neutral both in the Tyr(D) and Tyr(D)() states in the whole pH6-10 range. We show that the interactions between reduced Tyr(D) and D2-His189 are modulated by the pH. At pH greater than 7.5, the nu(CO) mode frequency of Tyr(D) indicates that Tyr(D) is involved in a strong hydrogen bond, as a hydrogen bond donor only, in a fraction of the PSII centers. At pH below 7.5, the hydrogen-bonding interaction formed by Tyr(D) is weaker and Tyr(D) could be also involved as a hydrogen bond acceptor, according to calculations performed by Takahashi and Noguchi [J. Phys. Chem. B 2007 111, 13833-13844]. The involvement of Tyr(D) in this strong hydrogen-bonding interaction correlates with the ability to oxidize Tyr(D) at cryogenic temperatures and rapidly at room temperature. A strong hydrogen-bonding interaction is also observed at pH 6 in the D2-Gln164Glu mutant, showing that the residue at position D2-164 regulates the properties of Tyr(D.) The IR data point to the role of a protonatable group(s) (with a pK(a) of approximately 7) other than D2-His189 and Tyr(D), in modifying the characteristics of the Tyr(D) hydrogen-bonding interactions, and hence its oxidation properties. It remains to be determined whether the strong hydrogen-bonding interaction involves D2-His189 and if Tyr(D) oxidation involves the same proton transfer route at low and at high pH.  相似文献   

12.
Peptidylglycine alpha-hydroxylating monooxygenase (PHM) and dopamine beta-monooxygenase (DbetaM) are homologous copper-containing enzymes that catalyze an oxygen-dependent hydroxylation of peptide-extended glycine residues and phenethylamines, respectively. The mechanism whereby these enzymes activate molecular oxygen and the C-H bond of substrate has been the subject of numerous studies, and various mechanisms have been put forth. From the magnitude of (18)O isotope effects as a function of substrate structure in DbetaM, an active site tyrosine had been proposed to function in the reductive activation of Cu(II)-OOH to generate a reactive copper-oxo species [Tian et al. (1994) Biochemistry 33, 226]. The presence of a tyrosine residue, Y318, in the active site of PHM was subsequently confirmed from crystallographic studies [Prigge et al. (1997) Science 278, 1300]. We now report extensive kinetic and isotope effect studies on the Y318F mutant form of PHM, analyzing the role of this tyrosine in the catalytic mechanism. It is found that the Y318F mutant has intrinsic hydrogen and (18)O isotope effects that are within experimental error of the wild-type enzyme and that the mutation causes only a slight reduction in the rate constant for C-H bond cleavage. These findings, together with the recent demonstration that C-H activation in PHM is dominated by quantum mechanical tunneling [Francisco et al. (2002) J. Am. Chem. Soc. 124, 8194], necessitate a reexamination of plausible mechanisms for this unique class of copper enzymes.  相似文献   

13.
Near-IR (NIR) excitation at liquid He temperatures of photosystem II (PSII) membranes from the cyanobacterium Synechococcus vulcanus or from spinach poised in the S2 state results in the production of a g = 2.035 EPR resonance, reminiscent of metalloradical signals. The signal is smaller in the spinach preparations, but it is significantly enhanced by the addition of exogenous quinones. Ethanol (2-3%, v/v) eliminates the ability to trap the signal. The g = 2.035 signal is identical to the one recently obtained by Nugent et al. by visible-light illumination of the S1 state, and preferably assigned to S1Y(Z*) [Nugent, J. H. A., Muhiuddin, I. P., and Evans, M. C. W. (2002) Biochemistry 41, 4117-4126]. The production of the g = 2.035 signal by liquid He temperature NIR excitation of the S2 state is paralleled by a significant reduction (typically 40-45% in S. vulcanus) of the S2 state multiline signal. This is in part due to the conversion of the Mn cluster to higher spin states, an effect documented by Boussac et al. [Boussac, A., Un, S., Horner, O., and Rutherford, A. W. (1998) Biochemistry 37, 4001-4007], and in part due to the conversion to the g = 2.035 configuration. Following the decay of the g = 2.035 signal at liquid helium temperatures (decay halftimes in the time range of a few to tens of minutes depending on the preparation), annealing at elevated temperatures (-80 degrees C) results in only partial restoration of the S2 state multiline signal. The full size of the signal can be restored by visible-light illumination at -80 degrees C, implying that during the near-IR excitation and subsequent storage at liquid helium temperatures recombination with Q(A-) (and therefore decay of the S2 state to the S1 state) occurred in a fraction of centers. In support of this conclusion, the g = 2.035 signal remains stable for several hours (at 11 K) in centers poised in the S2...Q(A) configuration before the NIR excitation. The extended stability of the signal under these conditions has allowed the measurement of the microwave power saturation and the temperature dependence in the temperature range of 3.8-11 K. The signal intensity follows Curie law temperature dependence, which suggests that it arises from a ground spin state, or a very low-lying excited spin state. The P1/2 (microwave power at half-saturation) value is 1.7 mW at 3.8 K and increases to 96 mW at 11 K. The large width of the g = 2.035 signal and its relatively fast relaxation support the assignment to a radical species in the proximity of the Mn cluster. The whole phenomenology of the g = 2.035 signal production is analogous to the effects of NIR excitation on the S3 state [Ioannidis, N., Nugent, J. H. A., and Petrouleas, V. (2002) Biochemistry 41, 9589-9600] producing an S2'Y(Z*) intermediate. In the present case, the intermediate is assigned to S1Y(Z*). The NIR-induced increase in the oxidative capability of the Mn cluster is discussed in relation to the photochemical properties of a Mn(III) ion that exists in both S2 and S3 states. The EPR properties of the S1Y(Z*) intermediate cannot be reconciled easily with our current understanding of the magnetic properties of the S1 state. It is suggested that oxidation of tyr Z alters the magnetic properties of the Mn cluster via exchange of a proton.  相似文献   

14.
Tarasev M  Pinto A  Kim D  Elliott SJ  Ballou DP 《Biochemistry》2006,45(34):10208-10216
Phthalate dioxygenase (PDO) and its reductase are parts of a two-component Rieske dioxygenase system that initiates the aerobic breakdown of phthalate by forming cis-4,5-dihydro-4,5-dihydroxyphthalate (DHD). Aspartate D178 in PDO, located near its ferrous mononuclear center, is highly conserved among Rieske dioxygenases. The analogous aspartate has been implicated in electron transfer between the mononuclear iron and Rieske center in naphthalene dioxygenase [Parales et al. (1999) J. Bacteriol. 181, 1831-1837] and in substrate binding and oxygen reactivity in anthranilate dioxygenase [Beharry et al. (2003) Biochemistry 42, 13625-13636]. The effects of substituting D178 in PDO with alanine or asparagine on the reactivity of the Rieske centers, phthalate hydroxylation, and coupling of Rieske center oxidation to DHD formation were studied previously [Pinto et al. (2006) Biochemistry 45, 9032-9041]. This work describes effects that D178N and D178A substitutions have on the interactions between the Rieske and mononuclear centers in PDO. The mutations affected protonation of the Rieske center histidine and conformation of subunits within the PDO multimer to create a more open structure with more solvent-accessible Rieske centers. When the Rieske centers in PDO were oxidized, D178N and D178A substitutions disrupted communication between the Rieske and Fe-mononuclear centers. This was shown by the lack of perturbations of the UV-vis spectra on phthalate binding to the D178N and D178A variants, as opposed to that observed in WT PDO. However, when the Rieske center was in the reduced state, communication between the centers was not disrupted. Phthalate binding similarly affected the rates of oxidation of the reduced Rieske center in both WT and mutant PDO. Nitric oxide binding at the Fe(II)-mononuclear center, as detected by EPR spectrometry of the Fe(II) nitrosyl complex, was regulated by the redox state of the Rieske center. When the Rieske center was oxidized in either WT or D178N PDO, NO bound to the mononuclear iron in the presence or absence of phthalate. However, when the Rieske center was reduced, NO bound only when phthalate was present. These findings are discussed in terms of the "communication functions" performed by the bridging Asp-178.  相似文献   

15.
The electrons extracted from the CaMn(4) cluster during water oxidation in photosystem II are transferred to P(680)(+) via the redox-active tyrosine D1-Tyr161 (Y(Z)). Upon Y(Z) oxidation a proton moves in a hydrogen bond toward D1-His190 (His(Z)). The deprotonation and reprotonation mechanism of Y(Z)-OH/Y(Z)-O is of key importance for the catalytic turnover of photosystem II. By light illumination at liquid helium temperatures (~5 K) Y(Z) can be oxidized to its neutral radical, Y(Z)(?). This can be followed by the induction of a split EPR signal from Y(Z)(?) in a magnetic interaction with the CaMn(4) cluster, offering a way to probe for Y(Z) oxidation in active photosystem II. In the S(3) state, light in the near-infrared region induces the split S(3) EPR signal, S(2)'Y(Z)(?). Here we report on the pH dependence for the induction of S(2)'Y(Z)(?) between pH 4.0 and pH 8.7. At acidic pH the split S(3) EPR signal decreases with the apparent pK(a) (pK(app)) ~ 4.1. This can be correlated to a titration event that disrupts the essential H-bond in the Y(Z)-His(Z) motif. At alkaline pH, the split S(3) EPR signal decreases with the pK(app) ~ 7.5. The analysis of this pH dependence is complicated by the presence of an alkaline-induced split EPR signal (pK(app) ~ 8.3) promoted by a change in the redox potential of Y(Z). Our results allow dissection of the proton-coupled electron transfer reactions in the S(3) state and provide further evidence that the radical involved in the split EPR signals is indeed Y(Z)(?).  相似文献   

16.
L Garfinkel  D Garfinkel 《Biochemistry》1984,23(15):3547-3552
We have attempted to resolve the differences between the levels of free Mg2+ in muscle calculated by Wu et al. [Wu, S. T., Pieper, G. M., Salhany, J. M., & Eliot, R. S. (1981) Biochemistry 20, 7399-7403] (2.5 mM in guinea pig heart) and by Gupta and Moore [Gupta, R. K., & Moore, R. D. (1980) J. Biol. Chem. 255, 3987-3993] (0.6 mM in frog skeletal muscle) on the basis of substantially identical measurements by 31P NMR of the phosphate peaks in the spectrum of MgATP2-. The differences depend on the methods of calculation, including which reactions in which multiple equilibria are being considered. Biochemists and physical chemists customarily use different working definitions of the stability constant for MgATP2- in particular. Wu et al. used in their calculations, without reconciliation, methods involving three different operational definitions of the chelation equilibria involved. An algorithm for calculating Mg2+ and total ATP, which can be carried out with a hand calculator, is described here. With it, we calculated Mg2+ levels that agree with those determined by Gupta et al. [Gupta, R. K., Benkovic, J. L., & Rose, Z. B. (1978) J. Biol. Chem. 253, 6165-6171] with their in vitro systems. We therefore agree with the finding of Gupta and Moore that the Mg2+ level in skeletal and cardiac muscle is 0.6 mM.  相似文献   

17.
We tried to confirm the antioxidative capability of lecithin:cholesterol acyltransferase (LCAT) reported by Vohl et al. [Biochemistry (1999) 38, 5976-5981]. The enzyme solution protected LDL against oxidation. However, this protection was not due to LCAT enzyme, but to some unknown low-molecular-weight substance(s) in the solution; LCAT itself exerted little protective effect against LDL oxidation.  相似文献   

18.
Electron paramagnetic resonance (EPR) and electron nuclear double resonance (ENDOR) were performed to investigate the difference in microenvironments and functions between tyrosine Z (Y(Z)) and tyrosine D (Y(D)). Mn-depletion or Ca(2+)-depletion causes extension of the lifetime of tyrosine radical Y(Z)(*), which can be trapped by rapid freezing after illumination at about 250 K. Above pH 6.5, Y(Z)(*) radical in Mn-depleted PS II shows similar EPR and ENDOR spectra similar to that of Y(D)(*) radical, which are ascribed to a typical neutral tyrosine radical. Below pH 6.5, Y(Z)(*) radical shows quite different EPR and ENDOR spectra. ENDOR spectra show the spin density distribution of the low-pH form of Y(Z)(*) that has been quite different from the high-pH form of Y(Z)(*). The spin density distribution of the low-pH Y(Z)(*) can be explained by a cation radical or the neutral radical induced by strong electrostatic interaction. The pH dependence of the activation energy of the recombination rate between Y(Z)(*) and Q(A)(-) shows a gap of 4.4 kJ/mol at pH 6.0-6.5. In the Ca(2+)-depleted PS II, Y(Z)(*) signal was the mixture of the cation-like and normal neutral radicals, and the pH dependence of Y(Z)(*) spectrum in Ca(2+)-depleted PS II is considerably different from the neutral radical found in Mn-depleted PS II. Based on the recent structure data of cyanobacterial PS II, the pH dependence of Y(Z)(*) could be ascribed to the modification of the local structure and hydrogen-bonding network induced by the dissociation of ASP170 near Y(Z).  相似文献   

19.
Nagai M  Aki M  Li R  Jin Y  Sakai H  Nagatomo S  Kitagawa T 《Biochemistry》2000,39(43):13093-13105
Heme structures of a natural mutant hemoglobin (Hb), Hb M Iwate [alpha87(F8)His-->Tyr], and protonation of its F8-Tyr were examined with the 244-nm excited UV resonance Raman (UVRR) and the 406.7- and 441.6-nm excited visible resonance Raman (RR) spectroscopy. It was clarified from the UVRR bands at 1605 and 1166 cm(-)(1) characteristic of tyrosinate that the tyrosine (F8) of the abnormal subunit in Hb M Iwate adopts a deprotonated form. UV Raman bands of other Tyr residues indicated that the protein takes the T-quaternary structure even in the met form. Although both hemes of alpha and beta subunits in metHb A take a six-coordinate (6c) high-spin structure, the 406.7-nm excited RR spectrum of metHb M Iwate indicated that the abnormal alpha subunit adopts a 5c high-spin structure. The present results and our previous observation of the nu(Fe)(-)(O(tyrosine)) Raman band [Nagai et al. (1989) Biochemistry 28, 2418-2422] have proved that F8-tyrosinate is covalently bound to Fe(III) heme in the alpha subunit of Hb M Iwate. As a result, peripheral groups of porphyrin ring, especially the vinyl and the propionate side chains, were so strongly influenced that the RR spectrum in the low-frequency region excited at 406.7 nm is distinctly changed from the normal pattern. When Hb M Iwate was fully reduced, the characteristic UVRR bands of tyrosinate disappeared and the Raman bands of tyrosine at 1620 (Y8a), 1207 (Y7a), and 1177 cm(-)(1) (Y9a) increased in intensity. Coordination of distal His(E7) to the Fe(II) heme in the reduced alpha subunit of Hb M Iwate was proved by the observation of the nu(Fe)(-)(His) RR band in the 441.6-nm excited RR spectrum at the same frequency as that of its isolated alpha chain. The effects of the distal-His coordination on the heme appeared as a distortion of the peripheral groups of heme. A possible mechanism for the formation of a Fe(III)-tyrosinate bond in Hb M Iwate is discussed.  相似文献   

20.
A comparative study using laser flash photolysis of the kinetics of reduction and intramolecular electron transfer among the redox centers of chicken liver xanthine dehydrogenase and of bovine milk xanthine oxidase is described. The photogenerated reductant, 5-deazariboflavin semiquinone, reacts with the dehydrogenase (presumably at the Mo center) in a second-order manner, with a rate constant (k = 6 x 10(7) M-1 s-1) similar to that observed with the oxidase [k = 3 x 10(7) M-1 s-1; Bhattacharyya et al. (1983) Biochemistry 22, 5270-5279]. In the case of the dehydrogenase, neutral FAD radical formation is found to occur by intramolecular electron transfer (kobs = 1600 s-1), presumably from the Mo center, whereas with the oxidase the flavin radical forms via a bimolecular process involving direct reduction by the deazaflavin semiquinone (k = 2 x 10(8) M-1 s-1). Biphasic rates of Fe/S center reduction are observed with both enzymes, which are due to intramolecular electron transfer (kobs approximately 100 s-1 and kobs = 8-11 s-1). Intramolecular oxidation of the FAD radical in each enzyme occurs with a rate constant comparable to that of the rapid phase of Fe/S center reduction. The methylviologen radical, generated by the reaction of the oxidized viologen with 5-deazariboflavin semiquinone, reacts with both the dehydrogenase and the oxidase in a second-order manner (k = 7 x 10(5) M-1 s-1 and 4 x 10(6) M-1 s-1, respectively). Alkylation of the FAD centers results in substantial alterations in the kinetics of the reaction of the viologen radical with the oxidase but not with the dehydrogenase. These results suggest that the viologen radical reacts directly with the FAD center in the oxidase but not in the dehydrogenase, as is the case with the deazaflavin radical. The data support the conclusion that the environments of the FAD centers differ in the two enzymes, which is in accord with other studies addressing this problem from a different perspective [Massey et al. (1989) J. Biol. Chem. 264, 10567-10573]. In contrast, the rate constants for intramolecular electron transfer among the Mo, FAD, and Fe/S centers in the two enzymes (where they can be determined) are quite similar.  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