Abstract
The cytochrome bc1 complex (ubiquinone:cytochrome c oxidoreductase) is the central integral membrane protein in the mitochondrial respiratory chain as well as the electron-transfer chains of many respiratory and photosynthetic prokaryotes. Based on X-ray crystallographic studies of cytochrome bc1, a mechanism has been proposed in which the extrinsic domain of the iron-sulfur protein first binds to cytochrome b where it accepts an electron from ubiquinol in the Qo site, and then rotates by 57o to a position close to cytochrome c1 where it transfers an electron to cytochrome c1. This review describes the development of a ruthenium photooxidation technique to measure key electron transfer steps in cytochrome bc1, including rapid electron transfer from the iron-sulfur protein to cytochrome c1. It was discovered that this reaction is rate-limited by the rotational dynamics of the iron-sulfur protein rather than true electron transfer. A conformational linkage between the occupant of the Qo ubiquinol binding site and the rotational dynamics of the iron-sulfur protein was discovered which could play a role in the bifurcated oxidation of ubiquinol. A ruthenium photoexcitation method is also described for the measurement of electron transfer from cytochrome c1 to cytochrome c. This article is part of a special issue entitled: Respiratory Complex III.
Keywords: Cytochrome bc1, cytochrome c, electron transfer, ruthenium
1. Introduction
The cytochrome bc1 complex (cyt bc1) (ubiquinone:cytochrome c oxidoreductase) is the central integral membrane protein in the mitochondrial respiratory chain as well as the electron transfer chains of many respiratory and photosynthetic prokaryotes [1,2]. The overall net reaction catalyzed by cyt bc1 involves the 2-electron oxidation of ubiquinol (QH2) to ubiquinone (Q), and the reduction of two molecules of cytochrome c (Cc). The energy of electron transfer is coupled to the uptake of two protons from the inside of the membrane, and the release four protons to the outside of the membrane (equation 1):
| (1) |
Mitochondrial cyt bc1 is a homodimer with 11 polypeptide chains, while prokaryotic cyt bc1 complexes contain as few as 3 polypeptide chains [1,2]. Cyt bc1 contains three redox proteins: cyt b with two b hemes (bL and bH), the Rieske iron-sulfur protein (ISP) containing a 2Fe2S cluster, and cyt c1 with one c-type heme. In the widely accepted Q-cycle mechanism, QH2 binds to the Qo site near the outside of the membrane and transfers its first electron to the Rieske iron-sulfur center 2Fe2S, which is then transferred to cyt c1 and finally to Cc [1-4]. The second electron is transferred from semiquinone in the Qo site to cyt bL, and then to cyt bH and ubiquinone in the Qi site to form semiquinone. The cycle is repeated to reduce semiquinone in the Qi site to QH2. X-ray crystallographic studies have shown that the conformation for the extrinsic domain of the Rieske iron-sulfur protein depends on the crystal form and the presence of Qo site inhibitors [5-8]. In crystals of cyt bc1 from all species grown in the presence of stigmatellin, the ISP is in a conformation with 2Fe2S close to the cyt bL heme, called the b state [5-11] (Figure 1). In contrast, the ISP is in a conformation with 2Fe2S close to cyt c1, called the c1 state, in native chicken or beef P6522 crystals in the absence of inhibitors [6,7]. The intensity of the anomalous signal for 2Fe2S close to cyt bL is small in bovine I4122 crystals in the absence of inhibitors, indicating that the ISP is conformationally mobile (5,8). A rotational shuttle mechanism involving the extrinsic domain of the ISP has been proposed that is supported by these structural studies (Figure 1, Scheme 1) [5-8]. QH2 in the Qo site transfers an electron to the oxidized 2Fe2S center of the ISP in the b state, followed by rotation of the ISP to the c1 state and electron transfer from reduced 2Fe2S to cyt c1. This mobile shuttle mechanism is supported by extensive mutational, cross-linking, and kinetic studies [12-28].
Figure 1.
X-ray crystal structure of chicken cyt bc1 in b state in presence of stigmatellin and antimycin (PDB: 3H1I ) [6], and of beef P65 22 crystals in c1 state (PDB:1BE3) [7]. The ISP, cyt c1 and cyt b subunits are colored blue, green, and gray, respectively. The hemes, 2Fe2S, stigmatellin, and antimycin are colored red, yellow, cyan, and green. The Rieske neck region residues 66-72 are colored orange, and the ef loop residues 252-268 are colored yellow.
Scheme 1.

An important goal is to measure the rate constants for all of the electron-transfer reactions in cyt bc1, as well as the dynamics of the conformational changes of the extrinsic domain of the ISP. This has been a challenging goal, since many of the electron-transfer reactions are very rapid. This review describes the development and use of a ruthenium photooxidation technique to study electron transfer in cytochrome bc1 with microsecond time resolution [13]. This technique has been used to study the key electron-transfer steps in the mechanism of cyt bc1 as well as the dynamics of the ISP extrinsic domain rotation.
2. Design of Photoactive Ruthenium Complexes
Ruthenium polypyridine complexes have a number of remarkable properties that make them excellent photoredox initiators [29]. They have a long-lived metal-to-ligand charge-transfer excited state that is both a strong oxidant and a strong reductant, and are very stable in both the ground state and the excited state. Ruthenium complexes can be used to rapidly photoreduce or photooxidize neighboring redox centers. In the photooxidation mechanism of scheme 2, the excited state Ru(II*) accepts an electron from Fe(II) to form Ru(I)---Fe(III) with rate constant k3. A sacrificial electron acceptor A can oxidize Ru(I) to Ru(II) with rate constant k8, preventing the k4 back reaction to Ru(II)---Fe(II). All three chelating ligands of ruthenium can be altered to tune the redox potentials over a wide range, and optimize the rate and yield of photooxidation or photoreduction. The effects of different ligands on the redox potentials of the ruthenium complexes are shown in Table I. The Ru(bpz)2(dmb) complex is particularly well-suited for photooxidation applications. We have introduced four different strategies for specifically labeling proteins with photoredox active ruthenium polypyridyl complexes [30-37]. The most useful method involves the formation of a thioether link between a protein cysteine residue and a ruthenium complex containing a bromomethyl group on the heterocyclic ring [34-37]. The location of the cysteine residue on the protein can be genetically engineered to address specific questions [30-37]. We have recently designed binuclear ruthenium complexes which bind non-covalently to CcO and cyt bc1 and initiate photoinduced electron transfer [13,38, 39]. [Ru(bpy)2]2(qpy)4+ (Ru2D) (Figure 2) has a net charge of +4 which allows it to bind strongly to the acidic domain on cyt bc1 and photooxidize cyt c1 with a yield of 25% according to Scheme 2 [13].
Scheme 2.
Table 1.
Standard reduction potentials (in V) of ruthenium complexes vs. normal hydrogen electrode
| Complex | (II)/(III) | (II*)/(III) | (II)/(I) | (II*)/(I) |
|---|---|---|---|---|
| Ru(bpy)3 | 1.27 | −0.87 | −1.31 | 0.83 |
| Ru(bpy)2(dmb) | 1.27 | −0.83 | −1.36 | 0.79 |
| Ru(bpz)2(dmb) | 1.76 | −0.25 | −0.79 | 1.22 |
| Ru(bpd)2(dmb) | 1.49 | −0.49 | −1.00 | 0.98 |
| ||||
Figure 2.
Modeled structure of Ru2D. The carbons are colored grey, the nitrogens are blue, and the rutheniums are green.
Extensive studies on a wide range of ruthenium-labeled proteins have provided important information on the dependence of electron transfer on driving force, distance and pathway [40-49]. The theory developed by Marcus has revealed that three important factors control the rate of electron transfer, the free energy change ΔGo’ of the redox reaction, the reorganization energy λ, and the electronic coupling between the redox centers [50]. The reorganization energy λ is a measure of the energy required to rearrange and repolarize the reactants and surrounding solvent before electron transfer can occur. Dutton and coworkers have reported that the rate constants in a broad range of biological systems can be described approximately by a simple exponential dependence on the distance between the redox centers, as originally proposed by Marcus [51]:
| (2) |
where r is the distance between the closest macrocycle atoms in the two redox centers, the van der Waals contact distance ro = 3.6 Å, β = 1.4 Å−1, and the nuclear frequency ko = 1013 s−1.
3. Kinetics of Electron Transfer within Cytochrome bc1
The electron-transfer reactions of mitochondrial cyt bc1 has been studied extensively by stopped-flow spectroscopy and rapid-mix/freeze-quench EPR [52,53]. In the absence of inhibitors, the reduction of heme bH by ubiquinol in bovine cyt bc1 is multiphasic. The fast phase is inhibited by antimycin, indicating that it occurs through the Qi site. The reduction of heme bH by 300 μM duroquinol through the Qo site in the presence of antimycin occurred with a half-time of 25 ms, while cyt c1 and the 2Fe2S center were also reduced with the same half-time, indicating that they are in rapid equilibrium [53]. A rapid technique using photo-releasable decylubiquinol was also used to study reduction of heme bH and cyt c1 [54]. In another study, it was found that the rate of reduction of yeast cyt bc1 by menaquinol was significantly decreased by antimycin, suggesting that the redox states of heme bL and heme bH control the reduction of the 2Fe2S center [55]. The effects of Qo and Qi site inhibitors on the kinetics of yeast cyt bc1 led Trumpower to propose an alternating, half-of-the-sites mechanism [56].
A photoexcitation method has been developed to study the kinetics of cyt bc1 electron transfer in chromatophores of photosynthetic bacteria including Rhodobacter sphaeroides and R. capsulatus [57]. Photoexcitation of the photosynthetic reaction center rapidly oxidizes cyt c2, which diffuses to cyt bc1 and oxidizes cyt c1 with a half-time of 150 μsec [58,59]. Cyt c1 then oxidizes the 2Fe2S center, allowing bifurcated electron transfer from QH2 in the Qo site to 2Fe2S and heme bL and heme bH. The following rate constants, which are identified in Scheme 1, have been estimated for the cyt bc1 reactions in R. sphaeroides chromatophores : k1 > 104 s−1 for electron transfer from cyt c1 to Cc, k2 > 105 s−1 for electron transfer from 2Fe2S to cyt c1, k3 = 1650 s−1 for electron transfer from QH2 to 2Fe2S in the Qo site, k4 > 109 s−1 for electron transfer from Q•− to heme bL in the Qo site, k5 > k3, and k6 > k3 [59]. Shinkarev et al. [60] have determined that electron transfer between cyt bL and cyt bH has a half-time of 0.1 ms using the transient electric field generated by excitation of the reaction center to initiate reverse electron transfer from cyt bH to cyt bL.
A laser flash photolysis method was developed to study electron transfer within the cyt bc1 complex using the binuclear ruthenium complex Ru2D [13] (Schemes 2, 3). Rb. sphaeroides cyt bc1 is typically redox poised with cyt c1 and 2Fe2S reduced, heme bL oxidized, and heme bH and quinol partially reduced (Scheme 3). Laser excitation of Ru2D to the metal-to-ligand charge transfer state, Ru2D*, a strong oxidant, leads to oxidation of cyt c1 within 700 ns, as indicated by the rapid decrease in the reduced cyt c1 absorbance band at 552 nm (Figure 3). A sacrificial electron acceptor A such as [Co(NH3)5Cl]2+ is present in the solution to oxidize RuII* and/or RuI and promote oxidation of cyt c1 by either of the pathways shown in Scheme 2. The +4 charge on Ru2D allows it to bind selectively to the negatively charged domain on the surface of cyt c1 with a dissociation constant of 8 μM [13]. The photooxidation of cyt c1 is complete within 670 ns, the lifetime of the Ru2D excited state, indicating that Ru2D must be very close to cyt c1 at the time of the flash [13,61]. Ru2D does not photooxidize other proteins that do not have a negatively charged domain, such as cytochrome c, and does not photooxidize the iron-sulfur center, cyt bL, or cyt bH in cyt bc1 [61].
Scheme 3.

Figure 3.
Electron transfer within wild-type R. sphaeroides cyt bc1 initiated by photooxidation of cyt c1 [68]. The sample contained 5 μM cyt bc1, 20 μM Ru2D, 5 mM [Co(NH3)5Cl]2+, in 20 mM sodium borate, pH 9.0 with 0.01% dodecylmaltoside. Treatment of cyt bc1 with 10 μM QoC10BrH2, 1 mM succinate, and 50 nM SCR completely reduced 2Fe2S and cyt c1, and reduced cyt bH by 30%. Cyt c1 was photooxidized within 1 μs, and then reduced with rate constants of 80,000 s−1 and 2,000 s−1, as indicated in the 552 nm transient. The rate constant for the reduction of cyt bH measured at 561 – 569 nm was 2,300 s−1. (Bottom two traces) Addition of 30 μM famoxadone decreased the rate of reduction of cyt c to 5,400 s−1 1 and eliminated reduction of cyt bH.
The photooxidation of cyt c1 by Ru2D* is followed by biphasic reduction of cyt c1 with rate constants of 80,000 s−1 and 2,000 s−1 (Figure 3) [13]. The fast phase has been assigned to electron transfer from reduced 2Fe2S to photooxidized cyt c1 with rate constant k2, while the slow phase of cyt c1 reduction is correlated with the oxidant-induced reduction of heme bH, monitored at 561 - 569 nm (Figure 3). The oxidant-induced reduction of cyt bH is rate-limited by transfer of the first electron from QH2 to 2Fe2S with rate constant k3. The subsequent transfer of the second electron from the semiquinone to heme bL and heme bH with rate constants k4 and k5 is much more rapid than k3, and not rate-limiting. The kinetics of both electron transfer from 2Fe2S to cyt c1 and from QH2 to 2Fe2S can thus be resolved by this technique. Electron transfer in the bovine cyt bc1 complex and Paracoccus denitrificans cyt bc1 has also been studied using this technique [13,61]. The rate constants for bovine cyt bc1 are k2 = 16,000 s−1 and k3 = 250 s−1, while for the P. denitrificans complex they are k2 = 10,700 s−1 and k3 = 700 s−1.
An important question is whether the rate constant k2 for electron transfer from 2Fe2S to cyt c1 is rate-limited by true electron transfer, proton gating, or conformational gating [21]. The most definitive way to discriminate between true electron transfer and conformational gating mechanisms is by changing the driving force of the reaction, since Marcus theory predicts a large dependence of the rate of true electron transfer on driving force. The redox potential of 2Fe2S is decreased significantly as the pH is increased from pH 7.0 to pH 10.0 due to the deprotonation of the 2Fe2S ligand His-161, leading to an increase in the driving force ΔGo’ from −0.02 V to +0.115 V [62,63]. However, the rate constant k2 for electron transfer from 2Fe2S to cyt c1 in R. sphaeroides cyt bc1 was independent of pH, even though Marcus theory predicts that the rate constant should increase 12-fold as the pH is increased from 7.0 to 10.0 [21]. This Marcus theory calculation using equation (2) is based on the assumption that λ = 1.0 V. Although λ has not been measured for this reaction, a value of 1.0 V is typical for cytochromes with a similar heme solvent exposure to that of cyt c1 [21,43,49]. Moreover, the rate constant k2 was not affected by the ISP mutations Y156W, S154A, and Y156F/ S154A which decrease the redox potential of 2Fe2S by 62 mV, 109 mV, and 159 mV, respectively (Table 2) [21]. Marcus theory, equation 2, predicts that the increase in the driving force in these mutants would increase the rate constant by up to 17-fold (Table 2), assuming λ = 1.0 V [21]. These studies indicate that the rate constant k2 for electron-transfer from the 2Fe2S center to cyt c1 is controlled by conformational gating rather than the rate of transfer of the electron. For this conformational gating mechanism to be valid, the fluctuations in the conformation of the Rieske iron-sulfur protein must be slow compared to electron transfer in the active c1 state. A model for the active c1 state is provided by the bovine P6522 crystal structure [6, 7], in which the 2Fe2S ligand His-161 forms a hydrogen bond with the heme c1 propionate oxygen (Figure 4). There is a pathway for electron transfer from the 2Fe2S center to heme c1 that has a distance of 7.8 Å from the His-161 nitrogen to the closest heme c1 macrocycle atom C3D (Figure 4). Equation 1 predicts a rate constant k2 ranging from 1.5 × 106 to 3 × 107 s−1 for this pathway, assuming λ values between 1.0 and 0.7 eV, which are typical for cytochromes [43,49]. The observed rate constant k2 for electron transfer from 2Fe2S to cyt c1 is considerably smaller than the predicted value for the c1 state, consistent with a conformational gating mechanism.
Table 2.
Kinetic properties of R. sphaeroides cyt bc1 mutants [21]. Enzymatic activity in μmol cyt c reduced/min/μmol cyt b at 25 °C. ΔEm is the difference in redox potential between 2Fe2S and cyt c1 at pH 8.0, 25 °C. k2 is the experimental rate constant for electron transfer from 2Fe2S to cyt c1 at pH 8.0, 25 °C. Theoretical rate constant for electron transfer is calculated from equation 2 with r = 9.9 Å, λ = 1.0 eV, and ΔGo calculated from the ΔEm value of the mutant cyt bc1.
| Mutant | Enzymatic activity |
ΔEm (mV) | k2 (104 s−1) (experimental) |
k2 (104 s−1) (theory) |
|---|---|---|---|---|
| Wild-type | 2.5 | 0 | 8.0 | 8.0 |
| Y156W | 0.58 | −62 | 15.0 | 26.0 |
| S154A | 0.23 | −109 | 7.8 | 60.0 |
| S154A/Y156F | 0.03 | −159 | 9.0 | 140.0 |
Figure 4.
Structure of bovine cyt bc1 P6522 crystals in the c1 state (PDB:1BE3) [7]. The Rieske and cyt c1 subunits are colored dark blue and light blue, respectively, the 2Fe2S center is shown as a CPK model colored red/yellow, and heme c1 is colored by element. His-161, Ser-163,Tyr-165, and Cys-139 are shown as sticks. The hydrogen bond between the Nε2 nitrogen of His-161 and the heme c1 propionate oxygen is shown with a line. The distance of 7.8 Å from the His-161 nitrogen to the closest heme c1 macrocycle atom C3D is indicated by the red arrow. Tyr-165 and Ser-163 in the bovine ISP are homologous to Tyr-156 and Ser-154 in the R. sphaeroides ISP.
A central question about the bifurcated reaction at the Qo site is how QH2 can deliver 2 electrons sequentially to the high and low potential chains, while avoiding short–circuit and bypass reactions. It has been proposed that the controlled motion of the ISP extrinsic domain may play a role in a gating mechanism for the electron bifurcation reaction at the Qo site [10,12,28]. X-ray crystallography studies have shown that inhibitors bound at the Qo site have a significant influence on the position and mobility of the ISP. Binding Pm type inhibitors such as myxothioazol to the Qo site lead to the release of ISP from the b-state to a disordered “mobile” state not detected in the crystal, while structures complexed with Pf type inhibitors such as stigmatellin show that ISP binds to the b subunit in a “fixed” state [8, 10, 25, 28]. EPR studies have also provided valuable information on the effect of Qo site inhibitors on the position and orientation of the ISP [17,22,26,27]. Havens et al. [61] have studied the effects of six different Qo site inhibitors on electron transfer from ISP to cyt c1 in P. denitrificans cyt bc1 using the Ru2D photooxidation technique. Binding any of the Pm inhibitors MOA-stilbene, myxothiazol, or azoxystrobin to cyt bc1 increased the rate and extent of electron transfer from ISP to cyt c1, consistent with release of ISP from the b state which causes a linked decrease in the redox potential and increase in the mobility of the ISP (Table 3). Binding the Pf inhibitor stigmatellin completely prevented electron transfer from ISP to cyt c1, consistent with X-ray crystallography studies showing that stigmatellin locks the ISP in the b-state with a hydrogen bond between a carbonyl group of the inhibitor and His-161, a ligand of the 2Fe2S cluster [8,10,64-66]. In contrast, binding the Pf type inhibitors JG-144 and famoxadone decreased the rate constant by 5 to 10-fold, and increased the amplitude over 2-fold. These inhibitors therefore do not lock the ISP in the b state, but rather decrease the rate of its release from the b state and rotation to the c1 state. Binding reduced QH2 leads to a two-fold increase in the amplitude of the fast phase, A1f. These results indicate that the species occupying the Qo site has a significant effect on the dynamics of the ISP domain rotation.
Table 3.
Effects of Qo site inhibitors on electron transfer in P.denitricans cyt bc1 [61]. k2f and A2f are the rate constant and amplitude of the fast phase of electron transfer between 2Fe2S and cyt c1, respectively. Solutions contained 5 μM cyt bc1, 20 μM Ru2D, 1 mM ascorbate, 4 μM TMPD, and 5 mM [Co(NH3)5Cl]2+ in 20 mM TRIS-HCl pH 8.0. Inhibitor concentrations were 25 μM.
| inhibitor/substrate | type | k2f (s−1) | A2f |
|---|---|---|---|
| none | -- | 6,300 | 10 % |
| MOAS | Pm | 9,900 | 33 % |
| myxothiazol | Pm | 8,900 | 30 % |
| azoxystrobin | Pm | 8,000 | 28 % |
| stigmatellin | Pf | 0 | 0 |
| JG-144 | Pf | 1,300 | 16 % |
| famoxadone | Pf | 600 | 26 % |
| Q | -- | 5,300 | 10 % |
| QH2 | -- | 10,700 | 18 % |
X-ray crystallography studies of bovine cyt bc1 have shown that binding Pf type inhibitors such as famoxadone displace the cd1 helix and the ef helix away from each other to widen the Qo pocket and form a binding crater for the capture of the ISP in the b-state (Figures 5, 6) [28]. Photoactivated ruthenium kinetic studies have shown that famoxadone binding does not completely immobilize the ISP in the b state, but rather slows down the rate of its release and rotation to the c1 state. Famoxadone binding to bovine cyt bc1 decreased k2 from 16,000 s−1 to 1,500 s−1, while in R. sphaeroides cyt bc1 k2 was decreased from 60,000 s−1 to 5,400 s−1 [67] (Figure 3). A series of mutants at residues in the ef loop were constructed to explore the role of the ef loop in regulating the dynamics of the ISP [68] (Table 4; Figure 5). The mutation Y280A caused a decrease in k2 from 60,000 s−1 to 7,900 s−1, but famoxadone binding only decreased k2 to 3,200 s−1. Similarly, the I292A mutation decreased k2 to 4,400 s−1, but famoxadone binding only decreased it to 3,000 s−1. These mutations might cause a conformational change similar to that of famoxadone, limiting the additional effect of famoxadone binding. The I292A mutation caused a decrease in the rate constant k3 for electron transfer from QH2 to 2Fe2S from 2,300 s−1 to 350 s−1, indicating an effect on the conformation of the QH2 reaction site. Mutation of L286A at the tip of the ef loop had an interesting effect, decreasing k2 to 33,000 s−1 and k3 to 740 s−1. However, famoxadone binding does not lead to any further decrease in k2, suggesting that this mutation might block the famoxadone-induced conformational change in the wild-type protein. Darrouzet et al. have also carried out experiments indicating the importance of L286 [18,23].
Figure 5.
X-ray crystal structure of bovine cyt bc1 in the presence of famoxadone (PDB:1LOL) [58]. Famoxadone is colored cyan, the cyt c1 and cyt bL hemes are red, and the 2Fe2S center is represented by a CPK model. The ISP is blue, and cyt b is grey. Residues 252-268 in the ef loop are colored orange while residues 269-283 in the PEWY sequence and the ef helix are red. Residues 136-152 in the cd1 helix are green and residues 163-171 in the neck-contacting domain are colored yellow. Residues of interest are indicated by sticks, and labeled with R. sphaeroides sequence numbering.
Figure 6.
X-ray crystal structure of ISP bound to the docking crater on cyt b with stigmatellin bound to bovine cyt bc1 (PDB:1SQX) [28]. View is parallel to membrane, showing ef loop and helix (gray), cd1 helix (green), and ISP (orange). The H-bond between stigmatellin and the His-152 ligand to 2Fe2S is shown with a line. Residues on cyt b near the Qo site or interacting with the ISP are shown as sticks, and labeled with R. sphaeroides sequence numbering.
Table 4.
Effect of mutations on electron transfer within R sphaeroides. cyt bc1 [68]. The rate constant k2 for electron transfer from 2Fe2S to cyt c1 was measured in a solution containing 5 μM cyt bc1, 20 μM Ru2D, 5 mM [Co(NH3)5Cl]2+, in 20 mM sodium borate, pH 9.0, 0.01% dodecylmaltoside. The cyt bc1 was treated with 10 μM QoC10BrH2, 1 mM succinate, and 50 nM SCR to completely reduce 2Fe2S and cyt c1, and reduce cyt bH by about 30%. Famoxadone (30 μM) was added where indicated. The rate constant k3 for electron transfer from QH2 in the Qo site to 2Fe2S was measured without famoxadone.
| Mutant | Activity | k2 (s−1) | k2 (s−1) with famoxadone |
k3 (s−1) |
|---|---|---|---|---|
| WT | 2.35 | 60,000 | 5,400 | 2,300 |
| Y280A (b) | 1.34 | 7,900 | 3,200 | 2,800 |
| L286A (b) | 0.78 | 33,000 | 35,000 | 740 |
| I292A (b) | 0.81 | 4,400 | 3,000 | 350 |
| P150C (ISP) | 0.23 | 50,000 | 2,000 | 4 |
| G153C (ISP) | 0.78 | 13,000 | 1,470 | 450 |
Extensive research has been carried out on the bifurcated electron transfer reaction at the Qo site where QH2 transfers the first electron to the ISP and cyt c1, and the resulting semiquinone transfers the second electron to cyt bL and cyt bH [28,69-79]. Potential mechanisms must be consistent with the reversibility of the reaction, and prevent short-circuit reactions, including the delivery of both electrons from ubiquinol to 2Fe2S and cyt c1 in the high potential chain. Double gating mechanisms have been proposed in which QH2 can only react if b-state ISP and cyt bL are both initially oxidized [69,76-79]. Crofts and colleagues have proposed a coulombic gating mechanism in which the semiquinone anion moves from the distal site near the ISP to a site near oxidized cyt bL in a process controlled by electrostatics [71,77]. Another possibility is simultaneous transfer of two electrons from QH2 to the ISP and cyt bL in a concerted reaction without the formation of a semiquinone intermediate [69,76,78-80]. However, a semiquinone radical at the Qo site has been detected at the Qo site by two different methods [81,82], and a semiquinone is also thought to be required in the bypass reaction during formation of superoxide [83].
Unfortunately, it has not been possible to experimentally detect QH2 or Q in the Qo binding pocket by X-ray crystallography. However, the effects of Qo site inhibitors on the structural linkage between the conformations of cyt b and the ISP have led to a proposal for the mechanism for bifurcated electron transfer [12,28]. It was proposed that binding QH2 to the Qo site widens the Qo pocket between the cd1 helix and ef helices forming a crater to bind the ISP in the b-state (Figure 6) [12,28]. This would promote the formation of a hydrogen bond between QH2 and His-161 and lead to proton-coupled electron transfer from QH2 to oxidized 2Fe2S. After the second electron was transferred from semiquinone to cyt bL and cyt bH, the resulting oxidized Q would leave the distal Qo binding pocket, triggering the cd1 and ef helices to come closer together and release the ISP from the docking crater, allowing it to rotate to the c1 position and transfer an electron to cyt c1 [12,28]. The effects of Pm and Pf inhibitors on the rapid kinetics of bovine, R. sphaeroides, andP. denitiricans cyt bc1 provide evidence that the conformations of the Qo site, the ISP docking crater, and the ISP extrinsic domain orientation and dynamics are tightly linked. It is suggested that binding Pf inhibitors to the Qo site leads to a conformation similar to that of the active QH2 – oxidized ISP complex, while binding Pm inhibitors leads to a conformation in which the ISP is released from the b-state to a mobile state. The linkage between the Qo site and the ISP conformation and dynamics may play an important role in gating the electron transfer bifurcation reaction in the Qo site to minimize short-circuit and bypass reactions. Other factors are also likely to be involved in gating, including coulombic gating of the motion of the semiquinone [72,77], and the conformations of water or amino acid side chains in the Qo pocket [69,75-79]. There is also evidence that events at the Qi site and the bL and bH hemes might be linked to turnover at the Qo site [26,27,84-89]. In addition, conformational interactions and electron transfer between the two monomers of the cyt bc1 dimer might play a role in the mechanism of the reaction at the Qo site [85,89-92].
4. Reaction between Cytochrome bc1 and Cytochrome c
Cytochrome c is a hydrophilic heme protein with a molecular weight of 12,500 Da that transfers electrons from the cyt bc1 complex to cytochrome c oxidase by a diffusional shuttle mechanism. The electron-transfer reaction from the cyt bc1 complex to Cc involves three steps: a) formation of a 1:1 reactant complex between reduced cyt bc1 and Cc3+, b) intracomplex electron transfer from cyt c12+ to Cc3+, and c) release of the product Cc2+. The overall rate of the reaction is optimized when the interaction between Cc and cyt bc1 stabilizes a reactant complex which allows rapid electron transfer, and also promotes rapid reactant complex formation and product complex dissociation. The steady-state reaction rate decreases with increasing ionic strength, indicating the involvement of electrostatic interactions between the two proteins [93-95]. Extensive chemical modification studies have demonstrated that six lysine amino groups surrounding the heme crevice of Cc are involved in binding to cyt bc1 [93-96]. Chemical modification and cross-linking studies have shown that acidic residues on cyt c1 and subunit 8 in bovine cyt bc1 are involved in binding Cc [97, 98]. X-ray crystal structures of beef, chicken, yeast, and R. sphaeroides cyt bc1 reveal that the cyt c1 heme edge on the cytoplasmic surface is surrounded by acidic residues that could form a binding site for Cc [5-9]. Most importantly, the X-ray crystal structure of the complex between yeast Cc and yeast cyt bc1 revealed that it is stabilized by non-polar interactions at the center of the binding domain, including a planar stacking interaction between yCc Arg-13 and Phe-230 of cyt c1 (Figure 7) [99, 100]. There are only two direct polar interactions in the binding domain, but additional charged residues around the periphery of the binding domain may contribute to the electrostatic interaction. The distance between the edges of the heme c and heme c1 groups is 9.4 Å.
Figure 7.
X-ray crystal structure of the complex between yeast cyt bc1 and yCc [99] (PDB: 1KYO). yCc is colored light blue, Cyt c1 is gray, the heme groups are red, basic residues on yCc are blue, acidic residues on cyt c1 are red, and Phe-230 is purple. The ruthenium complex on Cys-39 (green) was attached to the crystal structure by molecular modeling.
Stopped-flow spectroscopy has been used to measure the second-order electron transfer reaction between Cc and bovine cyt c1 at high ionic strength, but the reaction becomes too fast to resolve by this technique below 200 mM ionic strength [52]. In R. sphaeroides chromatophores, the reaction is rate-limited by the diffusion of photooxidized cyt c2 from the reaction center to the cyt bc complex with an apparent rate constant of 5000 s−1 1 [58, 59]. These techniques have provided valuable information about the reaction between cyt c1 and Cc, but it was not possible to measure the intracomplex rate constant.
It was necessary to design a new ruthenium-labeled Cc derivative to study rapid electron transfer from cyt bc1 to Cc in the forward, physiological direction. A brominated Ru(bpz)2(dmb) reagent was used to label the Cys-39 sulfhydryl group on yeast H39C,C102T Cc to form Ruz-39-Cc (Figure 7) [101]. The ruthenium complex is on the surface opposite from the heme crevice of Cc, and does not affect the interaction with yeast cyt bc1. There is an efficient pathway for electron transfer between the heme and the ruthenium complex consisting of 13 covalent bonds and one hydrogen bond, with a distance of 12.6 Å. The new Ru(bpz)2(dmb) complex has a reduction potential of 1.22 V for the Ru(II*)/Ru(I) transition. The driving force of 1.0 volt for the Ru(II*)-Fe(II) → Ru(I)-Fe(III) reaction is close to the expected reorganization energy λ of 0.8 V, which should allow optimal photooxidation of the reduced heme c according to Scheme 2. The photoinitiated electron transfer from heme c Fe(II) to Ru(II*) in Ruz-39-Cc occurred with a rate constant of k3 = 1.5 × 106 s−1, followed by back electron transfer from Ru(I) to Fe(III) with a rate constant of k4 = 7000 s−1 [101]. The back reaction is prevented in the presence of atmospheric oxygen, which rapidly oxidizes Ru(I). The yield of photooxidized heme c is 20% in a single flash.
Laser excitation of reduced yeast Ruz-39-Cc and yeast cyt bc1 at 250 mM ionic strength led to rapid photooxidation of heme c, followed by electron transfer from cyt c1 to oxidized heme c with a rate constant of 3900 s−1, as monitored at 546 nm (Figure 8) [101]. The oxidation of cyt c1 was observed directly at 557 nm, which is an isobestic point for Cc. A fast phase with a rate constant of 14,000 s−1 was observed at ionic strengths below 150 mM due to electron transfer in a preformed Ru-39-Cc:cyt bc1 complex (Figure 9). The rate constant of this intracomplex electron transfer reaction was independent of ionic strength from 5 mM to 120 mM, indicating that the complex does not change its configuration. Equation 1 was used to calculate a theoretical rate constant for electron transfer from cyt c1 to Cc based on the X-ray crystal structure of the complex between yeast iso-1-Cc and yeast cyt bc1 [99] (Figure 7). The calculated rate constant is between 1.8 × 105 s−1 and 3.3 × 106 s−1, assuming a reorganization energy λ between 0.7 and 1.0 V, and an edge-to-edge separation of 9.4 Å between the heme c and heme c1 groups as given in the crystal structure. Although the theoretical value is larger than the experimental value of 1.4 × 104 s−1, the through-water jump of 4.5 Å between the two hemes in the crystallographic complex could give a large barrier to electron transfer that is not accounted for by equation (2).
Figure 8.
Photoinduced electron transfer between yeast Ru-39-Cc and yeast cyt bc1 [101]. The solution contained 5.2 μM yeast Ruz-39-Cc and 4.4 μM yeast cyt bc1 in 5 mM sodium phosphate, pH 7.0, 250 mM NaCl, and 0.1% lauryl maltoside. It was treated with 2 μM TMDP and 10 μM ascorbate to reduce the c1 and c hemes. The 550 nm transient shows the photooxidation and reduction of Ru-39-Cc, while the 557 nm transient shows the oxidation of cyt c1. The transients at both wavelengths indicate electron transfer from cyt c1 to Cc with a rate constant of 3900 ± 600 s−1.
Figure 9.
Ionic strength dependence of the rate constant for photoinduced electron transfer between yeast Ruz-39-Cc and yeast cyt bc1. The conditions were the same as in Figure 8 except that 0 to 800 mM NaCl was present. The square root of ionic strength is in units of [M]1/2.
Since both intracomplex and bimolecular phases are observed at 110 mM ionic strength, the bimolecular reaction involves formation of a 1:1 complex with rate constant kf = 2.0 × 109 M−1s−1, intracomplex electron transfer with ket = 14,000 s−1, and complex dissociation with kd = 1.7 × 103 s−1 [101]. The fast intracomplex phase disappears and the rate constant of the bimolecular phase increases to a maximum at 200 mM ionic strength (corresponding to (I)1/2 = 0.45 in (M)1/2 units), indicating an increase in kd (Figure 9). The second-order rate constant decreases with increasing ionic strength above 250 mM ionic strength, consistent with a reaction between oppositely charged proteins (Figure 9). Rajagukguk et al. [102] prepared a series of yeast Ruz-39-Cc mutants containing mutations of residues at the binding domain in order to characterize the interaction with cyt bc1. The rate constants were measured using the ruthenium photooxidation technique at 250 mM, where the reaction is bimolecular (Table 5). The largest effect was observed for the R13A mutant, where the rate constant was decreased from 3500 s−1 to 153 s−1. This indicates that the π-cation interaction between yCc Arg-13 and Phe-230 of cyt c1 observed in the crystal structure of the complex [99,100] is important for the reaction in solution. A substantial decrease in rate constant to 1090 s−1 and 190 s−1 for the K86A and K86D mutants, respectively, demonstrates that the charge-pair interaction between yCc Lys 86 and Glu 235 of cyt c1 in the crystal structure is also important for the reaction in solution. Mutation of other yCc lysines to alanine, including 11, 72, 73, 79, and 87, also led to significant decreases in the rate constant (Table 5). Although there is only one electrostatic charge-pair interaction in the binding domain of the yCc:ybc1 crystallographic complex, the 5 lysine amino groups on yCc and 5 carboxylate groups on cyt bc1 immediately surrounding the interaction domain could guide Cc to the binding site and contribute to complex formation [99,100].
Table 5.
Reaction of yeast Ru-39-Cc mutants with yeast cyt bc1 [102]. The reaction solution contained 5 μM yeast Ru-39-Cc mutant and 5 μM yeast cyt bc1 in 5 mM sodium phosphate, pH 7.0 with 250 mM NaCl and 0.1% lauryl maltoside. The c1 and c hemes were reduced with 10 μM ascorbate. The sequence numbering of horse Cc is used.
| Mutant | k (s−1) |
|---|---|
| None | 3500 |
| K11A | 1480 |
| T12A | 2540 |
| R13A | 153 |
| V28A | 2620 |
| K72A | 1960 |
| K73A | 2190 |
| K79A | 1530 |
| A81G | 2400 |
| K86A | 1090 |
| K86D | 190 |
| K87A | 1120 |
Sarewicz et al. [103] carried out EPR studies indicating that the lifetime of the complex between Rhodobacter capsulatus cyt c2 and cyt bc1 was longer than 100 μs at low ionic strength, decreasing to less than 400 ns at ionic strengths above 125 mM. Their results are consistent with a mechanism in which cyt c2 binds rapidly to cyt bc1 at low ionic strength allowing efficient intracomplex electron transfer, but product complex dissociation is slow, limiting enzyme turnover. At high ionic strength complex formation is slow and complex dissociation is so rapid that most collisions do not result in electron transfer. At intermediate physiological ionic strength, complex formation and dissociation are both moderately rapid, and there is a moderate ratio of electron transfer per collision. The rapid kinetic studies discussed above [101] are also consistent with this mechanism. Sarewicz et al. [104] carried out EPR studies indicating that the dipole moment of cyt c2 plays an important role in orienting the molecule for efficient electron transfer during the collision process.
The Paracocus denitrificans cyt bc1 complex contains just three subunits, the b subunit with heme bL and heme bH, the Rieske iron-sulfur protein (ISP), and cyt c1 [105]. The cyt c1 subunit has a tripartite domain structure consisting of a unique N-terminal acidic domain of 150 amino acids, a periplasmically oriented core domain containing the covalently attached heme c, and a C-terminal membrane anchor. The acidic domain may be analogous to the small acidic subunits of eukaryotic cyt bc1, including the hinge protein of bovine cyt bc1 and subunit 6 of yeast cyt bc1 [106,107]. An usual feature of P. denitrificans cyt bc1 is that it has a “dimer of dimers” quaternary structure rather than the dimeric structure found in other cyt bc1 complexes [108]. Cyt bc1 transfers electrons to membrane-bound cyt c552 [109]. Kinetic studies have been carried out on genetically engineered soluble modules of both redox partners [110]. The soluble cytochrome c1 core fragment (cyt c1CF) consists of only the central core domain, without the acidic domain and the membrane anchor. The soluble cytochrome c552 fragment (cyt c552F) contains only the C-terminal hydrophilic heme domain without the N-terminal membrane anchor and linker region. A new ruthenium cyt c552F derivative (Ruz-23-c552F) was designed to measure rapid electron transfer with cyt c1CF using the ruthenium photooxidation technique [110]. The bimolecular rate constant k12 decreased rapidly with increasing ionic strength above 40 mM, indicating that electrostatic interactions were important for the reaction between the two proteins. However, k12 was rapid, 3 × 109 M−1s−1, and nearly independent of ionic strength below 35 mM. These results are consistent with a two-step process involving very rapid formation of an initial complex guided by long-range electrostatic interactions, followed by short-range diffusion along the protein surfaces guided by hydrophobic interactions. No intracomplex electron transfer between Ruz-23-c552F and c1CF was observed even at the lowest ionic strength, indicating a low-affinity complex. In contrast, yeast Ruz-39-Cc formed a tight 1:1 complex with cyt c1CF at ionic strengths below 60 mM with an intracomplex electron transfer rate constant of 50,000 s−1. A group of cyt c1CF mutants in the presumed docking site were generated based on information from the yeast cyt bc1/cyt c co-crystal structure. Kinetic analysis of cyt c1CF mutants located near the heme crevice provided preliminary identification of the interaction site for cyt c552F, and suggest that formation of the encounter complex is guided primarily by the overall electrostatic surface potential rather than by defined ion pairs [110]. Ruthenium kinetic studies have shown that the acidic domain does not play a significant role in the reaction of cyt c552F with P. denitrificans cyt bc1[111]. The reaction between a ruthenium horse Cc derivatve and R. sphaeroides cyt bc1 was found to have an intracomplex rate constant of 60,000 s−1 [112]. Mutagenesis studies indicated that acidic residues near the heme crevice of cyt c1 are important in guiding Cc to the binding domain [112].
Conclusions and Future Prospects
Extensive structural, spectroscopic, and kinetics studies have provided considerable insight into the mechanism of the cytochrome bc1 complex. Most of the rate constants for the key electron transfer reactions have been measured for R. sphaeroides cyt bc1 as indicated in Scheme 1 and Table 6. X-ray crystallographic and EPR studies have shown that inhibitors bound at the Qo site affect the position and mobility of the ISP [8,10,17, 25-28]. Kinetic studies have revealed that the reduction of cyt c1 by 2Fe2S is rate-limited by the rotational diffusion of the ISP from the b-state to the c1 state, and that the rate of this rotation is controlled by the type of inhibitor bound in the Qo site [13,21,61,67-68]. On the basis of these studies it has been proposed that binding QH2 to the Qo site induces a conformational change in the ef and cd1 loops leading to capture of the ISP in the b state and proton-coupled electron transfer from 2Fe2S to cyt c1[12,28]. Following electron transfer to cyt bL and cyt bH, the ISP is released from the b state, rotates to the c1 state, and transfers an electron to cyt c1. The detailed mechanism of how this bifurcated electron transfer reaction avoids short-circuit and bypass reactions remains enigmatic, however. A number of possibilities have been proposed, including control of the rotation of the ISP by the release of Q from the Qo site or the reduction of cyt bL and cyt bH [12,28], double gating mechanisms [69,75-79], or coulombic gating of the motion of the semiquinone [72,77]. It has also been proposed that the reaction of Q at the Qi site could affect the bifurcated reduction of QH2 at the Qo site and the motion of the ISP [26,27,84-89]. Moreover, experiments from several different laboratories indicate that electrons can be transferred between the two cyt bL hemes in the homodimeric enzyme, suggesting that electrons can be distributed between the four quinone sites in the dimer[85,89-92]. However, it is not clear whether cross-monomer communication regulates reactions at the Qo and Qi sites. Future experiments are needed to address these important questions.
Table 6.
Rate constants of reactions in R. sphaeroides cyt bc1. Rate constants for the reactions shown in Scheme 1 were measured under the conditions reported in the references.
Highlights.
Photoactive ruthenium polypyridine complexes have been developed to study rapid electron transfer reactions in the cytochrome bc1 complex at a microsecond time scale.
A ruthenium photooxidation technique has been developed to measure the rate constant for electron transfer from the iron-sulfur protein to cytochrome c1.
The dynamics of rotation of the iron-sulfur protein from the b state to the c1 state has been measured.
A conformational linkage between the occupant of the Qo ubiquinol binding site and the rotational dynamics of the iron-sulfur protein was investigated which could play a role in the bifurcated oxidation of ubiquinol.
A ruthenium photoexcitation method is described for the measurement of electron transfer from cytochrome c1 to cytochrome c.
Acknowledgements
This work was supported in part by NIH grants GM20488 and 8P30GM103450.
Abbreviations
- Cc
cytochrome c
- yCc
yeast Cc
- cyt bc1
cytochrome bc1
- CcO
cytochrome c oxidase
- 2Fe2S
Rieske iron-sulfur center
- ISP
iron-sulfur protein
- bpy
2,2′-bipyridine
- dmb
4,4′-dimethyl-2,2′-bipridine
- bpz
2,2′-bipyrazine
- bpd
3,3′-bipyridazine
- qpy
2,2′:4′,4″:2″,2″′-quaterpyridine
- Ru2D
[Ru(bpy)2]2qpy4+
- R.sphaeroides
Rhodobacter sphaeroides
- MOAS
methoxyacrylate stilbene
- JG144
S-3-anilino-5-methyl-5-(4,6-difluorophenyl)-1,3-oxazolidine-2,4-dione
- Q
ubiquionone
- QH2
ubiquinol
- Qo
outside ubiquinol binding site
- Qi
inside ubiquinone binding site
- Pf
Qo site inhibitor which fixes ISP in b state
- Pm
Qo site inhibitor which promotes mobile state of ISP
Footnotes
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
References
- 1.Trumpower BL, Gennis RB. Energy transduction by cytochrome complexes in mitochondrial and bacterial respiration: the enzymology of coupling electron transfer reactions to transmembrane proton translocation. Annu. Rev. Biochem. 1994;63:675–716. doi: 10.1146/annurev.bi.63.070194.003331. [DOI] [PubMed] [Google Scholar]
- 2.Trumpower BL. The protonmotive Q cycle. Energy transduction by coupling of proton translocation to electron transfer by the cytochrome bc1 complex. J. Biol. Chem. 1990;265:11409–11412. [PubMed] [Google Scholar]
- 3.Mitchell P. The protonmotive Q cycle: a general formulation. FEBS Lett. 1975;59:137–139. doi: 10.1016/0014-5793(75)80359-0. [DOI] [PubMed] [Google Scholar]
- 4.Crofts AR. The cytochrome bc1 complex: function in the context of structure. Ann. Rev. Physiol. 2004;66:689–733. doi: 10.1146/annurev.physiol.66.032102.150251. [DOI] [PubMed] [Google Scholar]
- 5.Xia D, Yu CA, Kim H, Xia JZ, Kachurin AM, Zhang L, Yu L, Deisenhofer J. Crystal structure of the cytochrome bc1 complex from bovine heart mitochondria. Science. 1997;277:60–66. doi: 10.1126/science.277.5322.60. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Zhang Z, Huang L, Shulmeister VM, Chi YI, Kim KK, Hung LW, Crofts AR, Berry EA, Kim SH. Electron transfer by domain movement in cytochrome .bc1. Nature. 1998;392:677–684. doi: 10.1038/33612. [DOI] [PubMed] [Google Scholar]
- 7.Iwata S, Lee JW, Okada K, Lee JK, Iwata M, Rasmussen B, Link TA, Ramaswamy S, Jap BK. Complete structure of the 11-subunit bovine mitochondrial cytochrome bc1 complex. Science. 1998;281:64–71. doi: 10.1126/science.281.5373.64. [DOI] [PubMed] [Google Scholar]
- 8.Kim H, Xia D, Yu CA, Xia JZ, Kachurin AM, Zhang L, Yu L, Deisenhofer J. Inhibitor binding changes domain mobility in the iron-sulfur protein of the mitochondrial bc1 complex from bovine heart. Proc. Natl. Acad. Sci. USA. 1998;95:8026–8033. doi: 10.1073/pnas.95.14.8026. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Hunte C, Koepke J, Lange C, Rossmanith T, Michel H. Structure at 2.3 Å resolution of the cytochrome bc1 complex from the yeast Saccharomyces cerevisiae co-crystallized with an antibody Fv fragment. Structure Fold. Des. 2000;8:669–684. doi: 10.1016/s0969-2126(00)00152-0. [DOI] [PubMed] [Google Scholar]
- 10.Esser L, Elberry M, Zhou F, Yu CA, Yu L, Xia D. Inhibitor-complexed structures of the cytochrome bc1 from the photosynthetic bacterium Rhodobactersphaeroides. J. Biol. Chem. 2008;283:2846–2857. doi: 10.1074/jbc.M708608200. [DOI] [PubMed] [Google Scholar]
- 11.Berry EA, Huang LS, Saechao LK, Pon NG, Valkova-Valchanova M, Daldal F. X-Ray Structure of Rhodobacter Capsulatus Cytochrome bc1: Comparison with its Mitochondrial and Chloroplast Counterparts. Photosynth. Res. 2004;81:251–75. doi: 10.1023/B:PRES.0000036888.18223.0e. [DOI] [PubMed] [Google Scholar]
- 12.Xia D, Esser L, Yu L, Yu CA. Structural basis for the mechanism of electron bifurcation at the quinol oxidation site of the cytochrome bc1 complex. Photosyn. Res. 2007;92:17–34. doi: 10.1007/s11120-007-9155-3. [DOI] [PubMed] [Google Scholar]
- 13.Sadoski RC, Engstrom G, Tian H, Zhang L, Yu CA, Yu L, Durham B, Millett F. Use of a photoactivated ruthenium dimer complex to measure electron transfer between the Rieske iron-sulfur protein and cytochrome c1 in the cytochrome bc1 complex. Biochemistry. 2000;39:4231–4236. doi: 10.1021/bi000003o. [DOI] [PubMed] [Google Scholar]
- 14.Tian H, Yu L, Mather MW, Yu CA. Flexibility of the neck region of the Rieske iron-sulfur protein is functionally important in the cytochrome bc1 complex. J. Biol. Chem. 1998;273:27953–27959. doi: 10.1074/jbc.273.43.27953. [DOI] [PubMed] [Google Scholar]
- 15.Xiao K, Yu L, Yu CA. Confirmation of the involvement of protein domain movement during the catalytic cycle of the cytochrome bc1 complex by the formation of an intersubunit disulfide bond between cytochrome b and the iron-sulfur protein. J. Biol. Chem. 2000;275:38597–38604. doi: 10.1074/jbc.M007444200. [DOI] [PubMed] [Google Scholar]
- 16.Tian H, White S, Yu L, Yu CA. Evidence for the head domain movement of the Rieske iron-sulfur protein in electron transfer reaction of the cytochrome bc1 complex. J. Biol. Chem. 1999;274:7146–7152. doi: 10.1074/jbc.274.11.7146. [DOI] [PubMed] [Google Scholar]
- 17.Darrouzet E, Valkova-Valchanova M, Daldal F. Probing the role of the Fe-S subunit hinge region during Qo site catalysis in Rhodobacter capsulatusbc1 complex. Biochemistry. 2000;39:15475–15483. doi: 10.1021/bi000750l. [DOI] [PubMed] [Google Scholar]
- 18.Darrouzet E, Valkova-Valchanova M, Moser CC, Dutton PL, Daldal F. Uncovering the 2Fe2S domain movement in cytochrome bc1 and its implications for energy conversion. Proc. Natl. Acad. Sci. USA. 2000;97:4567–4572. doi: 10.1073/pnas.97.9.4567. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Nett JH, Hunte C, Trumpower BL. Changes to the length of the flexible linker region of the Rieske protein impair the interaction of ubiquinol with the cytochrome bc1 complex. Eur. J. Biochem. 2000;267:5777–5782. doi: 10.1046/j.1432-1327.2000.01650.x. [DOI] [PubMed] [Google Scholar]
- 20.Ghosh M, Wang Y, Ebert CE, Vadlamuri S, Beattie DS. Substituting leucine for alanine-86 in the tether region of the iron-sulfur protein of the cytochrome bc1 complex affects the mobility of the 2Fe2S domain. Biochemistry. 2001;40:327–335. doi: 10.1021/bi001708t. [DOI] [PubMed] [Google Scholar]
- 21.Engstrom G, Xiao K, Yu CA, Yu L, Durham B, Millett F. Photoinduced electron transfer between the Rieske iron-sulfur protein and cytochrome c1 in the Rhodobacter sphaeroides cytochrome bc1 complex. Effects of pH, temperature, and driving force. J. Biol. Chem. 2002;277:31072–31078. doi: 10.1074/jbc.M202594200. [DOI] [PubMed] [Google Scholar]
- 22.Darrouzet E, Valkova-Valchanova M, Daldal F. The [2Fe2S] cluster Em as an indicator of the iron-sulfur subunit position in the ubihydroquinone oxidation site of the cytochrome bc1 complex. J. Biol. Chem. 2002;277:3464–3470. doi: 10.1074/jbc.M107973200. F. [DOI] [PubMed] [Google Scholar]
- 23.Darrouzet E, Daldal F. Movement of the iron-sulfur subunit beyond the ef loop of cytochrome b is required for multiple turnovers of the bc1 complex but not for single turnover Qo site catalysis. J. Biol. Chem. 2002;277:3471–3476. doi: 10.1074/jbc.M107974200. [DOI] [PubMed] [Google Scholar]
- 24.Brugna M, Rodgers S, Schricker A, Montoya G, Kazmeier M, Nitschke W, Sinning I. A spectroscopic method for observing the domain movement of the Rieske iron-sulfur protein. Proc. Natl. Acad. Sci. USA. 2000;97:2069–2074. doi: 10.1073/pnas.030539897. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Berry EA, Huang LS. Conformationally linked interaction in the cytochrome bc1 complex between inhibitors of the Qo site and the Rieske iron-sulfur protein. Biochim Biophys Acta. 2011;1807:1349–63. doi: 10.1016/j.bbabio.2011.04.005. [DOI] [PubMed] [Google Scholar]
- 26.Cooley JW, Roberts AG, Bowman MK, Kramer DM, Daldal F. The raised midpoint potential of the 2Fe2S cluster of cytochrome bc1 is mediated by both the Qo site occupants and the head domain position of the Fe-S protein subunit. Biochemistry. 2004;43:2217–27. doi: 10.1021/bi035938u. [DOI] [PubMed] [Google Scholar]
- 27.Sarewicz M, Dutka M, Froncisz W, Osyczk A. Magnetic interactions sense changes in distance between heme bL and the iron-sulfur cluster in cytochrome bc1. Biochemistry. 2009;48:5708–20. doi: 10.1021/bi900511b. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Esser L, Gong X, Yang S, Yu L, Yu CA, Xia D. Surface-modulated motion switch: capture and release of iron-sulfur protein in the cytochrome bc1 complex. Proc. Natl. Acad. Sci. USA. 2006;103:13045–13050. doi: 10.1073/pnas.0601149103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Kalyanasundaram K. Photochemistry of Polypyridine and Porphyrin Complexes. Academic Press; London: 1992. [Google Scholar]
- 30.Pan LP, Durham B, Wolinska J, Millett F. Preparation and Characterization of Singly Labeled Ruthenium Polypyridine Cytochrome c Derivatives. Biochemistry. 1988;27:7180–7184. doi: 10.1021/bi00419a003. [DOI] [PubMed] [Google Scholar]
- 31.Durham B, Pan LP, Long J, Millett F. Photoinduced Electron-Transfer Kinetics of Singly Labeled Ruthenium Bis-(bipyridine) Dicarboxybipyridine Cytochrome c Derivatives. Biochemistry. 1989;28:8659–8665. doi: 10.1021/bi00447a057. [DOI] [PubMed] [Google Scholar]
- 32.Liu R-Q, Geren L, Anderson P, Fairris JL, Peffer N, McKee A, Durham B, Millett F. Design of Ruthenium-Cytochrome c Derivatives to Measure Electron Transfer to Cytochrome c Peroxidase. Biochimie. 1995;77:549–561. doi: 10.1016/0300-9084(96)88171-5. [DOI] [PubMed] [Google Scholar]
- 33.Durham B, Pan LP, Hahm S, Long J, Millett F. Electron Transfer in Biology and the Solid State. In: Johnson MK, et al., editors. ACS Advances in Chemistry Series. American Chemical Society; Washington, D. C.: 1990. pp. 181–193. [Google Scholar]
- 34.Geren L, Hahm S, Durham B, Millett F. Photoinduced Electron Transfer Between Cytochrome c Peroxidase and Yeast Cytochrome c Labeled at Cys 102 with (4-Bromomethyl-4′-methylbipyridine)[bis(bipyridine)]ruthenium. Biochemistry. 1991;30:9450–9457. doi: 10.1021/bi00103a009. [DOI] [PubMed] [Google Scholar]
- 35.Willie A, Stayton PS, Sligar SG, Durham B, Millett F. Genetic Engineering of Redox Donor Sites: Measurement of Intracomplex, Electron Transfer between Ruthenium-65-Cytochrome b5 and Cytochrome c. Biochemistry. 1992;31:7237–7242. doi: 10.1021/bi00147a005. [DOI] [PubMed] [Google Scholar]
- 36.Geren LM, Beasley JR, Fines BR, Saunders AJ, Hibdon S, Pielak GJ, Durham B, Millett F. Design of a Ruthenium-Cytochrome c Derivative to Measure the Electron Transfer to the Initial Acceptor in Cytochrome c Oxidase. J. Biol. Chem. 1995;270:2466–2472. doi: 10.1074/jbc.270.6.2466. [DOI] [PubMed] [Google Scholar]
- 37.Wang K, Mei H, Geren L, Miller MA, Saunders A, Wang X, Waldner JL, Pielak GJ, Durham B, Millett F. Design of a ruthenium-cytochrome c derivative to measure electron transfer to the radical cation and oxyferryl heme in cytochrome c peroxidase. Biochemistry. 1996;35:15107–15119. doi: 10.1021/bi9611117. [DOI] [PubMed] [Google Scholar]
- 38.Zaslavsky D, Sadoski RC, Wang K, Durham B, Gennis RB, Millett F. Single Electron Reduction of Cytochrome c Oxidase Compound F: Resolution of Partial Steps by Transient Spectroscopy. Biochemistry. 1998;37:14910–14916. doi: 10.1021/bi981490z. [DOI] [PubMed] [Google Scholar]
- 39.Brand SE, Rajagukguk S, Ganesan K, Geren L, Fabian M, Han D, Gennis RB, Durham B, Millett F. A new ruthenium complex to study single-electron reduction of the pulsed OH state of detergent-solubilized cytochrome oxidase. Biochemistry. 2007;46:14610–14618. doi: 10.1021/bi701424d. [DOI] [PubMed] [Google Scholar]
- 40.Winkler JR, Nocera DG, Yocom KM, Bordignon E, Gray HB. Electron-Transfer Kinetics of Ru(NH3)5(III)(His-33)-Ferricytochrome c. J. Am. Chem. Soc. 1982;104:5798–5800. [Google Scholar]
- 41.Isied SS, Worosila G, Atherton SJ. Electron Transfer across Polypeptides. Intramolecular Electron Transfer from Ruthenium (II) to Iron (III) in Histidine-33 Modified Horse Heart Cytochrome c. J. Am. Chem. Soc. 1982;104:7659–7661. [Google Scholar]
- 42.Scott JR, Fairris JL, McLean M, Wang K, Sligar SG, Durham B, Millett F. Intramolecular Electron-Transfer Reactions of Cytochrome b5 Covalently Bonded to Ruthenium(II) Polypyridine Complexes: Reorganizational Energy and Pressure Effects. Inorg. Chim. Acta. 1996;243:193. [Google Scholar]
- 43.Fairris JL, Wang K, Geren L, Pielak GJ, Durham B, Millett F. Intramolecular Electron Transfer in Yeast Cytochrome c Covalently Bonded to Ruthenium(II) Polypyridine Complexes at Cys39. Photochemistry and Radiation Chemistry, ACS; Washington, DC: 1998. pp. 99–110. (ACS Advances in Chemistry Series 254). [Google Scholar]
- 44.Millett F, Durham B. Design of Photoactive Ruthenium Complexes to Study Interprotein Electron Transfer. Biochemistry. 2002;41:11315–11324. doi: 10.1021/bi0262956. [DOI] [PubMed] [Google Scholar]
- 45.Meade TJ, Gray HB, Winkler JR. Driving-Force Effects on the Rate of Long-Range Electron Transfer in Ruthenium-Modified Cytochrome c. J. Am. Chem. Soc. 1989;111:4353–4356. [Google Scholar]
- 46.Chang I-J, Gray HB, Winkler JR. High-Driving Force Electron Transfer in Metalloproteins: Intramolecular Oxidation of Ferrocytochrome c by Ru(2,2′bpy)2(im)(His-33)3+ J. Am. Chem. Soc. 1991;113:7056–757. [Google Scholar]
- 47.Wuttke DS, Bjerrum MJ, Winkler JR, Gray HB. Electron-Tunneling Pathways in Cytochrome c. Science. 1992;256:1007–1009. doi: 10.1126/science.256.5059.1007. [DOI] [PubMed] [Google Scholar]
- 48.Di Bilio AJ, Dennison C, Gray HB, Ramirez BE, Sykes AG, Winkler JR. Electron Transfer in Ruthenium-Modified Plastocyanin. J. Am. Chem. Soc. 1998;120:7551–7556. [Google Scholar]
- 49.Bjerrum MJ, Casimiro DR, Chang I, De Bilio AJ, Gray HB, Hill MG, Langen R, Mines GA, Skov LK, Winkler JR, Wuttke DS. Electron Transfer in Ruthenium-Modified Proteins. J. Bioen. Biomem. 1995;27:295–302. doi: 10.1007/BF02110099. [DOI] [PubMed] [Google Scholar]
- 50.Marcus RA. On the Theory of Oxidation-Reduction Reactions Involving Electron Transfer. J. Chem. Phys. 1956;24:966–978. [Google Scholar]
- 51.Moser CC, Keske JM, Warncke K, Farid RS, Dutton PL. Nature of biological electron transfer. Nature. 1992;355:796–802. doi: 10.1038/355796a0. [DOI] [PubMed] [Google Scholar]
- 52.Yu CA, Yu L, King TE. Kinetics of electron transfer between cardiac cytochrome c1 and c. J. Biol. Chem. 1973;248:528–33. [PubMed] [Google Scholar]
- 53.De Vries S, Albracht SPJ, Berden JA, Slater EC. The pathway of electron through QH2:Cytochrome c Oxidoreductase studied by pre-steady-state kinetics. Biochim. Biophys. Acta. 1982;681:41–53. doi: 10.1016/0005-2728(82)90276-6. [DOI] [PubMed] [Google Scholar]
- 54.Hansen KC, Schultz BE, Wang F, Chan SI. Reaction of E. coli cytochrome bo3 and mitochondrial cytochrome bc1 with a photoreleasable decylubiquinol. Biochim. Biophy. Acta. 2000;1456:121–137. doi: 10.1016/s0005-2728(99)00107-3. [DOI] [PubMed] [Google Scholar]
- 55.Snyder CH, Guierrez-Cirlos EB, Trumpower BL. Evidence for a concerted mechanism of ubiquinol oxidation by the cytochrome bc1 complex. J. Biol. Chem. 2000;275:13555–13541. doi: 10.1074/jbc.275.18.13535. [DOI] [PubMed] [Google Scholar]
- 56.Trumpower BL. A concerted, alternating sites mechanism of ubiquinol oxidation by the dimeric cytochrome bc1 complex. Biochim. Biophys. Acta. 2002;1555:166–173. doi: 10.1016/s0005-2728(02)00273-6. [DOI] [PubMed] [Google Scholar]
- 57.Crofts AR, Meinhardt SW. A Q-cycle mechanism for the cyclic electron transfer chain of Rp. Sphaeroides. Biochem. Soc. Trans. 1982;10:210–203. doi: 10.1042/bst0100201. [DOI] [PubMed] [Google Scholar]
- 58.Meinhardt SW, Crofts AR. Kinetic and thermodynamic resolution of cytochrome c1 and cytchrome c2 from Rhodopseudomonas sphaeroides. FEBS let. 1982;149:223–227. [Google Scholar]
- 59.Crofts AR, Wang Z. How rapid are the internal reactions of the ubiquinol:cytochrome c2 oxidoreductase? Photosynthesis Res. 1989;22:69–87. doi: 10.1007/BF00114768. [DOI] [PubMed] [Google Scholar]
- 60.Shinkarev VP, Crofts AR, Wraight CA. The electric field generated by photosynthetic reaction center induces rapid reversed electron transfer in the bc1 complex. Biochemistry. 2001;40:12584–12590. doi: 10.1021/bi011334j. [DOI] [PubMed] [Google Scholar]
- 61.Havens J, Castellani M, Kleinschroth T, Ludwig B, Durham B, Millett F. Photoinitiated electron transfer within the Paracoccus denitrificans cytochrome bc1 complex: mobility of the iron-sulfur protein is modulated by the occupant of the Qo site. Biochemistry. 2011;50:10462–72. doi: 10.1021/bi200453r. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Link TA. Two pK values of the oxidized Rieske [2Fe2S] cluster observed by CD spectroscopy. Biochim. Biophys. Acta. 1994;1185:81–84. [Google Scholar]
- 63.Ugulava NB, Crofts AR. CD-monitored redox titration of the Rieske Fe-S protein of Rb. sphaeroides: pH dependence of the midpoint potential in isolated bc1 complex and in membranes. FEBS Lett. 1998;440:409–413. doi: 10.1016/s0014-5793(98)01493-8. [DOI] [PubMed] [Google Scholar]
- 64.Ritter M, Palsdottir H, Abe M, Mäntele W, Hunte C, Miyoshi H, Hellwig P. Direct evidence for the interaction of stigmatellin with a protonated acidic group in the bc1 complex from Saccharomyces cerevisiae as monitored by FTIR difference spectroscopy and 13C specific labeling. Biochemistry. 2004;43:8439–8446. doi: 10.1021/bi049649x. [DOI] [PubMed] [Google Scholar]
- 65.Esser L, Quinn B, Li YF, Zhang M, Elberry M, Yu L, Yu CA, Xia D. Crystallographic studies of quinol oxidation site inhibitors: a modified classification of inhibitors for the cytochrome bc1 complex. J. Mol. Biol. 2004;341:281–302. doi: 10.1016/j.jmb.2004.05.065. L. [DOI] [PubMed] [Google Scholar]
- 66.Lancaster CR, Hunte C, Kelley J, Trumpower BL, Ditchfield R. A comparison of stigmatellin conformations, free and bound to the photosynthetic reaction center and the cytochrome bc1 complex. J. Mol. Biol. 2007;368:197–208. doi: 10.1016/j.jmb.2007.02.013. [DOI] [PubMed] [Google Scholar]
- 67.Xiao K, Engstrom G, Rajagukguk S, Yu CA, Yu L, Durham B, Millett F. Effect of famoxadone on photoinduced electron transfer between the iron-sulfur center and cytochrome c1 in the cytochrome bc1 complex. J. Biol. Chem. 2003;278:11419–11426. doi: 10.1074/jbc.M211620200. [DOI] [PubMed] [Google Scholar]
- 68.Rajagukguk S, Yang S, Yu CA, Yu L, Durham B, Millett F. Effect of mutations in the cytochrome b ef loop on the electron-transfer reactions of the Rieske iron-sulfur protein in the cytochrome bc1 complex. Biochemistry. 2007;46:1791–1798. doi: 10.1021/bi062094g. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Osyczka A, Moser CC, Daldal F, Dutton PL. Reversible redox energy coupling in electron transfer chains. Nature. 2004;427:607–612. doi: 10.1038/nature02242. [DOI] [PubMed] [Google Scholar]
- 70.Chobot SE, Zhang H, Moser CC, Dutton PL. Breaking the Q-cycle: finding new ways to study Qo through thermodynamic manipulations. J. Bioenerg. Biomembr. 2008;40:501–507. doi: 10.1007/s10863-008-9175-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Hong S, Ugulava N, Guergova-Kuras M, Crofts AR. The energy landscape for ubihydroquinone oxidation at the Qo site of the bc1 complex in Rhodobacter sphaeroides. J. Biol. Chem. 1999;274:33931–33944. doi: 10.1074/jbc.274.48.33931. [DOI] [PubMed] [Google Scholar]
- 72.Crofts AR, Hong S, Zhangand Z, Berry EA. Physicochemical aspects of the movement of the Rieske iron sulfur protein during quinol oxidation by the bc1 complex from mitochondria and photosynthetic bacteria. Biochemistry. 1999;38:15827–15839. doi: 10.1021/bi990963e. [DOI] [PubMed] [Google Scholar]
- 73.Lhee S, Kolling D, Nair S, Dikanov S, Crofts AR. Modifications of protein environment of the [2Fe2S] cluster of the bc1 complex: Effects on the biophysical properties of the Rieske iron-sulfur protein and on the kinetics of the complex. J. Biol. Chem. 2010;285:9233–9248. doi: 10.1074/jbc.M109.043505. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Darrouzet E, Daldal F. Protein-protein interactions between cytochrome b and the Fe-S protein subunits during QH2 oxidation and large-scale domain movement in the bc1 complex. Biochemistry. 2003;42:1499–1507. doi: 10.1021/bi026656h. [DOI] [PubMed] [Google Scholar]
- 75.Rich PR. The quinone chemistry of bc complexes. Biochim. Biophys. Acta. 2004;1658:165–171. doi: 10.1016/j.bbabio.2004.04.021. [DOI] [PubMed] [Google Scholar]
- 76.Osyczka A, Moser CC, Dutton PL. Fixing the Q cycle. Trends Biochem. Sci. 2005;30:176–182. doi: 10.1016/j.tibs.2005.02.001. [DOI] [PubMed] [Google Scholar]
- 77.Crofts AR, Lhee S, Crofts SB, Cheng J, Rose S. Proton pumping in the bc1 complex: a new gating mechanism that prevents short circuits. Biochim. Biophys. Acta. 2006;1757:1019–1034. doi: 10.1016/j.bbabio.2006.02.009. [DOI] [PubMed] [Google Scholar]
- 78.Mulkidjanian AY. Ubiquinol oxidation in the cytochrome bc1 complex: reaction mechanism and prevention of short-circuiting. Biochim. Biophys. Acta. 2005;1709:5–34. doi: 10.1016/j.bbabio.2005.03.009. (2005) [DOI] [PubMed] [Google Scholar]
- 79.Osyczka A, Zhang H, Mathé C, Rich PR, Moser CC, Dutton PL. Role of the PEWY glutamate in hydroquinone-quinone oxidation-reduction catalysis in the Qo Site of cytochrome bc1. Biochemistry. 2006;45:10492–10503. doi: 10.1021/bi060013a. [DOI] [PubMed] [Google Scholar]
- 80.Zhu J, Egawa T, Yeh SR, Yu L, Yu CA. Simultaneous reduction of iron-sulfur protein and cytochrome bL during ubiquinol oxidation in cytochrome bc1 complex. Proc. Natl. Acad. Sci. USA. 2007;104:4864–4869. doi: 10.1073/pnas.0607812104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Cape JL, Bowman MK, Kramer DM. A semiquinone intermediate generated at the Qo site of the cytochrome bc1 complex: importance for the Q-cycle and superoxide production. Proc. Natl. Acad. Sc.i USA. 2007;104:7887–7892. doi: 10.1073/pnas.0702621104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Zhang H, Osyczka A, Dutton PL, Moser CC. Exposing the complex III Qo semiquinone radical. Biochim. Biophys. Acta. 2007;1767:883–887. doi: 10.1016/j.bbabio.2007.04.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Forquer I, Covian R, Bowman MK, Trumpower BL, Kramer DM. Similar transition states mediate the Q-cycle and superoxide production by the cytochrome bc1 complex. J. Biol. Chem. 2006;281:38459–38465. doi: 10.1074/jbc.M605119200. [DOI] [PubMed] [Google Scholar]
- 84.Covian R, Trumpower BL. Regulatory interactions in the dimeric cytochrome bc1 complex: the advantages of being a twin. Biochim. Biophys. Acta. 2008;1777:1079–1091. doi: 10.1016/j.bbabio.2008.04.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Castellani M, Covian R, Kleinschroth T, Anderka O, Ludwig B, Trumpower B. Direct Demonstration of Half-of-the-sites Reactivity in the Dimeric Cytochrome bc1 Complex: Enzyme with one inactive monomer is fully active but unable to activate the second ubiquinol oxidation site in response to ligand binding at the ubiquinone reduction site. J. Biol. Chem. 2010;285:502–510. doi: 10.1074/jbc.M109.072959. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Cooley J. A structural model for across membrane coupling between the Qo and Qi active sites of cytochrome bc1. Biochim. Biophys. Acta. 2010;1797:1842–1848. doi: 10.1016/j.bbabio.2010.05.013. [DOI] [PubMed] [Google Scholar]
- 87.Mulkidjanian AY. Proton translocation by the cytochrome bc1 complexes of phototrophic bacteria: introducing the activated Q-cycle. Photochem. Photobio. Sci. 2007;6:19–34. doi: 10.1039/b517522d. [DOI] [PubMed] [Google Scholar]
- 88.Cen X, Yu L, Yu CA. Domain movement of iron sulfur protein in cytochrome bc1 complex is facilitated by the electron transfer from cytochrome bL to bH. FEBS Lett. 2008;582:523–526. doi: 10.1016/j.febslet.2008.01.016. [DOI] [PubMed] [Google Scholar]
- 89.Lanciano P, Lee DW, Yang H, Darrouzet E, Daldal F. Intermonomer electron transfer between the low-potential b hemes of cytochrome bc1. Biochemistry. 2011;50:1651–63. doi: 10.1021/bi101736v. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Świerczek M, Cieluch E, Sarewicz M, Borek A, Moser CC, Dutton PL, Osyczka A. An electronic bus bar lies in the core of cytochrome bc1. Science. 2010;329:451–454. doi: 10.1126/science.1190899. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Czapla M, Borek A, Sarewicz M, Osyczka A. Enzymatic activities of isolated cytochrome bc1-like complexes containing fused cytochrome b subunits with asymmetrically inactivated segments of electron transfer chains. Biochemistry. 2012;51:829–35. doi: 10.1021/bi2016316. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Hong S, Victoria D, Crofts AR. Inter-monomer electron transfer is too slow to compete with monomeric turnover in bc1 complex. Biochim. Biophys. Acta. 2012;1817:1053–62. doi: 10.1016/j.bbabio.2012.03.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Ahmed AJ, Smith HT, Smith MB, Millett F. Effect of specific lysine modification on the reduction of cytochrome c by succinate-cytochrome c reductase. Biochemistry. 1978;17:2479–2483. doi: 10.1021/bi00606a003. [DOI] [PubMed] [Google Scholar]
- 94.Smith H, Ahmed A, Millett F. Electrostatic Interaction of Cytochrome c with Cytochrome c1 and Cytochrome Oxidase. J. Biol. Chem. 1981;256:4984–4990. [PubMed] [Google Scholar]
- 95.Speck SH, Ferguson-Miller S, Osheroff N, Margoliash E. Definition of cytochrome c binding domains by chemical modification: Kinetics of reaction with beef mitochondrial reductase. Proc. Nat. Acad. Sc. USA. 1979;76:155–160. doi: 10.1073/pnas.76.1.155. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Konig BW, Osheroff N, Wilms J, Muijsers AO, Dekker HL, Margoliash E. Mapping of the Interaction Domain for Purified Cytochrome c1 on Cytochrome c. FEBS Lett. 1980;111:359–398. doi: 10.1016/0014-5793(80)80835-0. [DOI] [PubMed] [Google Scholar]
- 97.Stonehuerner J, O’Brien P, Geren L, Millett F, Steidl J, Yu L, Yu C-A. Identification of the Binding site on Cytochrome c1 for Cytochrome c. J. Biol. Chem. 1983;260:5392–5398. [PubMed] [Google Scholar]
- 98.Gutweniger HE, Grassi C, Bisson C. Interaction between cytochrome c and ubiquinone-cytochrome c oxidoreductase: a study with water-soluble carbodiimides. Biochem. Biophys. Res. Comm. 1983;116:272–283. doi: 10.1016/0006-291x(83)90411-4. [DOI] [PubMed] [Google Scholar]
- 99.Lange C, Hunte C. Crystal structure of the yeast cytochrome bc1 complex with its bound substrate cytochrome c. Proc. Nat. Acad. Sci. USA. 2002;99:2800–2805. doi: 10.1073/pnas.052704699. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Solmaz SRN, Hunte C. Structure of Complex III with Bound Cytochrome c in Reduced State and Definition of a Minimal Core Interface for Electron Transfer. J. Biol. Chem. 2008;283:17542–17549. doi: 10.1074/jbc.M710126200. [DOI] [PubMed] [Google Scholar]
- 101.Engstrom G, Rajagukguk R, Saunders AJ, Patel C, Rajagukguk S, Merbitz-Zahradnik T, Xiao K, Pielak G, Trumpower B, Yu C-A, Yu L, Durham B, Millett F. Design of a Ruthenium-labeled Cytochrome c Derivative to Study Electron Transfer with the Cytochrome bc1 Complex. Biochemistry. 2003;42:2816–2824. doi: 10.1021/bi027213g. [DOI] [PubMed] [Google Scholar]
- 102.Rajagukguk S. Ph.D. Dissertation. University of Arkansas; 2006. Synthesis and use of Ruthenium Complexes in the Study of Electron Transfer Reactions within the Cytochrome bc1 Complex and with its Redox Partner Cytochrome c. [Google Scholar]
- 103.Sarewicz M, Bork A, Daldal F, Froncisz W, Osyczka A. Demonstration of short-lived complexes of cytochrome c with cytochrome bc1 by EPR Spectroscopy. J. Biol. Chem. 2008;283:24826–836. doi: 10.1074/jbc.M802174200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Sarewicz M, Pietras R, Froncisz W, Osyczka A. Reorientation of cytochrome c2 upon interaction with oppositely charged macromolecules probed by SR EPR: implications for the role of dipole moment to facilitate collisions in proper configuration for electron transfer. Metallomics. 2011;3:404–409. doi: 10.1039/c0mt00105h. [DOI] [PubMed] [Google Scholar]
- 105.Kurowski B, Ludwig B. The genes of the Paracoccus denitrificansbc1 complex. Nucleotide sequence and homologies between bacterial and mitochondrial subunits. J. Biol. Chem. 1987;262:13805–13811. [PubMed] [Google Scholar]
- 106.Kim CH, Balny C, King TE. Role of the hinge protein in the electron transfer between cardiac cytochrome c1 and c. Equilibrium constants and kinetic probes. J. Biol. Chem. 1987;262:8103–8108. [PubMed] [Google Scholar]
- 107.Schmitt ME, Trumpower BL. Subunit 6 regulates half-of-the-sites reactivity of the dimeric cytochrome bc1 complex in Saccharomyces cerevisiae. J. Biol. Chem. 1990;265:17005–17011. [PubMed] [Google Scholar]
- 108.Morgner N, Kleinschroth T, Barth HD, Ludwig B, Brutschy B. A novel approach to analyze membrane proteins by laser mass spectrometry: from protein subunits to the integral complex. J. Am. Soc. Mass Spectr. 2007;18:1429–1438. doi: 10.1016/j.jasms.2007.04.013. [DOI] [PubMed] [Google Scholar]
- 109.Turba A, Jetzek M, Ludwig B. Purification of Paracoccus denitrificans cytochrome c552 and sequence analysis of the gene. Eur. J. Biochem. 1995;231:259–65. [PubMed] [Google Scholar]
- 110.Janzon J, Yuan Q, Malatesta F, Hellwig P, Ludwig B, Durham B, Millett F. Probing the Paracoccus denitrificans cytochrome c1-cytochrome c552 interaction by mutagenesis and fast kinetics. Biochemistry. 2008;47:12974–84. doi: 10.1021/bi800932c. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Castellani M, Havens J, Kleinschroth T, Millett F, Durham B, Malatesta F, Ludwig B. The acidic domain of cytochrome c1 in paracoccus denitrificans, analogous to the acidic subunits in eukaryotic bc1 complexes, is not involved in the electron transfer reaction to its native substrate cytochrome c552. Biochim. Biophys. Acta. 2011;1807:1383–9. doi: 10.1016/j.bbabio.2011.08.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Tian H, Sadoski R, Zhang L, Yu C.-a., Yu L, Durham B, Millett F. Definition of the interaction domain for cytochrome c on the cytochrome bc1 complex. Steady-state and rapid kinetic analysis of electron transfer between cytochrome c and Rhodobacter sphaeroides cytochrome bc1 surface mutants. J. Biol. Chem. 2000;275:9587–9595. doi: 10.1074/jbc.275.13.9587. [DOI] [PubMed] [Google Scholar]










