Abstract
Like most bacteria, Escherichia coli has a flexible and branched respiratory chain that enables the prokaryote to live under a variety of environmental conditions, from highly aerobic to completely anaerobic. In general, the bacterial respiratory chain is composed of dehydrogenases, a quinone pool, and reductases. Substrate-specific dehydrogenases transfer reducing equivalents from various donor substrates (NADH, succinate, glycerophosphate, formate, hydrogen, pyruvate, and lactate) to a quinone pool (menaquinone, ubiquinone, and dimethylmenoquinone). Then electrons from reduced quinones (quinols) are transferred by terminal reductases to different electron acceptors. Under aerobic growth conditions, the terminal electron acceptor is molecular oxygen. A transfer of electrons from quinol to O2 is served by two major oxidoreductases (oxidases), cytochrome bo3 encoded by cyoABCDE and cytochrome bd encoded by cydABX. Terminal oxidases of aerobic respiratory chains of bacteria, which use O2 as the final electron acceptor, can oxidize one of two alternative electron donors, either cytochrome c or quinol. This review compares the effects of different inhibitors on the respiratory activities of cytochrome bo3 and cytochrome bd in E. coli. It also presents a discussion on the genetics and the prosthetic groups of cytochrome bo3 and cytochrome bd. The E. coli membrane contains three types of quinones that all have an octaprenyl side chain (C40). It has been proposed that the bo3 oxidase can have two ubiquinone-binding sites with different affinities.
“What’s new” in the revised article: The revised article comprises additional information about subunit composition of cytochrome bd and its role in bacterial resistance to nitrosative and oxidative stresses. Also, we present the novel data on the electrogenic function of appBCX-encoded cytochrome bd-II, a second bd-type oxidase that had been thought not to contribute to generation of a proton motive force in E. coli, although its spectral properties closely resemble those of cydABX-encoded cytochrome bd.
TWO TYPES OF METABOLISM, TWO TYPES OF OXIDASES
Anaerobiosis versus Aerobiosis in Escherichia coli
Like most bacteria, Escherichia coli has a flexible and branched respiratory chain that enables the prokaryote to live under a variety of environmental conditions, from highly aerobic to completely anaerobic. E. coli induces the expression of those respiratory components that are best suited to a particular environment. In general, the bacterial respiratory chain is composed of dehydrogenases, a quinone pool, and reductases. Substrate-specific dehydrogenases transfer reducing equivalents from various donor substrates (NADH, succinate, α-glycerophosphate, formate, hydrogen, pyruvate, and lactate) to a quinone pool (menaquinone, ubiquinone, and dimethylmenoquinone). Then electrons from reduced quinones (quinols) are transferred by terminal reductases to different electron acceptors. Under aerobic growth conditions, the terminal electron acceptor is molecular oxygen. A transfer of electrons from quinol to O2 is served by two major oxidoreductases (oxidases), cytochrome bo3 and cytochrome bd (it is worth noting that accumulated evidence over the past few years also suggests the contribution of a second bd-type oxidase, cytochrome bd-II, to the electron transfer and membrane potential generation). When oxygen is not available (under anaerobic conditions), alternative terminal electron acceptors, including nitrate, nitrite, dimethyl sulfoxide, trimethylamine N-oxide, and fumarate, can be used, and the reaction is catalyzed by nitrate reductases, nitrite reductase, dimethyl sulfoxide reductases, trimethylamine N-oxide reductase, and fumarate reductase, respectively (reviewed in references 1 and 2).
Two Types of Quinol Oxidases Only
Terminal oxidases of aerobic respiratory chains of bacteria, which use O2 as the final electron acceptor, can oxidize one of two alternative electron donors, either cytochrome c or quinol. Oxidases utilizing cytochrome c are called cytochrome c oxidases, whereas oxidases oxidizing quinol are referred to as quinol oxidases (3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13). Cytochrome c oxidases cannot directly accept reducing equivalents from quinol. For this purpose, there is an additional, middle component of the respiratory chain between dehydrogenase and oxidase, the cytochrome bc1 complex, which enables the transfer of electrons from quinol to cytochrome c. Respiratory chains of many aerobic bacteria, such as Paracoccus denitrificans and Azotobacter vinelandii, contain the bc1 complex and both types of terminal oxidase, cytochrome c and quinol oxidases. For instance, P. denitrificans has one quinol oxidase (ba3) and two cytochrome c oxidases (aa3 and cbb3) (14, 15). A. vinelandii has two quinol oxidases (bo3 and bd) and one cytochrome c oxidase (cbb3) (reviewed in references 16 and 17). As shown in Fig. 1, the aerobic respiratory chain of E. coli is much simpler than those of P. denitrificans and A. vinelandii. It lacks the cytochrome bc1 complex and any cytochrome c oxidase but contains instead three quinol oxidases, bo3, bd, and bd-II (reviewed in references 17, 18, 19, 20, and 21).
Figure 1.

Simplified view of the E. coli respiratory chain under aerobic and microaerobic conditions. The two NADH-quinone oxidoreductases called NDH-I and NDH-II and succinate-quinone oxidoreductase (SQR) transfer reducing equivalents to ubiquinone-8 (UQ-8) to yield reduced UQ-8, ubiquinol-8. Three quinol-oxygen oxidoreductases, cytochrome bo3 (CyoABCD), cytochrome bd (CydABX), and cytochrome bd-II (AppBCX), oxidize ubiquinol-8 and reduce O2 to 2H2O. CydABX and possibly AppBCX oxidize menaquinol-8. NDH-I, CyoABCD, CydABX, and AppBCX are coupled (ΔμH+ generators); NDH-II and SQR are uncoupled (no ΔμH+ generation). The energetic efficiency of each enzyme is indicated as the number of protons delivered to the periplasmic side of the membrane per electron (H+/e− ratio). doi:10.1128/ecosalplus.ESP-0012-2015.f1
Physiological Functions
Cytochrome bo3 predominates under high aeration, whereas cytochrome bd is expressed under low oxygen tension (22, 23, 24) (Table 1). It is of interest that the cytochrome bo3 level increases about 150-fold during aerobic growth, but the cytochrome bd level falls only 3-fold, i.e., the change in the cytochrome bd level in response to oxygen is much smaller than that of the cytochrome bo3 level (23). Both oxidases catalyze the oxidation of ubiquinol-8 to allow cellular respiration with oxygen as the terminal electron acceptor (25). Cytochrome bd can also oxidize menaquinol-8 (26, 27), which replaces ubiquinol-8 upon a change of growth conditions from aerobic to anaerobic (2).
Table 1.
Properties of cytochrome bo3 and cytochrome bd in E. coli
| Property | Cytochrome bo3 | Cytochrome bd |
|---|---|---|
| Level of O2 tension for expressiona | High | Low |
| Catalyzed reaction of oxidationb | Ubiquinol-8 → ubiquinone-8 | Ubi(mena)quinol-8 → ubi(mena)quinone-8 |
| Catalyzed reaction of reductionb | O2 → 2H2O | O2 → 2H2O |
| Energetic efficiency (H+/e− ratio)c | 2 (true proton pump) | 1 |
| KD (O2) (μM)d | >300 | 0.28 |
| Apparent Km for O2 (μM)e | 0.016–2.9 | 0.003–2 |
| Apparent Km for nonphysiological reductants (mM):f | ||
| Ubiquinol-1 | 0.05 | 0.23 |
| Menadiol | 38 | 1.67 |
| TMPD (N,N,N′,N′-tetramethyl-p-phenylenediamine) | 9.5 | 18.2 |
| Operon encoding oxidaseg | cyoABCDE | cydABX |
| Subunits (mass [kDa])h | CyoA (33.5) | CydA (57) |
| CyoB (75) | CydB (43) | |
| CyoC (20.5) | CydX (4) | |
| CyoD (12) | ||
| Redox cofactors (Em value[s] [mV])i | Heme b (+180, +280) | Heme b558 (+176) |
| Heme o3 (+180, +280) | Heme b595 (+168) | |
| CuB (+370) | Heme d (+258) |
Both cytochrome bo3 and cytochrome bd are primary generators of a transmembrane gradient of electrochemical H+ potentials (ΔμH+), because the reaction arising from the transfer of reducing equivalents from quinol to O2 is coupled directly to transmembrane charge separation (28, 29, 32, 50, 51, 52, 53, 54). The energy conserved in the form of ΔμH+ can be used by the E. coli cell for ATP synthesis, the transport of nutrients, and other useful work. Thus, the main function of both oxidases is energy conservation. The two enzymes, however, are different in their bioenergetic efficiencies (transmembrane proton translocation ratios, or the number of protons delivered to the periplasmic side of the membrane per electron [H+/e− ratios]), with an H+/e− ratio of 2 for cytochrome bo3 and an H+/e− ratio of 1 for cytochrome bd (28, 29) (Table 1). This difference is because cytochrome bo3 is a true proton pump, whereas cytochrome bd is not capable of proton pumping (28, 29). As sources of oxidizing power, cytochrome bo3 and cytochrome bd can support disulfide bond formation upon protein folding catalyzed by the DsbA-DsbB system (55).
Apart from energy conservation, cytochrome bd endows E. coli with a number of specific physiological functions. Cytochrome bd can serve as an oxygen scavenger and inhibit the degradation of O2-sensitive enzymes present under anaerobic and microaerophilic conditions (56). In a recent systematic mutational analysis to elucidate the contribution of the respiratory pathways to the abilities of commensal and pathogenic (enterohemorrhagic) E. coli strains to colonize a streptomycin-treated mouse intestine, mutants lacking cytochrome bd failed to colonize whereas cytochrome bo3 was found not to be necessary for colonization (57).
The cytochrome bd contents increase not only at low oxygen concentrations, but also under some unfavorable conditions, such as alkaline pH (58), high temperature (59, 60), the presence of uncouplers-protonophores (58, 61, 62), and low concentrations of cyanide (63) in growth media. Mutants defective in cytochrome bd are sensitive to H2O2 (60) and a self-produced extracellular factor that inhibits their growth (64, 65). Mutants that cannot synthesize cytochrome bd are also unable to exit from the stationary phase and resume aerobic growth at 37°C (66, 67). The expression of cytochrome bd, instead of cytochrome bo3, may enhance bacterial tolerance to oxidative and nitrosative stresses (68, 69, 70, 71). In particular, cytochrome bd contributes to bacterial resistance to peroxynitrite (71, 72), nitric oxide (68, 69, 70, 71, 73, 74, 75, 76, 77), carbon monoxide (78, 79), and hydrogen peroxide (59, 70, 71, 80, 81, 82, 83, 84, 85, 86).
Inhibitors
Table 2 compares the effects of different inhibitors on the respiratory activities of cytochrome bo3 and cytochrome bd in E. coli. Inhibitors of the quinol oxidases can be divided into two groups: quinol-like compounds acting at a quinol-binding site(s) and heme ligands (e.g., cyanide, azide, and NO) acting at the oxygen-binding/reducing site. A specific feature of cytochrome bd is that it is much less sensitive to cyanide and azide than cytochrome bo3 (32) (Table 2). The lower sensitivity of cytochrome bd to anionic heme ligands may be a result of an elevated electron density on the central ion of iron due to the breaking of the circle conjugate π-electron structure in the d-type porphyrin ring and/or may point to a more hydrophobic environment for the O2-reducing site of cytochrome bd than for that of cytochrome bo3. It is of interest that the quinolone-type compound aurachin D and its derivatives in submicromolar concentrations specifically inhibit cytochrome bd but virtually do not affect cytochrome bo3 (87). These outcomes may indicate some differences in the specific structures of quinol-binding sites in cytochrome bo3 and cytochrome bd. It has been shown recently that cytochrome bd in E. coli is a bacterial membrane target for a cationic cyclic decapeptide, gramicidin S (50% inhibitory concentration, ∼5.3 μM) (Table 2), although it has been generally accepted that the main target of gramicidin S is the membrane lipid bilayer rather than the protein components (88). This finding can provide new insight for the molecular design and development of novel gramicidin S-based antibiotics. The effect of gramicidin S on cytochrome bd and some other membrane-bound proteins may be the alteration of the protein structure through binding to the hydrophobic protein surface (88).
Table 2.
Effects of inhibitors on quinol oxidase activities of cytochrome bo3 and cytochrome bd in E. coli
| Inhibitora | Concentration, inhibition for: | |
|---|---|---|
| Cytochrome bo3 | Cytochrome bd | |
| KCN | 10 μM | 2 mM |
| NaN3 | 15 mM | 400 mM |
| H2O2 | 300 mM | 120 mM |
| HOQNO (2-n-heptyl-4-hydroxyquinoline N-oxide) | 2 μM | 7 μM |
| ZnSO4 | 1 μM | 60 μM |
| Piericidin A | 2 μM | 15 μM |
| Antimicin A | 50 μM, 18% | 50 μM, 80% |
| UHDBT (undecylhydroxydioxobenzothiazole) | 20 μM, 97% | 20 μM, 18% |
| (1,5-Dimethylhexyl)quinazolinamide | 100 μM, 23% | 100 μM, 88% |
| (1-Methyldecyl)quinazolinamide | 100 μM, 24% | 100 μM, 85% |
| Stigmatellin | 200 μM, 94% | 200 μM, 14% |
| Nigericin | 100 μM, 35% | 100 μM, 44% |
| Dibromothymoquinone | 100 μM, 82% | 100 μM, 38% |
| Aurachin A | 700 μM, 56% | 700 μM, 27% |
| Aurachin C | 214 nM, 90% | 214 nM, 90% |
| Aurachin D | 400 nM, 5% | 400 nM, 93% |
| NO | << 10−8 M | 100 nM |
| PCP | 200 μM | |
| TTFA | 1 mM, 35% | |
| Gramicidin S | 189 μM | 5.3 μM |
The concentrations shown for KCN, NaN3, H2O2, HOQNO, ZnSO4, and piericidin A (32) and gramicidin S (88) are the concentrations required for 50% inhibition of the ubiquinol-1 oxidase activities of the purified cytochromes bo3 and bd. For PCP (pentachlorophenol) and NO (nitric oxide), the inhibition constants (Ki values) are shown (73, 89). For TTFA (2-thenoyl trifluoroacetone), the data shown are the concentrations yielding the indicated percent inhibition of the ubiquinol-1 oxidase activity of purified cytochrome bd (47). For other inhibitors, the data shown are the concentrations yielding the indicated percent inhibition of the duroquinol oxidase activities of the membranes containing either cytochrome bo3 or cytochrome bd (87).
GENETICS
Oxidase Encoding
Cytochrome bo3
Cytochrome bo3 is composed of four different subunits (90, 91, 92) encoded by the cyoABCDE operon (39, 45) (Table 1). The cyoABCDE operon, located at 10.2 min on the E. coli genetic map (39, 93), has been cloned and sequenced (44). Subunits I (75 kDa), II (33.5 kDa), and III (20.5 kDa) of cytochrome bo3 appeared to be homologous to the counterparts of the eukaryotic and prokaryotic aa3-type cytochrome c oxidases (45) and were referred to as cyoB, cyoA, and cyoC gene products, respectively, as determined by protein sequencing (46) and other approaches (44, 94, 95). Thus, cytochrome bo3 is a member of the heme-copper terminal oxidase superfamily (45, 96). The cyoD gene encodes subunit IV (12 kDa) (95, 97, 98), which is homologous to the counterpart in cytochrome c oxidases from Gram-positive bacteria and terminal quinol oxidases but unrelated to eukaryotic cytochrome c oxidases (96, 99). The cyoE gene, located at the 3′ end of the cyo operon, encodes no subunit of cytochrome bo3 but does encode the enzyme that catalyzes the transformation of heme B (protoporphyrin IX, or protoheme) to heme O (uppercase letters in heme designations highlight the chemical nature of hemes, as opposed to protein-bound hemes) by attaching a long farnesyl side chain to the former (100, 101, 102). Heme O is specifically required for the binuclear oxygen-reducing site of cytochrome bo3. Subunit I (CyoB) carries all three metal redox cofactors: low-spin heme b, high-spin heme o3, and a copper ion (CuB) (3, 94, 103, 104). Heme o3 and CuB form the heme-copper binuclear center for dioxygen reduction. Unlike aa3-type cytochrome oxidase, subunit II (CyoA) does not have any metal redox cofactors. Subunits III (CyoC) and IV (CyoD) can be removed from cytochrome bo3 without any loss of catalytic activity (38) but seem to be required for the assembly of the metal redox cofactors in subunit I (19, 105).
Cytochrome bd
Until recently, cytochrome bd was thought to be a two-subunit oxidase (106) encoded by the cydAB operon (39, 40, 41), located at 16.6 min on the E. coli genetic map (39, 93), and it was cloned (107) and sequenced (41). The molecular masses of subunit I (CydA) and subunit II (CydB) determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis, 57 and 43 kDa (47), respectively, are consistent with those determined based on the corresponding DNA sequences, 58 and 42.5 kDa (41). Neither of these two major subunits of cytochrome bd has homology to any subunit of other known respiratory chain oxidases, such as cytochrome bo3 and cytochrome c oxidases (41, 108). However, very recently it has been reported that cytochrome bd has an additional polypeptide named CydX (42, 43). This small (4 kDa) protein seems to be the third subunit of cytochrome bd. CydX is required for maintenance of the cytochrome bd activity and contributes to stabilization of the hemes (42, 43, 80, 109, 110). Thus it is now clear that cytochrome bd is a three-subunit oxidase encoded by the cydABX operon (21, 42, 43). Cytochrome bd contains no copper atoms and does not pump protons (28, 29, 32, 47, 50). Therefore, cytochrome bd is not a member of the heme-copper terminal oxidase superfamily. Cytochrome bd subunits carry three hemes: b558, b595, and d (106, 111). Heme b558 is located on subunit I (CydA), whereas hemes b595 and d are likely to be in the area of the subunit contact (112). CydA can be expressed and purified without CydB by using mutant strains defective in cydB (113). The purified CydA retains heme b558 but lacks hemes b595 and d (113). In addition to the cydABX operon, two other genes, cydC and cydD of the cydCD operon, located at 19 min on the E. coli genetic map (114, 115, 116), are essential for the assembly of cytochrome bd (114, 116, 117, 118). CydC and CydD, however, are not subunits of cytochrome bd. It was shown previously that cydCD encodes a heterodimeric ATP-binding cassette-type transporter (ABC transporter) that is a transport system for the thiol-containing redox-active molecules cysteine and glutathione (21, 119).
Regulation of Gene Expression
Under high oxygen tension, E. coli expresses cytochrome bo3 (encoded by cyoABCDE), whereas cytochrome bd (encoded by cydABX) is moderately repressed (22, 23, 24). The expression of the cyoABCDE and cydABX operons is controlled by the two global transcriptional regulators Arc and Fnr (2, 23, 120, 121, 122, 123, 124, 125, 126). Arc is a two-component regulatory system that includes ArcA, a cytosolic response regulator, and ArcB, a transmembrane histidine kinase sensor. ArcA controls several hundred genes (127) and responds to the oxidation state of the quinone pool, which is sensed by ArcB (128). ArcB is activated in the course of the transition from aerobic to microaerobic growth and remains active during anaerobic growth. Upon stimulation, ArcB autophosphorylates and then transphosphorylates ArcA (128, 129). Under microaerobic conditions (i.e., oxygen tension of 2 to 15% of air saturation), the increased level of phosphorylated ArcA represses the cyoABCDE operon and activates the cydABX operon (130).
Another global regulator, Fnr (an oxygen-labile transcription factor regulating hundreds of genes), controls the induction of anaerobic processes in E. coli (131, 132). The Fnr protein has an Fe-S cluster which serves as a redox sensor. The levels of the Fnr protein are similar under both aerobic and anaerobic conditions (122, 133), but the protein is active only during anaerobic growth. The active Fnr protein represses both cyoABCDE and cydABX operons during a transition to anaerobic conditions (i.e., oxygen tension of less than 2% of air saturation) (124, 125, 133).
PROSTHETIC GROUPS
Quinones
Quinones are lipophilic molecules dissolved within the lipid bilayer of the cytoplasmic membrane. The E. coli membrane contains three types of quinones that all have an octaprenyl side chain (C40). These are a benzoquinone, ubiquinone, and two naphthoquinones, menaquinone and dimethylmenaquinone. Both cytochrome bo3 and cytochrome bd catalyze the two-electron oxidation of ubiquinol-8 to ubiquinone-8 (Fig. 2) (apparent midpoint redox potential [Em] = +110 mV [2]), coupled to the four-electron reduction of O2 to 2H2O (Em = +820 mV). Cytochrome bd can also oxidize menaquinol-8 to menaquinone-8 (Fig. 2) (Em = −80 mV [2]) (26, 27). Ubiquinone-8 is predominant during aerobic growth but is replaced by menaquinone-8 upon the transition from aerobiosis to anaerobiosis (2, 134, 135). It was shown previously that cytochrome bo3 contains a tightly bound ubiquinone (136). The presence or absence of bound quinone in solubilized cytochrome bd depends on the purification protocol (20). In some preparations of purified cytochrome bd, there is no quinone (32, 47, 51, 137), whereas others clearly contain bound quinone (52, 54, 138).
Figure 2.

Structures of ubiquinone-8, the reduced ubiquinone-8 (ubiquinol-8), menaquinone-8, and the reduced menaquinone-8 (menaquinol-8). doi:10.1128/ecosalplus.ESP-0012-2015.f2
Cytochrome bo3 Prosthetic Groups
Cytochrome bo3 contains three redox-active metal groups: a low-spin heme b, which is involved in quinol oxidation, heme o3, and CuB; the latter two groups compose a binuclear center, which is the site of the binding of O2 and its reduction to water. The chemical cofactors are heme b, corresponding to protoporphyrin IX (protoheme, or heme B); heme o3, corresponding to the protoheme with a long farnesyl side chain attached (heme O); and the CuB center, represented by the Cu atom ligated by three histidine residues. In the three-dimensional (3D) structure of cytochrome bo3, all prosthetic groups were found to be within subunit I (139). In addition to the metal centers, there is also one tightly bound ubiquinone cofactor. Although the X-ray structure did not show any bound quinone in the crystallized enzyme, the site-directed mutagenesis studies identified residues that modulate the properties of the bound quinone (140, 141, 142).
Heme b
Subunit I of the bo3 oxidase contains 15 transmembrane helices. Heme b is located between helices II and X at a depth of about one-third of the membrane thickness from the P (positive) side and oriented perpendicular to the membrane plane, such that its propionates are exposed to the P side of the membrane. In both reduced (S [spin quantum number] = 0) and oxidized (S = 1/2) states (143, 144), the low-spin heme iron is bound to the four nitrogen atoms of the porphyrin ring and to the two conserved histidines of subunit I (H106I and H421I; subscript “I” identifies the subunit where the residue is). Heme b is a direct electron donor for the catalytic binuclear site. It transfers electrons obtained from the bound ubiquinone. The Em of this heme is about +280 mV (without redox interaction) (48). The optical spectrum of this heme is rather characteristic for low-spin six-coordinate heme b, with the α-band of the reduced heme at 562 nm in the absolute and the difference (reduced-minus-oxidized) absorption spectra. The maxima of the β- and Soret bands of the reduced-minus-oxidized heme b spectrum are at 530 and 430 nm, respectively. At a high level of cytochrome bo3 expression, heme B in the low-spin heme site can be replaced by heme O (145). Such heme alteration does not decrease the enzyme activity but lowers the functional Km of the enzyme for oxygen (36).
Heme o3
Heme o3 is the oxygen-binding heme, and the subscript 3 has been used historically to indicate this feature by analogy to the other heme-copper oxidases. Heme o3 is a high-spin heme in both the fully reduced ferrous state (S = 2) (146) and the oxidized ferric state (S = 5/2) (147). Depending on the conditions, heme o3 may be penta- or hexacoordinate (147): the permanent bonds of the heme iron include four bonds with nitrogen atoms of the porphyrin ring and one extra bond with the conserved H419I from helix X at the same depth in the membrane as heme b. Fivefold coordination of the heme iron leaves one side of the heme empty and available for the binding of ligands such as O2. This free coordination points toward the CuB, together with which heme o3 forms the bimetallic catalytic site where the reduction of oxygen to water takes place. The spectrum of heme o3 is characteristic of high-spin hemes. It has a broad α-band centered at 560 nm in the reduced-minus-oxidized spectrum with small extinction and an intense Soret band with the maximum at 430 nm (145). The redox properties of the high-spin heme o3 are very similar to those of heme b (48). Both of these hemes have a redox potential of about 280 mV when the neighboring heme is oxidized, but the reduction of the neighbor results in a 100-mV lowering of the heme potential (redox interaction).
CuB
The second partner in the binuclear catalytic site is CuB. The oxidized CuB is a tetragonal center (148, 149); it has three permanent axial histidine imidazole ligands and one mobile oxygen ligand with an exchangeable proton(s) (148, 149). The imidazole ligands originate from H284I in helix VI and from H333I and H334I, both located in a loop fragment between helices VII and VIII. This redox center is often called the invisible Cu site because it does not show any changes in optical spectra upon enzyme reduction and oxidation. Usually, the electron paramagnetic resonance (EPR) signal from the oxidized Cu can be easily detected, but the close proximity of the iron atom of heme o3 results in strong magnetic interaction and the absence of any detectable spectrum of CuB. Such strong magnetic interaction, however, can help to define the Em of the tetragonal center. The reduction of the Cu ion brakes magnetic interaction with the high-spin heme, and the appearance of a high-spin EPR signal can give information on the reduction level of CuB. With such an approach, an Em of +370 mV for CuB was obtained (48).
Bound ubiquinone
It has been proposed that the bo3 oxidase can have two ubiquinone-binding sites with different affinities. The bound ubiquinone in the site with high affinity for ubiquinone (the QH site) can be considered to be the enzyme cofactor (136, 150, 151, 152). At the same time, in the low-affinity (QL) site, fast exchange of the ubiquinone molecules occurs, and it was proposed that electrons are transferred from the QL site to the next electron acceptor (heme b) via the QH site (136, 153). A functional study of mutants obtained by site-directed mutagenesis was used to create a model for the possible QH binding site, which is located in subunit I, close to heme b (139). According to this model, the QH site is predicted to form up to four hydrogen bonds with D75 and R71 to the 1-carbonyl oxygen and with H98 and Q101 to the 4-carbonyl oxygen. EPR spectroscopy demonstrated that the QH site stabilizes a semiquinone anion radical of bound ubiquinone (154, 155, 156).
Cytochrome bd Prosthetic Groups
Cytochrome bd is composed of three subunits, I (CydA), II (CydB), and III (CydX), which are typical integral membrane proteins. The subunits carry three metal-containing redox centers, such as two protoheme IX groups (hemes b558 and b595) and a chlorin molecule (heme d) (Fig. 3), which are proposed to be in 1:1:1 stoichiometry per the enzyme complex, but no copper ion (157, 158, 159, 160,161). Heme b558 appears to be located within subunit I. Subunits I and II are required for the assembly of heme b595 and heme d, suggesting that these two hemes may reside at the subunit interface (112). Both heme b558 and heme d are presumed to be oriented with their heme planes perpendicular to the membrane plane. Heme b595 is possibly oriented with its heme plane at ∼55° with respect to the plane of the membrane (162).
Figure 3.

Structures of heme B (protoheme IX), heme O, and heme D (chlorin), which are redox cofactors of cytochrome bo3 and/or cytochrome bd from E. coli. doi:10.1128/ecosalplus.ESP-0012-2015.f3
Heme b558
Heme b558 was shown to be located on subunit I. Although subunits I and II of the isolated cytochrome bd complex cannot be split apart without denaturing the enzyme, some genetic approaches have allowed subunit I to be synthesized in the absence of subunit II (113). Antibodies directed against subunit I (163, 164), as well as selective proteolysis of this subunit (165, 166), inhibit the ubiquinol oxidase activity of cytochrome bd. These findings suggest that heme b558 is associated with subunit I and involved in quinol oxidation. The α-band of the reduced heme b558 reveals a peak at 560 to 562 nm in the absolute and the difference (reduced-minus-oxidized) absorption spectra at room temperature. The maximum of the β-band of the reduced heme is at 531 to 532 nm (167, 168). The difference absorption spectrum of heme b558 in the Soret region reveals a maximum at 429 nm and a minimum at 413 nm (168). The positions of these bands were confirmed upon redox titration by separating the composite difference absorption spectra of the enzyme into the contributions of the individual heme components (111, 169) as well as by the detailed spectroelectrochemical redox titration and numerical modeling of the data (168). Low temperature (77 K) shifts all bands by 1 to 4 nm to the blue (167). Heme b558 is a low-spin hexacoordinate (170, 171, 172), and amino acid residues H186 and M393 of subunit I were identified as its axial ligands (173, 174, 175). The location of heme b558 is predicted to be near the periplasmic surface (176, 177).
Heme b595
The band with a maximum at ∼595 nm in the difference (reduced-minus-“air-oxidized”) absorption spectrum of cytochrome bd was long ascribed to heme a1 because of the relatively bathochromic position of the extremum (178). Subsequently, it was established that the prosthetic group of this component is not heme a but protoheme IX (47, 111), whereupon the component has been called heme b595. Magnetic circular dichroism (MCD) studies confirmed such a conclusion (170, 172). The difference absorption spectrum of heme b595 in the visible region was resolved from a set of composite spectra of the isolated cytochrome bd recorded at different redox potentials (111, 169). It turned out that this spectrum is similar to that of catalases and peroxidases, containing pentacoordinate (high-spin) protoheme IX (111, 168). Heme b595 shows an α-band at 594 nm and a β-band at 561 nm in the difference absorption spectrum (168). A trough at 643 nm in the difference spectrum of heme b595 (168) is indicative of the disappearance of a charge transfer to the ligand band, characteristic of the oxidized high-spin heme b, as in the case of peroxidases. The γ-band of ferrous heme b595 in the absolute absorption spectrum is characterized by a maximum at ∼440 nm, as clearly revealed by femtosecond spectroscopy (179). The difference absorption spectrum of heme b595 in the Soret region shows a maximum at 439 nm and a minimum at 400 nm (168). Heme b595 is a high-spin pentacoordinate (170, 172) ligated by the histidine (H19) of subunit I (180) and located on the periplasmic surface (176, 177). The role of heme b595 remains obscure. It is proposed that heme b595 participates in the reduction of oxygen, forming, together with heme d, a diheme oxygen-reducing site somewhat similar to the heme-Cu oxygen-reducing site in the aa3- and bo3-type oxidases (52, 170, 171, 179, 181, 182, 183, 184, 185, 186, 187, 188, 189). In favor of this hypothesis is the finding that the circular dichroism (CD) spectrum of the reduced wild-type cytochrome bd in the Soret band shows strong excitonic interaction between ferrous hemes d and b595 (171). Modeling the excitonic interactions in absorption and CD spectra yields an estimate of the Fe-to-Fe distance between heme d and heme b595 to be about 10 Å (171). In the opinion of other researchers, the function of heme b595 is limited to the transfer of an electron from heme b558 to heme d (190, 191). Some authors believe that heme b595 can form a second site capable of reacting with oxygen (35, 157).
Heme d
Heme d is a chlorin-type molecule (192). The α-band of the reduced heme d in the absolute absorption spectrum shows a peak at 628 to 630 nm (89). However, under usual conditions, heme d is in the stable oxygenated (oxygen-ligated ferrous) form, which is characterized by a band with a maximum at 647 to 650 nm in the absolute absorption spectrum (193). The reduced-minus-oxidized difference spectrum of heme d in the Soret region shows a maximum at 430 nm and a minimum at 405 nm (168). The spectral contribution of heme d to the complex γ-band is much smaller than those of either hemes b (168). Heme d is predicted to be located near the periplasmic surface (176, 177). This heme is the place for capturing and reducing O2 to 2H2O. Being free of external ligands, heme d seems to be in the high-spin state. The protein ligand of heme d is not known. The data on the nature of the heme d axial ligand are controversial. Authors of resonance Raman and electron nuclear double resonance studies have claimed that it cannot be an ordinary histidine, cysteine, or tyrosinate but is either a weakly coordinating protein donor or a water molecule (180, 194, 195). In contrast, EPR studies have indicated that the heme d axial ligand is histidine in an anomalous condition or another nitrogenous amino acid residue (196). Finally, it has been reported recently that the highly conserved glutamate 99 of subunit I may be a candidate for such a role (137).
The millimolar extinction coefficients used commonly for the determination of the E. coli cytochrome bd concentration are listed in Table 3.
Table 3.
Extinction coefficients used for determination of the E. coli cytochrome bd concentration
| Absorption spectrum | Heme(s) | Wavelength paira (nm) | Δεb (mM−1·cm−1) | Reference |
|---|---|---|---|---|
| Difference spectra | ||||
| Fully reduced minus as prepared | d | 628–607 | 10.8 | 170 |
| d | 628–651* | 27.9 | 182 | |
| d | 628–649* | 18.8 | 32 | |
| b 558 | 561–580 | 21 | 182 | |
| b 595 | 595–606.5 | 1.9 | 182 | |
| All | 429–700† | 303 | 182 | |
| Fully reduced CO bound minus fully reduced | d | 642–622 | 12.6 | 32 |
| d | 643–623 | 13.2 | 54 | |
| Absolute spectra | ||||
| Fully reduced | d | 628–670 | 25 | 52 |
| As prepared | All | 414–700† | 223 | 182 |
Values marked with an asterisk or dagger cannot be recommended for the determination of the cytochrome bd concentration since, for those marked with an asterisk (*), the as-prepared enzyme contains various amounts of the ferrous heme d-oxy complex that absorbs at 649 to 651 nm and, for those marked with a dagger (†), the intensity of the Soret band is variable depending on the purity of the preparation.
Δε, extinction coefficient.
Redox Potentials of Hemes in Cytochrome bd
The Em values of hemes b558, b595, and d in the E. coli cytochrome bd solubilized in n-dodecyl-β-d-maltoside at pH 7.0 are +176, +168, and +258 mV, respectively (49) (Table 1). Similar values were reported for cytochrome bd contained in bacterial membranes (158, 159, 197), reconstituted in liposomes (198), or solubilized in Tween 20 or Triton X-100, in which the enzyme is active (198). Notably, the Em value of heme b558 can depend on the detergent used for solubilization (198). In particular, octylglucoside and cholate cause a large decrease in the Em value of heme b558 that also correlates with the reversible inactivation of the enzyme (198). The Em values of all three heme components of cytochrome bd are sensitive to pH values between 5.8 and 8.3, with dependence of −61 mV per pH unit for heme d and −40 mV per pH unit for hemes b558 and b595, indicating that the reduction of the bd complex is accompanied by enzyme protonation (198). The spectroelectrochemical redox titration and spectral modeling show the strong redox interaction between heme b558 and heme b595, while the interaction between heme d and either hemes b is insignificant (168). Of interest, in the absence of heme d the interaction potential between heme b558 and heme b595 is much larger compared with the situation when heme d is present (168).
STRUCTURE OF bo3 AND bd OXIDASES
3D Structure of bo3 Oxidase
The 3D structure of the bo3 oxidase from E. coli at 3.5-Å resolution was reported in 2000 (139). The structure confirms that the overall architecture of this complex is very similar to those of all other members of the heme-copper oxidase superfamily. The whole integral protein contains 25 transmembrane helices, of which subunit I has 15, subunit II has only 2, subunit III has 5, and subunit IV has 3 helices. All known enzyme cofactors were found in subunit I. The transmembrane helices of this subunit are not perpendicular to the membrane plane but are tilted about 20 to 35° against it. When viewed from the top (P side), they are arranged in an anticlockwise direction and form three semicircular arcs, organized in a quasithreefold axis of symmetry. Three pores are formed in the center of the arcs. One pore houses the binuclear center (heme o3 and CuB) of the oxidase and includes the proton-conductive K channel directed from the binuclear center toward the N (negative) side of the membrane. Another pore retains heme b; the last pore is devoid of cofactors but is used for the proton-conductive D channel.
Heme b is located at a depth of about one-third of the membrane thickness from the P side and oriented perpendicular to the membrane plane, such that its propionates are pointing toward the P side of the membrane (Fig. 4). The heme o3 is located at the same depth as heme b, the plane of heme o3 is also perpendicular to the membrane, and the propionates point toward the P side in a manner similar to heme a. Heme b and heme o3 are facing each other at an angle of 104 to 108o. At an ∼5-Å distance from the iron of heme o3, a copper atom designated CuB is located. In addition to this Cu atom that is coordinated by the three histidines, there is another important structure identified in all heme-copper oxidases and represented by the covalently bound H284I and Y288. The three histidines and the tyrosine form a conjugated π-electron system around the CuB center. Cytochrome bo3 has been proposed to have two ubiquinone-binding sites, one with high (QH) and one with low (QL) affinity for ubiquinone (136). It has also been postulated that electrons are transferred from the QL site to heme b via the ubiquinone bound at the QH site. Site-directed mutagenesis studies (140, 142) have identified residues that modulate the properties of the QH site. The model of a QH quinone-binding site including R71, D75, H98, and Q101 residues is also supported by the results of X-ray crystallography (139).
Figure 4.

Structure of cytochrome bo3 from E. coli. Only two main subunits are shown: subunit I in gray and subunit II in yellow. Hemes are shown in red (heme o3 on the right and heme b on the left). The cyan sphere near heme o3 represents the CuB center. The amino acid residues of possible proton-conducting pathways, the D (red-tag) and K (blue-tag) channels, in subunit I are shown. The most likely position of the membrane is depicted by the gray background. doi:10.1128/ecosalplus.ESP-0012-2015.f4
The protein medium itself cannot facilitate proton delivery toward the binuclear center or across the membrane, and in order to overcome this limitation, the oxidase has special proton-conductive structures. It is proposed that these structures are based on chains of hydrogen bonds between hydrogen-bonding protein side groups (polar and/or protonatable) and water molecules, whereby the proton is transferred via a Grotthuss type mechanism. At least two proton-conductive channels have been identified primarily by site-directed mutagenesis (199, 200, 201), and these findings were later confirmed by X-ray crystallography (139). One of them is the so-called K pathway, named after the highly conserved lysine 362 (199, 200), which is situated approximately halfway through the channel (Fig. 4). This pathway starts with S315 and S299 and continues through conserved residues K362I and T359I toward the hydroxyethyl farnesyl side chain of heme o3 and Y288I in the proximity of the binuclear center. The other channel is named D after the highly conserved D135I (201, 202), which is situated near the surface of the enzyme on the N side. D135I, together with T211I, forms a mouth that leads via polar residues N124I, N142I, S145I, T149I, T201I, and T204I to E286I (Fig. 4), which is an important residue for the proton-pumping mechanism.
Proposed Structure of Cytochrome bd
To date, the X-ray structure of cytochrome bd is not available; however, the findings from conventional studies of the protein topology in the membrane suggest that all three hemes are located near the periplasmic side of the membrane (176, 177). Figure 5 shows topological models of subunits I (CydA) and II (CydB) of cytochrome bd from E. coli (27). Both subunits are integral membrane proteins showing no sequence homology to the subunits of the heme-copper oxidase superfamily (e.g., cytochrome bo3). Subunit I consists of nine transmembrane helices, with the N terminus in the periplasm and the C terminus in the cytoplasm (176). Subunit II is composed of eight transmembrane helices, with both N and C termini in the cytoplasm (176). Newly discovered subunit III, the 37 amino acid CydX protein (42, 43, 109), is proposed to consist of the only membrane-spanning helix, with the N terminus in the cytoplasm and the C terminus in the periplasm (21, 80, 109). Subunit I contains a large hydrophilic domain, which is called the Q-loop, connecting transmembrane helices VI and VII. As shown by many experimental approaches, the Q-loop is involved in quinol binding and oxidation (163, 164, 165, 166, 203, 204). Thus, the quinol-oxidizing site in cytochrome bd is located on the periplasmic side of the membrane. It is worth mentioning that the size of the Q-loop is taken into account to categorize the bacterial cytochromes bd (20, 172).
Figure 5.

Proposed topology of the CydA and CydB subunits of cytochrome bd from E. coli. The axial ligands of heme b595 (H19) and heme b558 (H186 and M393) in the CydA subunit are shown in purple and red, respectively. The protein sequence data have been taken from information available at http://genolist.pasteur.fr/Colibri/. The alignment has been made by using the TOPO2 program available at http://www.sacs.ucsf.edu/TOPO2. The model is very similar to that reported in reference 27. doi:10.1128/ecosalplus.ESP-0012-2015.f5
Based on 815 sequences of the cytochrome bd gene available, it was possible to search for highly conserved residues in the corresponding protein (27). It was shown previously that subunit II evolved significantly faster than subunit I, with the result that subunit II exhibits greater sequence diversity (205). A number of residues in subunit I are totally (>99%) conserved in the 815 sequences (27). These residues include H19 (the heme b595 axial ligand [180]), H186 and M393 (the heme b558 axial ligands [173, 174, 175]), K252 and E257 (involved in quinol binding [204]), R448 (having an unknown function), and E99, E107, and S140 (proposed to be components of a proton channel [54, 176] and important for binding in the heme d-heme b595 diheme site [27, 137]). Slightly less conserved (95 to 99%) are E445 (required for charge compensation at the heme b595-heme d oxygen-reducing site upon the full reduction of oxygen by two electrons [52]), N148 (a plausible component of a proton channel), and R9 (having an unknown function) (27). Somewhat less conserved (∼85%) are R391 (which stabilizes the reduced form of heme b558 [206]) and D239 (having an unknown function); however, these residues are totally conserved within the group of cytochromes bd with a long Q-loop, to which the E. coli enzyme belongs (27). Other conserved residues are glycines, prolines, phenylalanines, and tryptophans, which may play structural roles. There is only one totally (>99%) conserved residue (W57) in subunit II (27). Residues R100, D29, and D120 of subunit II are totally conserved within the family of long-Q-loop cytochrome bd, and the residue at position 58 in subunit II (according to the numbering for the E. coli enzyme) is either an aspartate or a glutamate. The N-terminal portion of subunit II is thought to be involved in the binding of heme d and heme b595 (27, 207).
A multiple sequence alignment of 299 homologues of the newly found CydX subunit reveals the conserved amino acid motif that includes the residues Y3, W6, G9, and E/D25 (109). Of these residues, the first three are part of the predicted transmembrane α-helix being localized to the same side of the helix (109). The latter suggests that this is possibly the side of the α-helix of CydX that interacts with the other subunits in cytochrome bd (109).
Since the active site of O2 reduction is located near the periplasmic surface and protons for H2O production are taken from the bacterial cytoplasm, there must be at least one transmembrane proton-conducting pathway to convey protons from the cytoplasm to the heme b595-heme d site (51, 52, 54, 176) (Fig. 6).
Figure 6.

Schematic model of electron and proton transfer pathways in cytochrome bd from E. coli. There are two protonatable groups, XP and XN, redox coupled to the heme b595-heme d active site. A highly conserved residue, E445, was proposed to be either the XP group or the gateway in a channel that connects XP with the cytoplasm or the periplasm (52). A strictly conserved E107 residue is a part of the channel mediating proton transfer to XN from the cytoplasm (54). Xb, a group at the periplasmic side of the membrane that picks up and releases a proton as heme b558, is reduced and oxidized. doi:10.1128/ecosalplus.ESP-0012-2015.f6
COMPARISON OF AFFINITIES OF THE TWO OXIDASES FOR OXYGEN
In E. coli, the affinity of cytochrome bd for oxygen, i.e., dissociation constant for O2, KD (O2) of 0.28 μM (31), is about 1,000-fold higher than that of cytochrome bo3, KD (O2) >0.3 mM (30), which allows us to consider the bd and bo3 enzymes as the high- and low-affinity oxidases, respectively. Such a striking difference in the KD (O2) values correlates with the following facts.
Cytochrome bo3 predominates under high aeration, whereas cytochrome bd is expressed maximally under low aeration (22, 23, 24).
A peculiar feature of cytochrome bd is that it is purified mainly as a stable oxygenated (oxy)complex (heme b5583+-heme b5953+-heme d2+-O2) characterized by an absorption peak at 645 to 650 nm (193, 208, 209, 210). The fact that a stable oxycomplex can be generated by the reversible binding of oxygen to the one-electron-reduced cytochrome bd can be used for direct measuring of the KD (O2) of cytochrome bd (31, 49). This is not the case for cytochrome bo3 or any other heme-copper oxidase, which in any redox state under normal conditions does not form a stable oxycomplex; therefore, the KD (O2) for cytochrome bo3 can be determined only indirectly (30). For cytochrome bo3, the KD (O2) is more than 100-fold higher than the apparent Km for O2, which allows us to conclude that the bo3 oxidase is designed to trap O2 kinetically by reducing it to an oxoferryl species (36). Because of its ability to function efficiently under microaerobic conditions, cytochrome bd is required for commensal and pathogenic E. coli strains to colonize mouse intestine (57). It turns out that E. coli mutants lacking cytochrome bd, with high affinity for O2, are eliminated by competition with wild-type strains competent in respiration and that cytochrome bo3, with low affinity for O2, is not necessary for colonization (57). The colonization defects of the cytochrome bd mutants challenge the traditional view that the intestine is anaerobic (211) but support the hypothesis that a microaerobic niche is critical for both establishing and maintaining E. coli in the intestine (57).
BINDING OF LIGANDS (OTHER THAN OXYGEN)
Ligand Binding to the Cytochrome bo3 Binuclear Center
Ligands that bind to the binuclear center of cytochrome bo3 can be divided into two classes: uncharged molecules like CO and NO, which preferably bind to the binuclear center in the reduced form and induce the transition of the high-spin heme to the low-spin state, making it six coordinate, and ionizable molecules, like HCN and NaN3, or formate, which preferably bind to the oxidized heme. The binding of the second group of ligands can result in a different spin state for heme o3. Some of the ligands can bind either to heme o3 only or to a site between the two metals forming the binuclear center. The dynamics of ligand exchange can be used to characterize the possible dynamics of the binding of the substrate and the partial products of the reaction generated during the catalytic cycle.
CO and NO binding
The molecules of CO and NO mimic the oxygen molecule and can be bound to the catalytic site of cytochrome bo3. This binding occurs with the high-spin heme of the reduced binuclear center. Carbon monoxide reacts with the reduced enzyme with a stoichiometry of 1:1, and the KD for this reaction was determined to be 1.7 × 10−6 M (147). The value of a second-order rate constant for association (kon) is 6.1 × 104 M−1·s−1.
The CO ligation of heme o3 results in a blue shift of the heme absorption bands. The characteristic CO-binding spectrum has two small peaks in the visible part of the spectrum at 530 and 570 nm and a pronounced spectral shift in the Soret region, with the maximum λ of ∼415 nm and the minimum λ of ∼430 nm.
The photolysis of the reduced CO-bound enzyme at low temperatures results in the dislocation of the iron-bound CO to CuB, where it can be recognized by the specific C≡O stretch at ∼2,065 cm−1 due to CO bound to copper (212). At room temperature, the photodissociation of CO from the heme iron and its subsequent binding to CuB is an extremely fast reaction, and then CO remains bound to CuB for only a short time (two-component dissociation with the time constants of ∼14 and 140 μs [213]). After carbon monoxide dissociation from the binuclear site, the rebinding to the heme iron via CuB occurs at a much lower rate (214). Yoshikawa et al. (215) have reported the crystallographic structures of the reduced bovine enzyme in the presence and absence of CO. While this X-ray study did not find any significant changes in the structure of CuB relative to the fully oxidized bovine enzyme, the extended X-ray absorption fine-structure (148) investigation of the first shell of CuB showed a dramatic change upon the addition of CO, which involves the dissociation of one of the CuB histidine ligands and its possible replacement by a chloride ion.
Fully reduced cytochrome bo3 also binds NO with a stoichiometry of 1:1 and very high affinity (KD < 10−8 M) (216), forming the ferrous nitrosyl (Fe2+-NO) species, as determined by EPR spectroscopy (147). The heme o3-NO complex yields well-resolved EPR signals from the 14N atoms of both NO itself and the proximal histidine ligand of heme o3, showing nuclear hyperfine coupling (147).
Reaction of cytochrome bo3 with cyanide
Cyanide reacts almost exclusively with oxidized cytochrome bo3. The reaction is manifested by a red shift of the Soret band from approximately 411 to 415 nm (217). A characteristic feature of the ligand-binding reaction in the α-band region is the loss of the broad charge transfer band at 630 nm. This band is attributable to the fully oxidized binuclear center in which heme o3 is in a high-spin state. The results of EPR studies (217) seem to indicate that these absorption changes occur because the ferric high-spin heme o3 becomes a ferric low-spin ligand complex. Because of the magnetic coupling with CuB2+, the EPR spectrum (gz [constant of magnetization] = 3.3) is observed only when the copper of the binuclear center is reduced. This signal is characteristic of the conversion of the ferric high-spin heme into the low-spin state by the binding of strong-field ligands such as cyanide (218). Cytochrome bo3 binds a single equivalent of cyanide (KD, 1 × 10−6 to 2 × 10−6 M) with monophasic kinetics (219) and kon of 37 M−1·s−1 (at pH 6.0) (220). This rate constant is slightly pH dependent and increases about 1.8-fold over the pH range between 6.0 and 8.5 (220). Room-temperature MCD spectra in the near-infrared region (221), results from infrared and EPR studies (222), and also resonance Raman detection (223) of the product of the reaction of the oxidized cytochrome bo3 with cyanide have led us to propose the bridging structure of the cyanide complex to be as follows: Feo33+—C=N—CuB2+, where Feo3 is the iron of heme o3. It was also shown that the reduction of the enzyme results in the release of the CuB ligation and the formation of an Feo32+—C=N moiety.
Reaction of cytochrome bo3 with azide
Azide binds to cytochrome bo3 with a stoichiometry of 1:1; the KD for this reaction is about 2 × 10−5 M (224). Contrary to the addition of cyanide, the addition of azide to the oxidized isolated enzyme causes a relatively rapid but small blue shift in the Soret band from approximately 411 to 409 nm. Simultaneously, the broad charge transfer band at 630 nm becomes more pronounced and shifts its maximum to 640 nm. These small changes induced in the electronic absorption spectrum are consistent with heme o becoming hexacoordinate but remaining high spin, which can also be seen by MCD spectroscopy in the range of 350 to 1,100 nm (221, 224). The kinetics of azide binding is an order of magnitude faster than that observed for the binding of cyanide. The calculated kon for the binding of azide to cytochrome bo3 is about 800 M−1·s−1 at pH 7.5. The kon shows a marked increase upon acidification, indicating that the active species is electroneutral hydrazoic acid. Analyses of EPR, electronic, and MCD spectra (219) were used to prove that, unlike cyanide, azide does not bind to heme o3 but rather to the CuB site, whereas the resolved 3D structure of the bovine cytochrome c oxidase in the presence of azide revealed the bridging structure of the complex (Fea33+—N=N=N—CuB2+) (215). The Fourier transform infrared (FTIR) study of the bovine enzyme complex (225) showed a major infrared band at 2,051 cm−1. Subsequently, an FTIR spectroscopic study of the E. coli bo3 oxidase (226) showed an infrared band at 2,041 cm−1, which was assigned to the bridging structure. Discrepancy in the ligation geometry found by the different methods was resolved by detailed analysis of the FTIR spectroscopy of the complex of cytochrome bo3 with asymmetrically 15N-labeled azide (227). The experiments showed time-dependent evolution of the geometry of azide binding. In the air-oxidized form, a major infrared azide antisymmetric stretching band corresponds to the bridging geometry. An additional band developing at 2,062.5 cm−1 during longer incubation reflects the appearance of the CuB2+—N=N=N structure. In addition, the partial reduction of the oxidase with β-NADH caused the appearance of new infrared bands indicating the emergence of the Feo3+—N=N=N configuration (227).
Ligand Binding to Cytochrome bd
Since hemes d and b595 in cytochrome bd are in the high-spin pentacoordinate state, these redox centers can potentially bind ligands. It has to be expected that the reduced form of the bd enzyme can bind mainly electroneutral molecules like O2, CO, and NO and that the oxidized cytochrome bd binds ligands in the anionic form, such as cyanide and azide. It appears that heme d binds ligands readily but that the ligand reactivity of heme b595 is minor (170, 184, 189). It was reported previously that heme b558, although a low-spin hexacoordinate, may also bind ligands (e.g., CO and cyanide) to some extent (170, 189). Such marginal reactivity is due possibly to the weakening of the bond of the methionine axial ligand (M393) with the heme b558 iron caused by the isolation procedure and/or protein denaturation (189).
CO binding
CO brings about a red shift of the heme d band, with a maximum at 643 to 644 nm, a minimum near 624 nm, and a peak at 540 nm. In the Soret band, CO binding to fully reduced cytochrome bd results in an absorption decrease and minima at 430 and 442 to 445 nm. Absorption perturbations in the Soret band and at 540 nm occur in parallel with the changes at 630 nm and reach saturation at 3 to 5 μM CO. The peak at 540 nm is probably either the β-band of the heme d-CO complex or part of its split α-band (189). A peculiar W-shaped curve in the Soret region of the difference spectrum can be caused by a small band shift for unligated heme b595 induced by CO interaction with the nearby heme d (179, 185). Only a small fraction of heme b595 in cytochrome bd binds CO at room temperature or low temperatures (from −70 to −100°C) (170, 184). CO binding with about 15% of heme b595 in the membrane-bound cytochrome bd at a cryogenic temperature (4 K) was observed with the aid of FTIR spectroscopy (181). The apparent KD for the CO complex with fully reduced cytochrome bd appeared to be ∼80 nM (189). This value is markedly higher than that for cytochrome bo3 (1.7 μM) (147). The fully reduced cytochrome bd can form a photosensitive heme d-CO complex, and following flash photolysis, CO recombines with ferrous heme d proportionally to the CO concentration, with kon of 8 × 107 M−1·s−1 (161) (Table 4). This value is much higher than that for cytochrome bo3 (6.1 × 104 M−1·s−1) (147).
Table 4.
Kinetic and thermodynamic parameters for the reaction of E. coli cytochrome bd with O2, CO, and NO at room temperature
| Value for ligandg: | ||||||
|---|---|---|---|---|---|---|
| Parameter | O2 | CO | NO | |||
| MV O2 | R-O2 | MV CO | R-CO | MV NO | R-NO | |
| kon (M−1 s−1) | 2 × 109b,e | 8 × 107b | ||||
| koff (s−1)f | 78 ± 0.5a | 4.2 ± 0.34a | 6.0 ± 0.2a | 0.036 ± 0.003a | 0.133 ± 0.005a | |
| KD (nM) | 280c | 80d | ||||
Data from reference 75.
Data from reference 161.
Data from reference 31.
Data from reference 189.
Data from reference 53.
koff values are means ± standard deviations.
R-O2, R-CO, and R-NO are complexes of fully reduced cytochrome bd with O2, CO, and NO, respectively. MV O2, MV CO, and MV NO are complexes of one-electron-reduced cytochrome bd with O2, CO, and NO, respectively.
CO can also react with one-electron-reduced oxygen-free cytochrome bd of E. coli, forming a mixed-valence (MV) CO compound in which both b-type hemes remain oxidized (heme b5583+-heme b5953+-heme d2+-CO) (75, 179, 185, 188). The heme d α-band is also positioned at 635 to 636 nm in the absolute absorption spectrum.
It was proposed that the redox state of the b-type hemes in cytochrome bd, presumably that of heme b595, controls the status of the pathway for ligand transfer between heme d and the bulk phase (75, 179). Two observations allowed us to draw such a conclusion.
Flash photolysis of the CO complex of the fully reduced cytochrome bd results in the complete photodissociation of the CO molecule into the bulk. If the experiment is repeated with the MV CO complex, a significant part of the CO flashed off heme d2+ (up to 70%) gets trapped inside the protein and undergoes geminate recombination with heme d2+ on a subnanosecond time scale (179).
The apparent off rate constant for the spontaneous dissociation (koff) of CO from heme d2+ is markedly lower for the MV state of cytochrome bd than for the fully reduced state of the bd oxidase (75) (Table 4).
Interaction with some nitrogen-containing ligands
A number of small nitrogen-containing molecules can react with fully reduced cytochrome bd from E. coli. NO3−, NO2−, N2O32− (trioxodinitrate), NH2OH, and NO, when added to membranes containing cytochrome bd or the purified enzyme, give the decrease in amplitude and shift the 630-nm band of ferrous heme d to 641 to 645 nm (73, 75, 157, 170, 196, 228, 229, 230, 231). The common peak position was interpreted as all ligands yielding the same or very similar heme-nitrosyl compounds (231). A red shift of the α-band of heme d2+, observed upon the interaction of nitrite with the fully reduced membrane-bound cytochrome bd, was accompanied by the slower formation of a trough at 438 nm in the difference (nitrite-treated-minus-reduced) absorption spectrum. These changes in the Soret band were ascribed to the formation of the product of the interaction of heme b595 with nitrite (heme b5952+-NO) (157).
Cytochrome bd can also form a stable complex with NO in an MV state, in which ligand-bound heme d is reduced (to heme d2+) while the other two hemes (b558 and b595) are oxidized (75, 196). The rates of NO dissociation from heme d2+ of cytochrome bd in both the fully reduced and MV states were determined previously (75). In the fully reduced state, NO dissociates from heme d2+ at an unusually high rate, 0.133 s−1 (75), which is ∼30-fold higher than koff measured for the ferrous heme a3 of the mitochondrial cytochrome c oxidase, 0.004 s−1 (232). These data are consistent with the proposal that in heme-copper oxidases, CuB acts as a gate, controlling ligand binding to the heme in the active site (233). Another remarkable feature of NO dissociation from cytochrome bd is that the koff value for the MV state (0.036 s−1), although still quite high, is significantly lower than that measured for the fully reduced enzyme (75) (Table 4). The same effect was observed with CO (see above) (75). These data show that the redox state of heme b595 controls the kinetic barrier for ligand dissociation from the active site of cytochrome bd. The unusually high rate of NO dissociation from cytochrome bd may explain the observation (73) that the NO-poisoned cytochrome bd recovers respiratory function much more rapidly than a heme-copper oxidase.
Of interest, cytochrome bd is not inactivated by up to 100 μM peroxynitrite, another nitrogen-containing harmful reactive species (72). Furthermore, cytochrome bd in turnover with O2 is able to rapidly metabolize peroxynitrite with an apparent turnover rate increasing at higher concentrations of the reducing substrates (72). Thus, cytochrome bd appears to be highly resistant to peroxynitrite damage (71, 72). It is suggested that the expression of bd-type instead of heme-copper-type oxidases enhances bacterial tolerance to nitrosative and oxidative stresses, thus promoting the colonization of host intestine or other microaerobic environments (68, 69, 70, 71, 72, 75, 76).
Reaction with cyanide
An earlier spectrophotometric study of the reaction of cytochrome bd from E. coli with KCN was performed with the membrane vesicles (234) and was confined to measurements in the α-band. That work revealed a decay of the absorbance at 650 nm induced by cyanide, and this finding was interpreted at that time to represent the disappearance of the free form of ferric heme d (234). This result was reinterpreted later as the decay of the ferrous heme d oxycomplex (167). Cyanide, interacting with the oxygenated form of the isolated cytochrome bd, brings about significant absorption changes in the γ-region: a maximum at 434 to 437 nm and a minimum near 405 to 410 nm in the difference (KCN-treated-minus-oxygenated) spectrum. There is also considerable increase in the MCD signal in the Soret region. These data were interpreted to indicate the ligand-induced transition of high-spin heme b595 to the low-spin state (235). Subsequently, a simple and fast method for the conversion of the oxygenated enzyme into the fully oxidized form with the use of lipophilic electron acceptors was developed (193). This approach enabled researchers to study the interaction of cyanide and other ligands with the homogenous oxidized preparation of cytochrome bd. It was found that the addition of KCN to the fully oxidized cytochrome bd brings about some absorption changes in the visible range of the difference absorption spectrum (the 624-nm peak is most pronounced) (170). These changes suggest the reaction of the ligand with heme d. A typical bathochromic shift of the γ-band is also observed. While the changes in the visible region are virtually saturated at 0.5 mM KCN, the Soret band effect continues to grow, indicating a second low-affinity ligand-binding site (170). The MCD spectrum of the fully oxidized cytochrome bd is dominated by an asymmetric signal in the Soret region. Submillimolar cyanide has no effect on the initial MCD spectrum. KCN at 50 mM induces minor changes of the MCD signal in the Soret band, which can be modeled as the transition of a part of the low-spin heme b558 (15 to 20%) to its low-spin cyanocomplex (170). There is no evidence of the interaction of high-spin ferric heme b595 with the ligand. The apparent discrepancy between data on the interactions of cyanide with oxygenated (235) and fully oxidized (170) forms of the bd enzyme may derive from the fact that, in the former case, there seemed to be partial reduction of heme b595 associated with ligand-induced electron transfer from heme d rather than a change of the heme b595 spin state. On the basis of EPR spectra, Tsubaki et al. (182) proposed that the treatment of air-oxidized cytochrome bd with cyanide results in a cyanide-bridging species with a heme d3+—C=N—heme b5953+ structure. However, Tsubaki et al. (182) did not account for the electron released from heme d upon cyanide binding to as-prepared cytochrome bd.
Interaction with H2O2
The addition of excess H2O2 to E. coli membranes containing cytochrome bd (236) and the purified enzyme in the as-prepared (183, 209) or the fully oxidized (51, 183, 237) state gives rise to an absorption band at ∼680 nm. The reaction of H2O2 with fully oxidized cytochrome bd also induces a bathochromic shift of the γ-band (183, 237). H2O2 binds to ferric heme d with an apparent KD of 30 μM, but it seems not to interact with heme b595 (183, 237). The fully ferric cytochrome bd reacts with peroxide with kon of 600 M−1·s−1. The decay of the H2O2-induced spectral changes upon the addition of catalase (at a rate of ∼10−3 s−1) is about 20-fold slower than expected for the dissociation of H2O2 from the complex, with heme d assuming a simple reversible binding of peroxide (koff = KD × kon ∼ 2 × 10−2 s−1) (237). This finding suggests that the interaction of H2O2 with cytochrome bd is essentially irreversible, giving rise to the oxoferryl state of heme d (237). The assignment of compound 680 to the oxoferryl state of heme d is confirmed by resonance Raman spectroscopy data (160). The results of resonance Raman spectroscopy studies suggest that heme d is in the high-spin pentacoordinate state when it is oxygenated (Fe2+-O2) or exists as an oxoferryl species (Fe4+-O2−) or in a compound with cyanide (238).
Remarkably, the isolated cytochrome bd displays some peroxidase enzymatic activity (84). Moreover, cytochrome bd, either purified or overexpressed in a catalase-deficient E. coli strain, shows a significant catalase activity that is insensitive to NO (71, 85, 86). Therefore, one may conclude that this bd-type terminal oxidase actually contributes to protect E. coli from H2O2 stress.
PROPOSED MECHANISMS OF FUNCTIONING
Mechanism of Cytochrome bo3 Functioning
Cytochrome bo3 catalyzes the final step of E. coli respiration—the oxidation of ubiquinol-8 and the reduction of molecular oxygen. The reduction of one dioxygen molecule to water requires four electrons, which are supplied by two molecules of ubiquinol on the P side, and four protons, taken up from the N side of the membrane. The reduction of oxygen to water is an exergonic process, coupled with the release of large amounts of energy. This energy is conserved in the form of ΔμH+. The formation of ΔμH+ by cytochrome bo3 is based on two principles: vectorial chemistry and proton pumping. Since the protons and electrons for oxygen reduction to water are taken from different sides of the membrane, the reduction results in the net transfer of four charges across the membrane. At the same time, the enzyme is able not only to catalyze the oxygen reduction, but also to utilize the released energy for proton pumping. This process was discovered in 1977 (239) by the demonstration that the reduction of molecular oxygen to water by mitochondrial cytochrome c oxidase is linked to the pumping of additional four protons across the membrane dielectric (239). Later, such functional ability was shown for cytochrome bo3 (240). Hence, the overall reaction catalyzed by cytochrome bo3 can be described by the following equation:
where Q is ubiquinone, QH2 is ubiquinol, H+N is proton taken from the N side of the membrane (the cytosol), and H+P is proton released to the P side of the membrane (the periplasmic space). The mechanism coupling electron transfer reactions with the transmembrane proton translocation is still under debate. Let’s look through the main elements of this mechanism.
Electron transfer reactions in cytochrome bo3
The general sequence of electron transfer in the bo3 oxidase is well established. The ubiquinone molecule occupying the QH-binding site has a stable semiquinone form (241), which ensures the coupling of a one-electron redox reaction to a two-electron donor. A pulse radiolysis study showed that the quinone bound at the QH site is important for the rapid reduction of heme b but not for rapid electron transfer from heme b to the heme o3-CuB binuclear center (242). The rate constant of electron transfer between semiquinone and heme b was found to be 1.5 × 103 s−1 (242).
In the next step, the low-spin heme delivers electrons to the binuclear site in a controlled fashion that is coupled to proton translocation across the membrane (243). The rate of this electron transfer without coupling to the reaction with protons has been studied extensively, in particular, by the photodissociation of CO from the oxygen-binding heme in the so-called mixed valence form of the enzyme, where only the binuclear-site metals are initially reduced (244, 245). (Note that the term “mixed valence” means two-electron-reduced form for cytochrome bo3 but one-electron-reduced form for cytochrome bd.) Primarily, it was found that this rate is about 3 μs (3 × 105 s−1), which is too low to correspond to the pure electron tunneling, especially since this rate is independent of pH and substitution with heavy water (246, 247) and therefore may not be presumed to be linked to proton transfer. The obtained value is, however, about three orders of magnitude lower than the value predicted by the empirical electron transfer theory (248, 249). The results of recent femtosecond experiments show that the electron transfer from CO-dissociated ferrous heme o3 to the low-spin ferric heme b takes place at a rate of 8.3 × 108 s−1 (1.2 ns).
The final step of electron transfer reactions between heme o3 and CuB has not been a subject for experimental determination, but, from our modern understanding of electron transfer reactions, the rate of transfer between the two metal atoms in the binuclear catalytic site can be predicted to be in the order of picoseconds, taking into account the very small distance (∼5 Å) between these atoms.
The suggested electron transfer sequence in cytochrome bo3 can be described by the following equation:
Proton transfer reactions in cytochrome bo3
The proton transfer pathways in cytochrome bo3 have been investigated much less than the electron transfer pathways. There are no well-identified places for proton localization during the transfer of protons across the membrane, except the glutamate residue at position 286 (250, 251, 252) in the middle of the membrane. The quality of the X-ray structure of the bo3 oxidase was not sufficient to resolve individual water molecules in the proton-conducting channels. However, a very high level of structural homology to the other members of the heme-copper superfamily allows us to draw some conclusions about proton movements in the protein milieu including structural water arrays (253, 254) in two well-defined proton channels. Both of these channels serve as proton delivery pathways and cross about two-thirds of the membrane dielectric from the cytoplasmic side to the binuclear catalytic site. The functional separation of the channel is not yet completely clear. Originally, models of the proposed molecular mechanism of proton pumping predicted the existence of two proton-conductive structures with different functions. One channel was proposed to be responsible for the translocation of “pumped” protons and the other for the uptake of “chemical” protons for water formation (255, 256). The resolved structure revealed the existence of two such channels (named the D and K channels), and, originally (257), it was proposed that the D channel was responsible for the translocation of pumped protons and that the K channel was used for the uptake of chemical protons for water formation. However, more recent results indicate that the D channel is involved in the uptake not only of all four pumped protons, but also of two chemical protons used in the oxidative part of the catalytic cycle (258) and that the K channel is responsible for the uptake of another two chemical protons during the reductive part of the cycle (259, 260). The proton translocation mechanism also requires two proton exit channels—one for proton release upon the oxidation of bound ubiquinol and the other for the release of the pumped protons. The 3D structure did not show clear exit pathways. However, the exit pathways should be much shorter than the D and K channels, only one-third of the membrane dielectric. In addition, in the well-resolved structures of bovine (261) and Rhodobacter sphaeroides (254) aa3 oxidases, areas rich in structural water molecules, which can serve as exit channels, were found above the heme propionates. The results of site-directed mutagenesis studies suggested that the exit for pumped protons may start at conserved residues R481 and R482 (262), which are hydrogen bonded to the Δ-propionates of the hemes, and then continue further through the chains of mobile water molecules.
Cytochrome bo3 catalytic cycle
The catalytic cycle for the reduction of oxygen to water is a rather complex process. One enzyme turnover includes the delivery of four electrons and four protons to the catalytic site, the binding of oxygen in this site, the translocation of four protons across the membrane, and the release of the product (two water molecules). This complexity is the reason why the real-time measurement of a single catalytic cycle shows a large number of intermediate states of the enzyme.
The reaction starts from the very fast (kon = 1.6 × 108 M−1·s−1 [30, 263]) formation of an oxygen adduct, so-called compound A, which is characterized by an oxygen molecule bound to a high-spin heme o3, as in hemoglobin.
Unlike that of hemoglobin, the lifetime of this intermediate is very short. In 24 μs (263), the bound oxygen accepts electrons from CuB and heme b and a proton, which results in O—O bond splitting and the formation of a peroxy intermediate. This compound was named “peroxy” because it can be produced by an oxidase containing only two electrons per enzyme molecule, which formally corresponds to the reduction of O2 to peroxide. However, more recent examination by kinetic resonance Raman spectroscopy (264) and mass spectrometry (265) clearly demonstrated that the oxygen-oxygen bond is already broken and that heme o3 is in the oxoferryl state (Feo34+=O) with another oxygen atom being bound to CuB as a hydroxide ion. The “peroxy” state can also be generated directly in the reaction of the oxidized bo3 with H2O2 (266), and the resulting spectrum has a characteristic peak at 582 nm and a shoulder at 550 nm.
Upon the arrival of the next electron and the proton, the “peroxy” form (30) is converted into a ferryl intermediate. The results of time-resolved resonance Raman studies showed the formation of a ferryl intermediate with a rate constant of about 2 × 104 s−1 (267, 268). A very similar rate was also detected by visible spectroscopy (30, 269, 270). The stable ferryl intermediate can be obtained as the end product in the reaction of the fully reduced enzyme with oxygen when no bound ubiquinone is in the QH site (269). Such an enzyme contains only three of the four electrons required for complete O2 reduction, and so the catalytic cycle stalls at the ferryl state. In the visible spectroscopy, this state is characterized by a spectrum with peaks at 557 and about 420 nm.
The presence of bound ubiquinone at the QH site of cytochrome bo3 (150) increases the electron capacitance of the enzyme and allows the reaction to proceed further to yield the oxidized form. The rate of formation of the oxidized state in the single-turnover experiments was estimated to be 0.3 × 103 s−1 by recording the electron and proton transfer reactions (271).
Mechanism of ΔμH+ formation by cytochrome bo3
The oxidation of ubiquinol (Em of redox couple Q/QH2 ∼ +0.1 V) by oxygen (Em of redox couple O2/H2O ∼ +0.8 V) is linked with free energy release in the order of ∼0.7 V. The proton motive force created by the respiratory chain on the E. coli cellular membrane is about −0.2 V (272, 273). It is clear that excess free energy (∼0.5 V) can be used for the translocation of more than one charge across the membrane. Indeed, measurement of the stoichiometry (28) of the proton transfer by the bo3 oxidase on the E. coli cellular membrane showed that two protons are transferred per each electron used for the O2 reduction. Of these two charges, the first is driven by the vectorial chemistry and the second is driven by the proton pump (240). The vectorial chemistry includes the oxidation of the ubiquinol molecule and the reduction of dioxygen, with the release and uptake, respectively, of one proton per electron. The charge separation in this case is achieved by the separation in the space of these two events.
The protons extracted from QH2 owing to its oxidation by heme b are released to the P side, and the protons for the O2 reduction in the binuclear site are taken up from the N side of the membrane. Because heme b and the binuclear site are located at the same depth in the membrane, the release and uptake of the protons correspond to the translocation of one complete charge across the membrane. A second charge is transferred by the proton pump, the mechanism of which is still unclear. Several proposals for the proton translocation machinery of the heme-copper oxidases have been presented during the past 30 years since the discovery of this machinery in mitochondria (239) and bacteria (240, 274), but a definitive explanation of how it functions has not yet been presented. A high level of structural homology between different members of the heme-copper oxidase superfamily argues for similarity in the molecular mechanism of the transmembrane proton translocation. For cytochrome c oxidase, it was shown that, in the continuous-turnover regimen, the catalytic cycle consists of four sequential proton translocation steps (275, 276). During each step, the delivery of an electron to the binuclear site initiates a proton pump cycle, which is likely to occur by essentially the same mechanism every time when an electron arrives. There are a number of models explaining how the coupling between electron transfer and proton translocation occurs (277, 278, 279, 280, 281). Despite the different postulations regarding the drivers of these processes, the models are quite similar and employ electrostatic coupling between the movements of electrons and protons in the low-dielectric medium of the membrane protein. The delivery of an electron to the binuclear center drives proton pumping across the hydrophobic barrier. This translocation is followed by the uptake of the chemical proton to the active site, which leads to the release of the pumped proton out of the protein at the P side.
Mechanism of Cytochrome bd Functioning
Under physiological conditions, cytochrome bd can oxidize ubiquinol-8 and menaquinol-8. In vitro, the bd enzyme can also utilize shorter-chained ubiquinols, menadiol, duroquinol, and artificial electron donors such as N,N,N′,N′-tetramethyl-p-phenylendiamine (TMPD) (in the presence of excess ascorbate). Of the in vitro substrates, ubiquinol-1 (plus excess dithiothreitol) shows the highest turnover numbers (34, 198). The apparent Km values for some reductants are shown in Table 1. As noted above, the activity of the purified bd enzyme strongly depends on the nature of the detergent in which the enzyme is solubilized. Cytochrome bd is inactive in octylglucoside or cholate but shows high activity in Tween 20 or Triton X-100 (198) or N-lauroyl-sarcosine (73). The ubiquinol-1 oxidase activity of cytochrome bd has a broad optimum above pH 7.5 but decreases at more acidic pH values (198). Cytochrome bd possesses three distinct active sites—for quinol oxidation, TMPD oxidation, and oxygen reduction. All three sites seem to be located at or close to the periplasmic surface of the membrane. Electrons donated from quinol transfer to heme b558 and then to the b595-d diheme site, whereas electrons donated from TMPD transfer directly to the b595-d site, bypassing the quinol-binding site and heme b558 (165).
Mechanism of ΔμH+ formation by cytochrome bd
Cytochrome bd was shown to generate a transmembrane electric potential both in single turnover (51, 52, 53, 54) and under multiple-turnover conditions (32, 50, 282) but without invoking a proton pump (H+/e− ratio, ∼1 [28, 29, 106]; q/e− ratio [the number of charges translocated across the membrane per electron], ∼1 [256]). When reconstituted into liposomes, cytochrome bd generates an uncoupler-sensitive transmembrane voltage difference with a value of 160 to 180 mV (negative inside) (32, 50). The ubi(mena)quinol molecule generated by the dehydrogenases of the respiratory chain can diffuse laterally within the bilayer, finding its way into the quinol oxidase site located near the outer side of the membrane. Upon the oxidation of a quinol, two protons are released into the periplasmic space and two electrons are transferred through heme b558 to the heme b595-heme d oxygen-reducing site also located near the periplasmic surface of the membrane. The four protons used for O2 reduction are taken up from the cytoplasm. Single-turnover electrometric experiments showed that membrane potential generation is associated with electron transfer from heme b558 to the b595-d active site (51, 52, 53, 54). However, since all three hemes are likely to be located at the same depth of the membrane, close to the periplasmic side (176, 177), the electron transfer itself cannot be electrogenic (51). Rather, vectorial proton movement from the cytoplasm toward the active site on the opposite (periplasmic) side of the membrane, coupled with a redox reaction between hemes b and d, must occur (51, 52, 53, 54). The latter means that there must be a proton-conducting channel connecting the cytoplasm to the b595-d active site (51, 52, 54) (Fig. 6). Thus, the generated potential must result primarily from protons moving from the cytoplasm to the O2-reducing site on the opposite side of the membrane.
Reaction of cytochrome bd with oxygen: sequence of catalytic intermediates
For cytochrome bd, the following species were detected: fully ferrous (32, 47), fully ferric (193, 209), MV O2 bound (193, 208, 209, 210, 283), fully reduced O2 bound (AR) (51), oxoferryl (160, 283), and peroxy (53). AR and the peroxy intermediate are transient, whereas the others can be generated in relatively stable forms. In addition, a recent stopped-flow multiwavelength absorption spectroscopy study of the isolated cytochrome bd at steady state revealed one electron-reduced species with the electron being on heme b558 (283).
The reaction of the fully ferrous cytochrome bd with oxygen was studied with the use of the flow-flash method (284) by means of spectroscopic and electrometric techniques (51, 52, 53, 54). Recording absorption spectra and membrane potential development with 1-μs time resolution allowed us (i) to observe the sequence of the catalytic intermediates and (ii) to establish which catalytic steps are linked to electric potential generation (53).
The scheme describing the reaction is shown in Fig. 7. The initial complex of the fully ferrous cytochrome bd with CO is photolyzed in the presence of oxygen. The unligated ferrous enzyme generated by the CO photolysis binds O2 very rapidly, forming the ferrous heme d oxyspecies (AR). The transition of the unligated fully ferrous enzyme to AR is not electrogenic, and its rate is proportional to [O2] (kon = 1.9 × 109 M−1·s−1 [53, 161]). The AR formation is followed by electron transfer from heme b595 to form the peroxy intermediate. The AR→peroxy transition occurs at a rate of 2.2 × 105 s−1 (4.5 μs) and is also nonelectrogenic (53). Thus, electron transfer from heme b595 to heme d is not coupled with membrane potential generation (52, 53). The peroxy intermediate might be a true peroxy complex of ferric heme d (53). If this is the case, the bound peroxide is likely not to be in the anionic form but at least singly protonated. The proton may come from one of the two protonatable groups linked to the b595-d diheme site upon its oxidation (52). Alternatively, the peroxy intermediate could be an oxoferryl species with a π-cation radical on the porphyrin ring of heme d and one electron on heme b558 (285, 286). The peroxy intermediate is further converted into the ferryl intermediate at a rate of 2.1 × 104 s−1 (48 μs). This conversion is accompanied by the oxidation of heme b558. The formation of the ferryl intermediate is coupled to the generation of a membrane potential (51, 52, 53, 54). At the ferryl intermediate stage, the b-type hemes are in a ferric state and heme d is in an oxoferryl state. When cytochrome bd contains bound quinol, the reaction proceeds further to form the oxidized enzyme. The transition from the ferryl intermediate to the oxidized enzyme occurs at a rate of 0.9 × 103 s−1 (1.1 ms) and is electrogenic (52, 53, 54).
Figure 7.

Cytochrome bd reaction scheme. The three rhombuses represent hemes b558, b595, and d, respectively. The minus signs and red backgrounds in the rhombuses denote that the heme is in the ferrous state. R-CO, R, AR, P, F, O are fully ferrous CO bound, fully ferrous unligated, fully ferrous O2 bound, peroxy, oxoferryl, and fully ferric species, respectively. Transient peroxy species (P) discovered by Belevich et al. (53) is shown as a true peroxy complex of ferric heme d. It is possible, however, that P is an oxoferryl form with a π-cation radical on the porphyrin ring of heme d (285, 286). doi:10.1128/ecosalplus.ESP-0012-2015.f7
Under turnover conditions, the oxoferryl and MV O2-bound species dominate, with a small fraction of cytochrome bd containing ferric heme d and one electron on heme b558 (283). The fully oxidized species is possibly not part of the normal catalytic cycle of cytochrome bd (287).
Role of heme b595
Upon the addition of a ligand (e.g., CO or cyanide) to cytochrome bd, most of heme b595 does not bind the ligand (170, 172, 184, 187, 188, 189). It is likely that heme b595, although in the high-spin pentacoordinate state, is resistant to interaction with the classical ligands of the high-spin iron-porphyrin complexes. It cannot be excluded that despite the high-spin pentacoordinate state of the iron-porphyrin group, the specific features of the protein environment are such that this redox cofactor is protected from interaction with ligands. In such a case, the participation of heme b595 in O2 reduction in cooperation with heme d is unlikely and the role of heme b595 is limited to the transfer of an electron to heme d. Another possible explanation that we favor is the following.
Both heme b595 and heme d potentially can bind ligands.
The hemes are located very close to each other, forming a diheme active site.
The spatial proximity of hemes b595 and d results in steric restrictions, allowing such a diheme site to bind only one ligand molecule.
Heme d has a higher affinity for ligands than heme b595, in which case the final result observed upon the addition of a ligand will always be the binding of the ligand to heme d, whereas heme b595 will remain mainly in the free state (170, 183, 184, 189).
The data on the redox coupling of the two hemes to the same ionizable groups (52) and the migration of CO within the protein from heme d to heme b595 at cryogenic temperatures (181) are in agreement with this proposal. Modeling the excitonic interactions in absorption and CD spectra of cytochrome bd yields an estimate of the Fed-to-Feb595 distance of about 10 Å (171). This distance is markedly larger than that between the Fe-CuB pair in heme-copper oxidases (4 to 5 Å). If this is the case, heme b595 cannot be functionally identical to CuB. A possible role of heme b595, apart from electron delivery to heme d and/or to heme d-bound oxygen intermediates, would be as a binding site for hydroxide produced from heme d-bound oxygen upon the reductive cleavage of the O—O bond (171).
CYTOCHROME bd-II (appBCX)
Originally, it had been reported that a second cytochrome bd in E. coli (cytochrome bd-II) is encoded by appBC genes (also have been called cyxAB or cbdAB) (288). The appBC genes, located at 22 min on the E. coli genetic map, are upstream from a pH 2.5 acid phosphatase (appA) gene (288). The appBC and appA genes constitute the complex operon. The appB and appC genes encode 58.1- and 42.4-kDa integral membrane proteins, respectively. The deduced amino acid sequences of appB and appC gene products show homologies of 60 and 57%, respectively, to the sequences of subunit I (CydA) and subunit II (CydB) of the major cytochrome bd (encoded by cydABX) (288). A mutant lacking the cyo and cyd operons and a functional appC gene is unable to grow aerobically in rich medium, in contrast to a mutant lacking only the cyo and cyd operons, suggesting that appBC genes encode a third terminal oxidase in E. coli (288). The expression of the appBC-appA operon is induced by entry into the stationary phase and phosphate starvation (289). The appBC genes are also induced by anaerobic growth, and this induction is controlled by transcriptional regulators AppY and ArcA but is independent of Fnr, in contrast to the induction of the cyd operon (289, 290). Only very recently it has become clear that cytochrome bd-II, like the major cytochrome bd, constitutes the third subunit, appX, encoded by the appX gene (21, 43, 109). Thus, in E. coli cytochrome bd-II is a trisubunit protein encoded by the appCBX locus (21, 43, 109). Cytochrome bd-II is likely to function under even more oxygen-limiting conditions than cydABX-encoded cytochrome bd (290). Cytochrome bd-II has been extracted and purified (29, 291). With the use of sodium dodecyl sulfate-polyacrylamide gel electrophoresis, two polypeptides, 43 kDa (subunit I) and 27 kDa (subunit II), were resolved from the preparation (291). These subunits show no cross-reactivity to subunit-specific polyclonal antibodies directed against the two major subunits of cydABX-encoded cytochrome bd (291). The spectral properties of cytochrome bd-II closely resemble those of cydABX-encoded cytochrome bd. Of the quinols tested in the work of Sturr et al. (291), cytochrome bd-II utilizes menadiol as the preferred substrate (although ubiquinol-1, the most efficient in vitro substrate for cydABX-encoded cytochrome bd, was not examined in that study). The TMPD oxidase activity of cytochrome bd-II is much more sensitive to cyanide than that of cydABX-encoded cytochrome bd (291). The electron flux through cytochrome bd-II appeared to be significant (292, 293). Bekker et al. reported that cytochrome bd-II does not contribute to the production of the proton motive force (H+/e− = 0) (292). That conclusion was made based on the growth properties of the strain that lacks NDH-I, cytochrome bo3 and cytochrome bd under glucose-limited conditions in continuous culture (292). Later, however, Borisov et al., using cytochrome bd from the same strain, directly measured generation of a proton motive force (29). Intact cells, spheroplasts, membrane vesicles, and the purified enzyme were tested (29). The authors showed clearly that the catalytic turnover of cytochrome bd-II does produce a proton motive force with the same energetic efficiency as cydABX-encoded cytochrome bd (H+/e− = 1) (29). Reexamining the data reported in reference 292, Sharma et al. (294) confirmed the H+/e− stoichiometry of 1, determined by Borisov et al. (29). Nonetheless, cytochrome bd-II still remains poorly studied.
ACKNOWLEDGMENTS
Studies in our laboratories were supported by the Biocentrum Helsinki, the Sigrid Jusélius Foundation, a grant from the Academy of Finland (to M.I.V.), and the Russian Foundation for Basic Research grants 08-04-00093, 11-04-00031, 14-04-00153, and 15-04-06266 (to V.B.B.).
REFERENCES
- 1.Gennis RB, Stewart V. 1996. Respiration, p 217–261. In Neidhardt FC, Curtiss R III, Ingraham JL, Lin ECC, Low KB, Magasanik B, Reznikoff WS, Riley M, Schaechter M, Umbarger HE (ed), Escherichia coli and Salmonella: Cellular and Molecular Biology, 2nd ed. ASM Press, Washington, DC. [Google Scholar]
- 2.Unden G, Bongaerts J. 1997. Alternative respiratory pathways of Escherichia coli: energetics and transcriptional regulation in response to electron acceptors. Biochim Biophys Acta 1320:217–234. [PubMed] 10.1016/S0005-2728(97)00034-0 [DOI] [PubMed] [Google Scholar]
- 3.Garcia-Horsman JA, Barquera B, Rumbley J, Ma J, Gennis RB. 1994. The superfamily of heme-copper respiratory oxidases. J Bacteriol 176:5587–5600. [PubMed] [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Pereira MM, Santana M, Teixeira M. 2001. A novel scenario for the evolution of haem-copper oxygen reductases. Biochim Biophys Acta 1505:185–208. [PubMed] 10.1016/S0005-2728(01)00169-4 [DOI] [PubMed] [Google Scholar]
- 5.Brunori M, Giuffrè A, Sarti P. 2005. Cytochrome c oxidase, ligands and electrons. J Inorg Biochem 99:324–336. [PubMed] 10.1016/j.jinorgbio.2004.10.011 [DOI] [PubMed] [Google Scholar]
- 6.Azarkina N, Borisov V, Konstantinov AA. 1997. Spontaneous spectral changes of the reduced cytochrome bd. FEBS Lett 416:171–174. [PubMed] 10.1016/S0014-5793(97)01196-4 [DOI] [PubMed] [Google Scholar]
- 7.Azarkina N, Siletsky S, Borisov V, von Wachenfeldt C, Hederstedt L, Konstantinov AA. 1999. A cytochrome bb’-type quinol oxidase in Bacillus subtilis strain 168. J Biol Chem 274:32810–32817. [PubMed] 10.1074/jbc.274.46.32810 [DOI] [PubMed] [Google Scholar]
- 8.Gavrikova EV, Grivennikova VG, Borisov VB, Cecchini G, Vinogradov AD. 2009. Assembly of a chimeric respiratory chain from bovine heart submitochondrial particles and cytochrome bd terminal oxidase of Escherichia coli. FEBS Lett 583:1287–1291. [PubMed] 10.1016/j.febslet.2009.03.022 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Siletsky SA, Konstantinov AA. 2012. Cytochrome c oxidase: charge translocation coupled to single-electron partial steps of the catalytic cycle. Biochim Biophys Acta 1817:476–488. [PubMed] 10.1016/j.bbabio.2011.08.003 [DOI] [PubMed] [Google Scholar]
- 10.Siletsky SA. 2013. Steps of the coupled charge translocation in the catalytic cycle of cytochrome c oxidase. Front Biosci 18:36–57. [PubMed] 10.2741/4086 [DOI] [PubMed] [Google Scholar]
- 11.Borisov VB. 2002. Defects in mitochondrial respiratory complexes III and IV, and human pathologies. Mol Aspects Med 23:385–412. [PubMed] 10.1016/S0098-2997(02)00013-4 [DOI] [PubMed] [Google Scholar]
- 12.Borisov VB. 2004. Mutations in respiratory chain complexes and human diseases. Ital J Biochem 53:34–40. [PubMed] [PubMed] [Google Scholar]
- 13.Szundi I, Kittredge C, Choi SK, McDonald W, Ray J, Gennis RB, Einarsdottir O. 2014. Kinetics and intermediates of the reaction of fully reduced Escherichia coli bo3 ubiquinol oxidase with O2. Biochemistry 53:5393–5404. [PubMed] 10.1021/bi500567m [DOI] [PubMed] [Google Scholar]
- 14.de Gier J-WL, Lübben M, Reijnders WNM, Tipker CA, van Spanning RJM, Stouthamer AH, van der Oost J. 1994. The terminal oxidase of Paracoccus denitrificans. Mol Microbiol 13:183–196. [PubMed] 10.1111/j.1365-2958.1994.tb00414.x [DOI] [PubMed] [Google Scholar]
- 15.Richardson DJ. 2000. Bacterial respiration: a flexible process for a changing environment. Microbiology 146:551–571. [PubMed] 10.1099/00221287-146-3-551 [DOI] [PubMed] [Google Scholar]
- 16.Bertsova YV, Demin OV, Bogachev AV. 2005. Respiratory protection of nitrogenase complex in Azotobacter vinelandii. Uspekhi Biolog Khim 45:205–234. (in Russian) [Google Scholar]
- 17.Poole RK, Cook GM. 2000. Redundancy of aerobic respiratory chains in bacteria? Routes, reasons and regulation. Adv Microb Physiol 43:165–224. [PubMed] 10.1016/S0065-2911(00)43005-5 [DOI] [PubMed] [Google Scholar]
- 18.Anraku Y, Gennis RB. 1987. The aerobic respiratory chain of Escherichia coli. Trends Biochem Sci 12:262–266. 10.1016/0968-0004(87)90131-9 [DOI] [Google Scholar]
- 19.Mogi T, Tsubaki M, Hori H, Miyoshi H, Nakamura H, Anraku Y. 1998. Two terminal quinol oxidase families in Escherichia coli: variations on molecular machinery for dioxygen reduction. J Biochem Mol Biol Biophys 2:79–110. [Google Scholar]
- 20.Borisov VB, Gennis RB, Hemp J, Verkhovsky MI. 2011. The cytochrome bd respiratory oxygen reductases. Biochim Biophys Acta 1807:1398–1413. [PubMed] 10.1016/j.bbabio.2011.06.016 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Holyoake LV, Poole RK, Shepherd M. 2015. The CydDC family of transporters and their roles in oxidase assembly and homeostasis. Adv Microb Physiol 66:1–53. [PubMed] 10.1016/bs.ampbs.2015.04.002 [DOI] [PubMed] [Google Scholar]
- 22.Rice CW, Hempfling WP. 1978. Oxygen-limited continuous culture and respiratory energy conservation in Escherichia coli. J Bacteriol 134:115–124. [PubMed] [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Cotter PA, Chepuri V, Gennis RB, Gunsalus RP. 1990. Cytochrome o (cyoABCDE) and d (cydAB) oxidase gene expression in Escherichia coli is regulated by oxygen, pH, and the fnr gene product. J Bacteriol 172:6333–6338. [PubMed] [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Fu H-A, Iuchi S, Lin ECC. 1991. The requirement of ArcA and Fnr for peak expression of the cyd operon in Escherichia coli under microaerobic conditions. Mol Gen Genet 226:209–213. [PubMed] 10.1007/BF00273605 [DOI] [PubMed] [Google Scholar]
- 25.Sharma P, Teixeira de Mattos MJ, Hellingwerf KJ, Bekker M. 2012. On the function of the various quinone species in Escherichia coli. FEBS J 279:3364–3373. [PubMed] 10.1111/j.1742-4658.2012.08608.x [DOI] [PubMed] [Google Scholar]
- 26.Zhang J, Oettmeier W, Gennis RB, Hellwig P. 2002. FTIR spectroscopic evidence for the involvement of an acidic residue in quinone binding in cytochrome bd from Escherichia coli. Biochemistry 41:4612–4617. [PubMed] 10.1021/bi011784b [DOI] [PubMed] [Google Scholar]
- 27.Yang K, Zhang J, Vakkasoglu AS, Hielscher R, Osborne JP, Hemp J, Miyoshi H, Hellwig P, Gennis RB. 2007. Glutamate 107 in subunit I of the cytochrome bd quinol oxidase from Escherichia coli is protonated and near the heme d/heme b595 binuclear center. Biochemistry 46:3270–3278. [PubMed] 10.1021/bi061946+ [DOI] [PubMed] [Google Scholar]
- 28.Puustinen A, Finel M, Haltia T, Gennis RB, Wikström M. 1991. Properties of the two terminal oxidases of Escherichia coli. Biochemistry 30:3936–3942. [PubMed] 10.1021/bi00230a019 [DOI] [PubMed] [Google Scholar]
- 29.Borisov VB, Murali R, Verkhovskaya ML, Bloch DA, Han H, Gennis RB, Verkhovsky MI. 2011. Aerobic respiratory chain of Escherichia coli is not allowed to work in fully uncoupled mode. Proc Natl Acad Sci USA 108:17320–17324. [PubMed] 10.1073/pnas.1108217108 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Svensson M, Nilsson T. 1993. Flow-flash study of the reaction between cytochrome bo and oxygen. Biochemistry 32:5442–5447. [PubMed] 10.1021/bi00071a021 [DOI] [PubMed] [Google Scholar]
- 31.Belevich I, Borisov VB, Konstantinov AA, Verkhovsky MI. 2005. Oxygenated complex of cytochrome bd from Escherichia coli: stability and photolability. FEBS Lett 579:4567–4570. [PubMed] 10.1016/j.febslet.2005.07.011 [DOI] [PubMed] [Google Scholar]
- 32.Kita K, Konishi K, Anraku Y. 1984. Terminal oxidases of Escherichia coli aerobic respiratory chain. II. Purification and properties of cytochrome b558-d complex from cells grown with limited oxygen and evidence of branched electron-carrying systems. J Biol Chem 259:3375–3381. [PubMed] [PubMed] [Google Scholar]
- 33.Kolonay JF, Jr., Moshiri F, Gennis RB, Kaysser TM, Maier RJ. 1994. Purification and characterization of the cytochrome bd complex from Azotobacter vinelandii: comparison to the complex from Escherichia coli. J Bacteriol 176:4177–4181. [PubMed] [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Jünemann S, Butterworth PJ, Wrigglesworth JM. 1995. A suggested mechanism for the catalytic cycle of cytochrome bd terminal oxidase based on kinetic analysis. Biochemistry 34:14861–14867. [PubMed] 10.1021/bi00045a029 [DOI] [PubMed] [Google Scholar]
- 35.D’mello R, Hill S, Poole RK. 1996. The cytochrome bd quinol oxidase in Escherichia coli has an extremely high oxygen affinity and two-oxygen-binding haems: implications for regulation of activity in vivo by oxygen inhibition. Microbiology 142:755–763. [PubMed] 10.1099/00221287-142-4-755 [DOI] [PubMed] [Google Scholar]
- 36.Verkhovsky MI, Morgan JE, Puustinen A, Wikström M. 1996. Kinetic trapping of oxygen in cell respiration. Nature 380:268–270. [PubMed] 10.1038/380268a0 [DOI] [PubMed] [Google Scholar]
- 37.D’Mello R, Hill S, Poole RK. 1995. The oxygen affinity of cytochrome bo’ in Escherichia coli determined by the deoxygenation of oxyleghemoglobin and oxymyoglobin: Km values for oxygen are in the submicromolar range. J Bacteriol 177:867–870. [PubMed] [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Kita K, Konishi K, Anraku Y. 1984. Terminal oxidases of Escherichia coli aerobic respiratory chain. I. Purification and properties of cytochrome b562-o complex from cells in the early exponential phase of aerobic growth. J Biol Chem 259:3368–3374. [PubMed] [PubMed] [Google Scholar]
- 39.Calhoun MW, Newton G, Gennis RB. 1991. E. coli map. Physical map locations of genes encoding components of the aerobic respiratory chain of Escherichia coli. J Bacteriol 173:1569–1570. [PubMed] [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Kranz RG, Barassi CA, Miller MJ, Green GN, Gennis RB. 1983. Immunological characterization of an E. coli strain which is lacking cytochrome d. J Bacteriol 156:115–121. [PubMed] [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Green GN, Fang H, Lin R-J, Newton G, Mather M, Georgiou CD, Gennis RB. 1988. The nucleotide sequence of the cyd locus encoding the two subunits of the cytochrome d terminal oxidase complex of Escherichia coli. J Biol Chem 263:13138–13143. [PubMed] [PubMed] [Google Scholar]
- 42.VanOrsdel CE, Bhatt S, Allen RJ, Brenner EP, Hobson JJ, Jamil A, Haynes BM, Genson AM, Hemm MR. 2013. The Escherichia coli CydX protein is a member of the CydAB cytochrome bd oxidase complex and is required for cytochrome bd oxidase activity. J Bacteriol 195:3640–3650. [PubMed] 10.1128/JB.00324-13 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Hoeser J, Hong S, Gehmann G, Gennis RB, Friedrich T. 2014. Subunit CydX of Escherichia coli cytochrome bd ubiquinol oxidase is essential for assembly and stability of the di-heme active site. FEBS Lett 588:1537–1541. [PubMed] 10.1016/j.febslet.2014.03.036 [DOI] [PubMed] [Google Scholar]
- 44.Au DC-T, Gennis RB. 1987. Cloning of the cyo locus encoding the cytochrome o terminal oxidase complex of Escherichia coli. J Bacteriol 169:3237–3242. [PubMed] [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Chepuri V, Lemieux LJ, Au DC-T, Gennis RB. 1990. The sequence of the cyo operon indicates substantial structural similarities between the cytochrome o ubiquinol oxidase of Escherichia coli and the aa3-type family of the cytochrome c oxidases. J Biol Chem 265:11185–11192. [PubMed] [PubMed] [Google Scholar]
- 46.Minghetti KC, Goswitz VC, Gabriel NE, Hill JJ, Barassi C, Georgiou CD, Chan SI, Gennis RB. 1992. Modified, large-scale purification of the cytochrome o complex of Escherichia coli yields a two heme/one copper terminal oxidase with high specific activity. Biochemistry 31:6917–6924. [PubMed] 10.1021/bi00145a008 [DOI] [PubMed] [Google Scholar]
- 47.Miller MJ, Gennis RB. 1983. The purification and characterization of the cytochrome d terminal oxidase complex of the Escherichia coli aerobic respiratory chain. J Biol Chem 258:9159–9165. [PubMed] [PubMed] [Google Scholar]
- 48.Salerno JC, Bolgiano B, Poole RK, Gennis RB, Ingledew WJ. 1990. Heme-copper and heme-heme interactions in the cytochrome bo-containing quinol oxidase of Escherichia coli. J Biol Chem 265:4364–4368. [PubMed] [PubMed] [Google Scholar]
- 49.Belevich I, Borisov VB, Bloch DA, Konstantinov AA, Verkhovsky MI. 2007. Cytochrome bd from Azotobacter vinelandii: evidence for high-affinity oxygen binding. Biochemistry 46:11177–11184. [PubMed] 10.1021/bi700862u [DOI] [PubMed] [Google Scholar]
- 50.Miller MJ, Gennis RB. 1985. The cytochrome d complex is a coupling site in the aerobic respiratory chain of Escherichia coli. J Biol Chem 260:14003–14008. [PubMed] [PubMed] [Google Scholar]
- 51.Jasaitis A, Borisov VB, Belevich NP, Morgan JE, Konstantinov AA, Verkhovsky MI. 2000. Electrogenic reactions of cytochrome bd. Biochemistry 39:13800–13809. [PubMed] 10.1021/bi001165n [DOI] [PubMed] [Google Scholar]
- 52.Belevich I, Borisov VB, Zhang J, Yang K, Konstantinov AA, Gennis RB, Verkhovsky MI. 2005. Time-resolved electrometric and optical studies on cytochrome bd suggest a mechanism of electron-proton coupling in the di-heme active site. Proc Natl Acad Sci USA 102:3657–3662. [PubMed] 10.1073/pnas.0405683102 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Belevich I, Borisov VB, Verkhovsky MI. 2007. Discovery of the true peroxy intermediate in the catalytic cycle of terminal oxidases by real-time measurement. J Biol Chem 282:28514–28519. [PubMed] 10.1074/jbc.M705562200 [DOI] [PubMed] [Google Scholar]
- 54.Borisov VB, Belevich I, Bloch DA, Mogi T, Verkhovsky MI. 2008. Glutamate 107 in subunit I of cytochrome bd from Escherichia coli is part of a transmembrane intraprotein pathway conducting protons from the cytoplasm to the heme b595/heme d active site. Biochemistry 47:7907–7914. [PubMed] 10.1021/bi800435a [DOI] [PubMed] [Google Scholar]
- 55.Bader M, Muse W, Ballou DP, Gassner C, Bardwell JCA. 1999. Oxidative protein folding is driven by the electron transport system. Cell 98:217–227. [PubMed] 10.1016/S0092-8674(00)81016-8 [DOI] [PubMed] [Google Scholar]
- 56.Hill S, Viollet S, Smith AT, Anthony C. 1990. Roles for enteric d-type cytochrome oxidase in N2 fixation and microaerobiosis. J Bacteriol 172:2071–2078. [PubMed] [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Jones SA, Chowdhury FZ, Fabich AJ, Anderson A, Schreiner DM, House AL, Autieri SM, Leatham MP, Lins JJ, Jorgensen M, Cohen PS, Conway T. 2007. Respiration of Escherichia coli in the mouse intestine. Infect Immun 75:4891–4899. [PubMed] 10.1128/IAI.00484-07 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Avetisyan AV, Bogachev AV, Murtasina RA, Skulachev VP. 1992. Involvement of a d-type oxidase in the Na+-motive respiratory chain of Escherichia coli growing under low ΔμH+ conditions. FEBS Lett 306:199–202. [PubMed] 10.1016/0014-5793(92)80999-W [DOI] [PubMed] [Google Scholar]
- 59.Wall D, Delaney JM, Fayet O, Lipinska B, Yamamoto T, Georgopoulos C. 1992. arc-Dependent thermal regulation and extragenic suppression of the Escherichia coli cytochrome d operon. J Bacteriol 174:6554–6562. [PubMed] [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Delaney JM, Wall D, Georgopoulos C. 1993. Molecular characterization of the Escherichia coli htrD gene: Cloning, sequence, regulation, and involvement with cytochrome d oxidase. J Bacteriol 175:166–175. [PubMed] [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Bogachev AV, Murtazina RA, Skulachev VP. 1993. Cytochrome d induction in Escherichia coli growing under unfavorable conditions. FEBS Lett 336:75–78. [PubMed] 10.1016/0014-5793(93)81612-4 [DOI] [PubMed] [Google Scholar]
- 62.Bogachev AV, Murtazina RA, Shestopalov AI, Skulachev VP. 1995. Induction of the Escherichia coli cytochrome d by low ΔμH+ and by sodium ions. Eur J Biochem 232:304–308. [PubMed] 10.1111/j.1432-1033.1995.tb20812.x [DOI] [PubMed] [Google Scholar]
- 63.Ashcroft JR, Haddock BA. 1975. Synthesis of alternative membrane-bound redox carriers during aerobic growth of Escherichia coli in the presence of potassium cyanide. Biochem J 148:349–352. [PubMed] 10.1042/bj1480349 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Macinga DR, Rather PN. 1996. aarD, a Providencia stuartii homologue of cydD : role in 2′-N-acetyltransferase expression, cell morphology and growth in the presence of an extracellular factor. Mol Microbiol 19:511–520. 10.1046/j.1365-2958.1996.385912.x [DOI] [PubMed] [Google Scholar]
- 65.Cook GM, Loder C, Soballe B, Stafford GP, Membrillo-Hernandez J, Poole RK. 1998. A factor produced by Escherichia coli K-12 inhibits the growth of E. coli mutants defective in the cytochrome bd quinol oxidase complex: enterochelin rediscovered. Microbiology 144:3297–3308. [PubMed] 10.1099/00221287-144-12-3297 [DOI] [PubMed] [Google Scholar]
- 66.Siegele DA, Kolter R. 1993. Isolation and characterization of an Escherichia coli mutant defective in resuming growth after starvation. Genes Dev 7:2629–2640. [PubMed] 10.1101/gad.7.12b.2629 [DOI] [PubMed] [Google Scholar]
- 67.Siegele DA, Imlay KR, Imlay JA. 1996. The stationary-phase-exit defect of cydC (surB ) mutants is due to the lack of a functional terminal cytochrome oxidase. J Bacteriol 178:6091–6096. [PubMed] [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Forte E, Borisov VB, Konstantinov AA, Brunori M, Giuffrè A, Sarti P. 2007. Cytochrome bd, a key oxidase in bacterial survival and tolerance to nitrosative stress. Ital J Biochem 56:265–269. [PubMed] [PubMed] [Google Scholar]
- 69.Giuffrè A, Borisov VB, Mastronicola D, Sarti P, Forte E. 2012. Cytochrome bd oxidase and nitric oxide: From reaction mechanisms to bacterial physiology. FEBS Lett 586:622–629. [PubMed] 10.1016/j.febslet.2011.07.035 [DOI] [PubMed] [Google Scholar]
- 70.Giuffrè A, Borisov VB, Arese M, Sarti P, Forte E. 2014. Cytochrome bd oxidase and bacterial tolerance to oxidative and nitrosative stress. Biochim Biophys Acta 1837:1178–1187. [PubMed] 10.1016/j.bbabio.2014.01.016 [DOI] [PubMed] [Google Scholar]
- 71.Borisov VB, Forte E, Siletsky SA, Arese M, Davletshin AI, Sarti P, Giuffrè A. 2015. Cytochrome bd protects bacteria against oxidative and nitrosative stress: a potential target for next-generation antimicrobial agents. Biochemistry (Mosc) 80:565–575. [PubMed] 10.1134/S0006297915050077 [DOI] [PubMed] [Google Scholar]
- 72.Borisov VB, Forte E, Siletsky SA, Sarti P, Giuffrè A. 2015. Cytochrome bd from Escherichia coli catalyzes peroxynitrite decomposition. Biochim Biophys Acta 1847:182–188. [PubMed] 10.1016/j.bbabio.2014.10.006 [DOI] [PubMed] [Google Scholar]
- 73.Borisov VB, Forte E, Konstantinov AA, Poole RK, Sarti P, Giuffrè A. 2004. Interaction of the bacterial terminal oxidase cytochrome bd with nitric oxide. FEBS Lett 576:201–204. [PubMed] 10.1016/j.febslet.2004.09.013 [DOI] [PubMed] [Google Scholar]
- 74.Borisov VB, Forte E, Sarti P, Brunori M, Konstantinov AA, Giuffrè A. 2006. Nitric oxide reacts with the ferryl-oxo catalytic intermediate of the CuB-lacking cytochrome bd terminal oxidase. FEBS Lett 580:4823–4826. [PubMed] 10.1016/j.febslet.2006.07.072 [DOI] [PubMed] [Google Scholar]
- 75.Borisov VB, Forte E, Sarti P, Brunori M, Konstantinov AA, Giuffrè A. 2007. Redox control of fast ligand dissociation from Escherichia coli cytochrome bd. Biochem Biophys Res Commun 355:97–102. [PubMed] 10.1016/j.bbrc.2007.01.118 [DOI] [PubMed] [Google Scholar]
- 76.Mason MG, Shepherd M, Nicholls P, Dobbin PS, Dodsworth KS, Poole RK, Cooper CE. 2009. Cytochrome bd confers nitric oxide resistance to Escherichia coli. Nat Chem Biol 5:94–96. [PubMed] 10.1038/nchembio.135 [DOI] [PubMed] [Google Scholar]
- 77.Borisov VB, Forte E, Giuffrè A, Konstantinov A, Sarti P. 2009. Reaction of nitric oxide with the oxidized di-heme and heme-copper oxygen-reducing centers of terminal oxidases: Different reaction pathways and end-products. J Inorg Biochem 103:1185–1187. [PubMed] 10.1016/j.jinorgbio.2009.06.002 [DOI] [PubMed] [Google Scholar]
- 78.Jesse HE, Nye TL, McLean S, Green J, Mann BE, Poole RK. 2013. Cytochrome bd-I in Escherichia coli is less sensitive than cytochromes bd-II or bo” to inhibition by the carbon monoxide-releasing molecule, CORM-3: N-acetylcysteine reduces CO-RM uptake and inhibition of respiration. Biochim Biophys Acta 1834:1693–1703. [PubMed] 10.1016/j.bbapap.2013.04.019 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Tinajero-Trejo M, Jesse HE, Poole RK. 2013. Gasotransmitters, poisons, and antimicrobials: it’s a gas, gas, gas! F1000Prime Rep 5:28. [PubMed] 10.12703/P5-28 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Sun YH, de Jong MF, den Hartigh AB, Roux CM, Rolan HG, Tsolis RM. 2012. The small protein CydX is required for function of cytochrome bd oxidase in Brucella abortus. Front Cell Infect Microbiol 2:47. [PubMed] 10.3389/fcimb.2012.00047 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Lindqvist A, Membrillo-Hernandez J, Poole RK, Cook GM. 2000. Roles of respiratory oxidases in protecting Escherichia coli K12 from oxidative stress. Antonie Van Leeuwenhoek 78:23–31. [PubMed] 10.1023/A:1002779201379 [DOI] [PubMed] [Google Scholar]
- 82.Edwards SE, Loder CS, Wu G, Corker H, Bainbridge BW, Hill S, Poole RK. 2000. Mutation of cytochrome bd quinol oxidase results in reduced stationary phase survival, iron deprivation, metal toxicity and oxidative stress in Azotobacter vinelandii. FEMS Microbiol Lett 185:71–77. [PubMed] 10.1111/j.1574-6968.2000.tb09042.x [DOI] [PubMed] [Google Scholar]
- 83.Korshunov S, Imlay JA. 2010. Two sources of endogenous hydrogen peroxide in Escherichia coli. Mol Microbiol 75:1389–1401. [PubMed] 10.1111/j.1365-2958.2010.07059.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Borisov VB, Davletshin AI, Konstantinov AA. 2010. Peroxidase activity of cytochrome bd from Escherichia coli. Biochemistry (Mosc) 75:428–436. [PubMed] 10.1134/S000629791004005X [DOI] [PubMed] [Google Scholar]
- 85.Borisov VB, Forte E, Davletshin A, Mastronicola D, Sarti P, Giuffrè A. 2013. Cytochrome bd oxidase from Escherichia coli displays high catalase activity: an additional defense against oxidative stress. FEBS Lett 587:2214–2218. [PubMed] 10.1016/j.febslet.2013.05.047 [DOI] [PubMed] [Google Scholar]
- 86.Forte E, Borisov VB, Davletshin A, Mastronicola D, Sarti P, Giuffrè A. 2013. Cytochrome bd oxidase and hydrogen peroxide resistance in Mycobacterium tuberculosis. MBio 4:e01006–01013. [PubMed] 10.1128/mBio.01006-13 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Meunier B, Madgwick SA, Reil E, Oettmeier W, Rich PR. 1995. New inhibitors of the quinol oxidation sites of bacterial cytochromes bo and bd. Biochemistry 34:1076–1083. [PubMed] 10.1021/bi00003a044 [DOI] [PubMed] [Google Scholar]
- 88.Mogi T, Ui H, Shiomi K, Omura S, Kita K. 2008. Gramicidin S identified as a potent inhibitor for cytochrome bd-type quinol oxidase. FEBS Lett 582:2299–2302. [PubMed] 10.1016/j.febslet.2008.05.031 [DOI] [PubMed] [Google Scholar]
- 89.Borisov VB. 1996. Cytochrome bd: structure and properties. Biochemistry (Mosc) 61:565–574. [Google Scholar]
- 90.Matsushita K, Patel L, Gennis RB, Kaback HR. 1983. Reconstitution of active transport in proteoliposomes containing cytochromes o oxidase and lac carrier protein purified from E. coli. Proc Natl Acad Sci USA 80:4889–4893. [PubMed] 10.1073/pnas.80.16.4889 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Matsushita K, Patel L, Kaback HR. 1984. Cytochrome o type oxidase from Escherichia coli. Characterization of the enzyme and mechanism of electrochemical proton gradient generation. Biochemistry 23:4703–4714. [PubMed] 10.1021/bi00315a028 [DOI] [PubMed] [Google Scholar]
- 92.Georgiou C, Cokic P, Carter K, Webster DA, Gennis RB. 1988. Relationships between membrane-bound cytochrome o from Vitreoscilla and that of Escherichia coli. Biochim Biophys Acta 933:179–183. [PubMed] 10.1016/0005-2728(88)90068-0 [DOI] [PubMed] [Google Scholar]
- 93.Bachmann BJ. 1990. Linkage map of Escherichia coli K-12, edition 8. Microbiol Rev 54:130–197. [PubMed] [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Nakamura H, Yamoto I, Anraku Y, Lemieux L, Gennis RB. 1990. Expression of cyoA and cyoB demonstrates that the CO-binding heme component of the E. coli cytochrome o complex is in subunit I. J Biol Chem 265:11193–11197. [PubMed] [PubMed] [Google Scholar]
- 95.Nakamura H, Saiki K, Mogi T, Anraku Y. 1997. Assignment and functional roles of the cyoABCDE gene products required for the Escherichia coli bo-type quinol oxidase. J Biochem (Tokyo) 122:415–421. [PubMed] 10.1093/oxfordjournals.jbchem.a021769 [DOI] [PubMed] [Google Scholar]
- 96.Saraste M. 1990. Structural features of cytochrome oxidase. Q Rev Biophys 23:331–366. [PubMed] 10.1017/S0033583500005588 [DOI] [PubMed] [Google Scholar]
- 97.Chepuri V, Gennis RB. 1990. The use of gene fusions to determine the topology of all of the subunits of the cytochrome o terminal oxidase complex of Escherichia coli. J Biol Chem 265:12978–12986. [PubMed] [PubMed] [Google Scholar]
- 98.Saiki K, Mogi T, Tsubaki M, Hori H, Anraku Y. 1997. Exploring subunit-subunit interactions in the Escherichia coli bo-type ubiquinol oxidase by extragenic suppressor mutation analysis. J Biol Chem 272:14721–14726. [PubMed] 10.1074/jbc.272.23.14721 [DOI] [PubMed] [Google Scholar]
- 99.Castresana J, Lübben M, Saraste M, Higgins DG. 1994. Evolution of cytochrome oxidase, an enzyme older than atmospheric oxygen. EMBO J 13:2516–2525. [PubMed] [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Saiki K, Mogi T, Anraku Y. 1992. Heme O biosynthesis in Escherichia coli: the CYOE gene in the cytochrome bo operon encodes a protoheme IX farnesyltransferase. Biochem Biophys Res Commun 189:1491–1497. 10.1016/0006-291X(92)90243-E [DOI] [PubMed] [Google Scholar]
- 101.Saiki K, Mogi T, Ogura K, Anraku Y. 1993. In vitro heme O synthesis by the cyoE gene product from Escherichia coli. J Biol Chem 268:26041–26045. [PubMed] [PubMed] [Google Scholar]
- 102.Mogi T, Saiki K, Anraku Y. 1994. Biosynthesis and functional role of haem O and haem A. Mol Microbiol 14:391–398. [PubMed] 10.1111/j.1365-2958.1994.tb02174.x [DOI] [PubMed] [Google Scholar]
- 103.Minagawa J, Mogi T, Gennis RB, Anraku Y. 1992. Identification of heme ligands in subunit I of the cytochrome bo complex in Escherichia coli. J Biol Chem 267:2096–2104. [PubMed] [PubMed] [Google Scholar]
- 104.Lemieux LJ, Calhoun MW, Thomas JW, Ingledew WJ, Gennis RB. 1992. Determination of the ligands of the low-spin heme of the cytochrome O ubiquinol oxidase complex using site-directed mutagenesis. J Biol Chem 267:2105–2113. [PubMed] [PubMed] [Google Scholar]
- 105.Saiki K, Nakamura H, Mogi T, Anraku Y. 1996. Probing a role of subunit IV of the Escherichia coli bo-type ubiquinol oxidase by deletion and cross-linking analyses. J Biol Chem 271:15336–15340. [PubMed] 10.1074/jbc.271.26.15336 [DOI] [PubMed] [Google Scholar]
- 106.Miller MJ, Hermodson M, Gennis RB. 1988. The active form of the cytochrome d terminal oxidase complex of Escherichia coli is a heterodimer containing one copy of each of the two subunits. J Biol Chem 263:5235–5240. [PubMed] [PubMed] [Google Scholar]
- 107.Green GN, Kranz JE, Gennis RB. 1984. Cloning the cyd gene locus coding for the cytochrome d complex of Escherichia coli. Gene 32:99–106. [PubMed] 10.1016/0378-1119(84)90037-4 [DOI] [PubMed] [Google Scholar]
- 108.Poole RK. 1994. Oxygen reactions with bacterial oxidases and globins: binding, reduction and regulation. Anthonie van Leeuwenhoek 65:289–310. [PubMed] 10.1007/BF00872215 [DOI] [PubMed] [Google Scholar]
- 109.Allen RJ, Brenner EP, VanOrsdel CE, Hobson JJ, Hearn DJ, Hemm MR. 2014. Conservation analysis of the CydX protein yields insights into small protein identification and evolution. BMC Genomics 15:946. [PubMed] 10.1186/1471-2164-15-946 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Chen H, Luo Q, Yin J, Gao T, Gao H. 2015. Evidence for requirement of CydX in function but not assembly of the cytochrome bd oxidase in Shewanella oneidensis. Biochim Biophys Acta 1850:318–328. [PubMed] 10.1016/j.bbagen.2014.10.005 [DOI] [PubMed] [Google Scholar]
- 111.Lorence RM, Koland JG, Gennis RB. 1986. Coulometric and spectroscopic analysis of the purified cytochrome d complex of Escherichia coli: evidence for the identification of “cytochrome a1” as cytochrome b595. Biochemistry 25:2314–2321. [PubMed] 10.1021/bi00357a003 [DOI] [PubMed] [Google Scholar]
- 112.Newton G, Gennis RB. 1991. In vivo assembly of the cytochrome d terminal oxidase complex of Escherichia coli from genes encoding the two subunits expressed on separate plasmids. Biochim Biophys Acta 1089:8–12. [PubMed] 10.1016/0167-4781(91)90077-Y [DOI] [PubMed] [Google Scholar]
- 113.Green GN, Lorence RM, Gennis RB. 1986. Specific overproduction and purification of the cytochrome b558 component of the cytochrome d complex from Escherichia coli. Biochemistry 25:2309–2314. [PubMed] 10.1021/bi00357a002 [DOI] [PubMed] [Google Scholar]
- 114.Georgiou CD, Fang H, Gennis RB. 1987. Identification of the cydC locus required for the expression of the functional form of the cytochrome d terminal oxidase complex in Escherichia coli. J Bacteriol 169:2107–2112. [PubMed] [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Poole RK, Williams HD, Downie JA, Gibson F. 1989. Mutations affecting the cytochrome d-containing oxidase complex of Escherichia coli K12: Identification and mapping of a fourth locus, cydD . J Gen Microbiol 135:1865–1874. [PubMed] 10.1099/00221287-135-7-1865 [DOI] [PubMed] [Google Scholar]
- 116.Poole RK, Hatch L, Cleeter MWJ, Gibson F, Cox GB, Wu G. 1993. Cytochrome bd biosynthesis in Escherichia coli: the sequences of the cydC and cydD genes suggest that they encode the components of an ABC membrane transporter. Mol Microbiol 10:421–430. [PubMed] 10.1111/j.1365-2958.1993.tb02673.x [DOI] [PubMed] [Google Scholar]
- 117.Bebbington KJ, Williams HD. 1993. Investigation of the role of the cydD gene product in production of a functional cytochrome d oxidase in Escherichia coli. FEMS Microbiol Lett 112:19–24. [PubMed] 10.1111/j.1574-6968.1993.tb06417.x [DOI] [PubMed] [Google Scholar]
- 118.Poole RK, Gibson F, Wu G. 1994. The cydD gene product, component of a heterodimeric ABC transporter, is required for assembly of periplasmic cytochrome c and of cytochrome bd in Escherichia coli. FEMS Microbiol Lett 117:217–224. [PubMed] 10.1111/j.1574-6968.1994.tb06768.x [DOI] [PubMed] [Google Scholar]
- 119.Pittman MS, Robinson HC, Poole RK. 2005. A bacterial glutathione transporter (Escherichia coli CydDC) exports reductant to the periplasm. J Biol Chem 280:32254–32261. [PubMed] 10.1074/jbc.M503075200 [DOI] [PubMed] [Google Scholar]
- 120.Iuchi S, Chepuri V, Fu HA, Gennis RB, Lin EC. 1990. Requirement for terminal cytochromes in generation of the aerobic signal for the arc regulatory system in Escherichia coli: study utilizing deletions and lac fusions of cyo and cyd. J Bacteriol 172:6020–6025. [PubMed] [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121.Cotter PA, Gunsalus RP. 1992. Contribution of the fnr and arcA gene products in coordinate regulation of cytochrome o and d oxidase (cyoABCDE and cydAB) genes in Escherichia coli. FEMS Microbiol Lett 91:31–36. 10.1111/j.1574-6968.1992.tb05179.x [DOI] [PubMed] [Google Scholar]
- 122.Gunsalus RP. 1992. Control of electron flow in Escherichia coli: coordinated transcription of respiratory pathway genes. J Bacteriol 174:7069–7074. [PubMed] [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Tseng C-P, Albrecht J, Gunsalus RP. 1996. Effect of microaerophilic cell growth conditions on expression of the aerobic (cyoABCDE and cydAB) and anaerobic (narGHJI, frdABCD, and dmsABC) respiratory pathway genes in Escherichia coli. J Bacteriol 178:1094–1098. [PubMed] [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Cotter PA, Melville SB, Albrecht JA, Gunsalus RP. 1997. Aerobic regulation of cytochrome d oxidase (cydAB) operon expression in Escherichia coli: roles of Fnr and ArcA in repression and activation. Mol Microbiol 25:605–615. [PubMed] 10.1046/j.1365-2958.1997.5031860.x [DOI] [PubMed] [Google Scholar]
- 125.Govantes F, Albrecht JA, Gunsalus RP. 2000. Oxygen regulation of the Escherichia coli cytochrome d oxidase (cydAB) operon: roles of multiple promoters and the Fnr-1 and Fnr-2 binding sites. Mol Microbiol 37:1456–1469. [PubMed] 10.1046/j.1365-2958.2000.02100.x [DOI] [PubMed] [Google Scholar]
- 126.Shalel-Levanon S, San KY, Bennett GN. 2005. Effect of oxygen, and ArcA and FNR regulators on the expression of genes related to the electron transfer chain and the TCA cycle in Escherichia coli. Metab Eng 7:364–374. [PubMed] 10.1016/j.ymben.2005.07.001 [DOI] [PubMed] [Google Scholar]
- 127.Lynch AS, Lin ECC. 1996. Responses to molecular oxygen, p 1526–1538. In Neidhardt FCea (ed), Escherichia coli and Salmonella typhimurium: Cellular and Molecular Biology, 2nd ed. ASM Press, Washington, DC. [Google Scholar]
- 128.Georgellis D, Kwon O, Lin EC. 2001. Quinones as the redox signal for the arc two-component system of bacteria. Science 292:2314–2316. [PubMed] 10.1126/science.1059361 [DOI] [PubMed] [Google Scholar]
- 129.Georgellis D, Kwon O, Lin EC. 1999. Amplification of signaling activity of the arc two-component system of Escherichia coli by anaerobic metabolites. An in vitro study with different protein modules. J Biol Chem 274:35950–35954. [PubMed] 10.1074/jbc.274.50.35950 [DOI] [PubMed] [Google Scholar]
- 130.Alexeeva S, Hellingwerf KJ, Teixeira de Mattos MJ. 2003. Requirement of ArcA for redox regulation in Escherichia coli under microaerobic but not anaerobic or aerobic conditions. J Bacteriol 185:204–209. [PubMed] 10.1128/JB.185.1.204-209.2003 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131.Kiley PJ, Beinert H. 1998. Oxygen sensing by the global regulator, FNR: the role of the iron-sulfur cluster. FEMS Microbiol Rev 22:341–352. [PubMed] 10.1111/j.1574-6976.1998.tb00375.x [DOI] [PubMed] [Google Scholar]
- 132.Overton TW, Griffiths L, Patel MD, Hobman JL, Penn CW, Cole JA, Constantinidou C. 2006. Microarray analysis of gene regulation by oxygen, nitrate, nitrite, FNR, NarL and NarP during anaerobic growth of Escherichia coli: new insights into microbial physiology. Biochem Soc Trans 34:104–107. [PubMed] 10.1042/BST0340104 [DOI] [PubMed] [Google Scholar]
- 133.Becker S, Holighaus G, Gabrielczyk T, Unden G. 1996. O2 as the regulatory signal for FNR-dependent gene regulation in Escherichia coli. J Bacteriol 178:4515–4521. [PubMed] [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134.Collins MD, Jones D. 1981. Distribution of isoprenoid quinone structural types in bacteria and their taxonomic implication. Microbiol Rev 45:316–354. [PubMed] [DOI] [PMC free article] [PubMed] [Google Scholar]
- 135.Shestopalov AI, Bogachev AV, Murtazina RA, Viryasov MB, Skulachev VP. 1997. Aeration-dependent changes in composition of the quinone pool in Escherichia coli. Evidence of post-transcriptional regulation of the quinone biosynthesis. FEBS Lett 404:272–274. [PubMed] 10.1016/s0014-5793(97)00143-9 [DOI] [PubMed] [Google Scholar]
- 136.Sato-Watanabe M, Mogi T, Ogura T, Kitagawa T, Miyoshi H, Iwamura H, Anraku Y. 1994. Identification of a Novel Quinone Binding Site in the Cytochrome bo Complex from Escherichia coli. J Biol Chem 269:28908–28912. [PubMed] [PubMed] [Google Scholar]
- 137.Mogi T, Endou S, Akimoto S, Morimoto-Tadokoro M, Miyoshi H. 2006. Glutamates 99 and 107 in transmembrane helix III of subunit I of cytochrome bd are critical for binding of the heme b595-d binuclear center and enzyme activity. Biochemistry 45:15785–15792. [PubMed] 10.1021/bi0615792 [DOI] [PubMed] [Google Scholar]
- 138.Belevich I, Bloch DA, Belevich N, Wikström M, Verkhovsky MI. 2007. Exploring the proton pump mechanism of cytochrome c oxidase in real time. Proc Natl Acad Sci USA 104:2685–2690. [PubMed] 10.1073/pnas.0608794104 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 139.Abramson J, Riistama S, Larsson G, Jasaitis A, Svensson-Ek M, Laakkonen L, Puustinen A, Iwata S, Wikström M. 2000. The structure of the ubiquinol oxidase from Escherichia coli and its ubiquinone binding site. Nat Struct Biol 7:910–917. [PubMed] 10.1038/82824 [DOI] [PubMed] [Google Scholar]
- 140.Hellwig P, Mogi T, Tomson FL, Gennis RB, Iwata J, Miyoshi H, Mantele W. 1999. Vibrational modes of ubiquinone in cytochrome bo3 from Escherichia coli identified by Fourier transform infrared difference spectroscopy and specific 13C labeling. Biochemistry 38:14683–14689. [PubMed] 10.1021/bi991267h [DOI] [PubMed] [Google Scholar]
- 141.Ma J, Tsatsos PH, Zaslavsky D, Barquera B, Thomas JW, Katsonouri A, Puustinen A, Wikström M, Brzezinski P, Alben JO, Gennis RB. 1999. Glutamate-89 in subunit II of cytochrome bo3 from Escherichia coli is required for the function of the heme-copper oxidase. Biochemistry 38:15150–15156. [PubMed] 10.1021/bi991764y [DOI] [PubMed] [Google Scholar]
- 142.Hellwig P, Yano T, Ohnishi T, Gennis RB. 2002. Identification of the residues involved in stabilization of the semiquinone radical in the high-affinity ubiquinone binding site in cytochrome bo3 from Escherichia coli by site-directed mutagenesis and EPR spectroscopy. Biochemistry 41:10675–10679. [PubMed] 10.1021/bi012146w [DOI] [PubMed] [Google Scholar]
- 143.Babcock GT, Callahan PM, Ondrias MR, Salmeen I. 1981. Coordination geometries and vibrational properties of cytochromes a and a3 in cytochrome oxidase from soret excitation Raman spectroscopy. Biochemistry 20:959–966. [PubMed] 10.1021/bi00507a049 [DOI] [PubMed] [Google Scholar]
- 144.Salerno JC, Bolgiano B, Ingledew WJ. 1989. Potentiometric titration of cytochrome-bo type quinol oxidase of Escherichia coli: evidence for heme-heme and copper-heme interaction. FEBS Lett 247:101–105. [PubMed] 10.1016/0014-5793(89)81249-9 [DOI] [PubMed] [Google Scholar]
- 145.Puustinen A, Morgan JE, Verkhovsky M, Thomas JW, Gennis RB, Wikström M. 1992. The low spin heme site of cytochrome o from E. coli is promiscuous with respect to heme type. Biochemistry 31:10363–10369. [PubMed] 10.1021/bi00157a026 [DOI] [PubMed] [Google Scholar]
- 146.Babcock GT, Vickery LE, Palmer G. 1976. Electronic state of heme in cytochrome oxidase. I. Magnetic circular dichroism of the isolated enzyme and its derivatives. J Biol Chem 251:7907–7919. [PubMed] [PubMed] [Google Scholar]
- 147.Cheesman MR, Watmough NJ, Pires CA, Turner R, Brittain T, Gennis RB, Greenwood C, Thomson AJ. 1993. Cytochrome bo from Escherichia coli: identification of haem ligands and reaction of the reduced enzyme with carbon monoxide. Biochem J 289:709–718. [PubMed] 10.1042/bj2890709 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148.Ralle M, Verkhovskaya ML, Morgan JE, Verkhovsky MI, Wikström M, Blackburn NJ. 1999. Coordination of CuB in reduced and CO-liganded states of cytochrome bo3 from Escherichia coli. Is chloride ion a cofactor? Biochemistry 38:7185–7194. [PubMed] 10.1021/bi982885l [DOI] [PubMed] [Google Scholar]
- 149.Osborne JP, Cosper NJ, Stalhandske CM, Scott RA, Alben JO, Gennis RB. 1999. Cu XAS shows a change in the ligation of CuB upon reduction of cytochrome bo3 from Escherichia coli. Biochemistry 38:4526–4532. [PubMed] 10.1021/bi982278y [DOI] [PubMed] [Google Scholar]
- 150.Sato-Watanabe M, Mogi T, Miyoshi H, Iwamura H, Matsushita K, Adachi O, Anraku Y. 1994. Structure-function studies on the ubiquinol oxidation site of the cytochrome bo complex from Escherichia coli using p-benzoquinones and substituted phenols. J Biol Chem 269:28899–28907. [PubMed] [PubMed] [Google Scholar]
- 151.Lin MT, Baldansuren A, Hart R, Samoilova RI, Narasimhulu KV, Yap LL, Choi SK, O’Malley PJ, Gennis RB, Dikanov SA. 2012. Interactions of intermediate semiquinone with surrounding protein residues at the QH site of wild-type and D75H mutant cytochrome bo3 from Escherichia coli. Biochemistry 51:3827–3838. [PubMed] 10.1021/bi300151q [DOI] [PMC free article] [PubMed] [Google Scholar]
- 152.Bossis F, De Grassi A, Palese LL, Pierri CL. 2014. Prediction of high- and low-affinity quinol-analogue-binding sites in the aa3 and bo3 terminal oxidases from Bacillus subtilis and Escherichia coli. Biochem J 461:305–314. [PubMed] 10.1042/BJ20140082 [DOI] [PubMed] [Google Scholar]
- 153.Yap LL, Lin MT, Ouyang H, Samoilova RI, Dikanov SA, Gennis RB. 2010. The quinone-binding sites of the cytochrome bo3 ubiquinol oxidase from Escherichia coli. Biochim Biophys Acta 1797:1924–1932. [PubMed] 10.1016/j.bbabio.2010.04.011 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154.Ingledew WJ, Ohnishi T, Salerno JC. 1995. Studies on a stabilisation of ubisemiquinone by Escherichia coli quinol oxidase, cytochrome bo. Eur J Biochem 227:903–908. [PubMed] 10.1111/j.1432-1033.1995.tb20217.x [DOI] [PubMed] [Google Scholar]
- 155.Sato-Watanabe M, Itoh S, Mogi T, Matsuura K, Miyoshi H, Anraku Y. 1995. Stabilization of a semiquinone radical at the high-affinity quinone-binding site (QH) of the Escherichia coli bo-type ubiquinol oxidase. FEBS Lett 374:265–269. [PubMed] 10.1016/0014-5793(95)01125-X [DOI] [PubMed] [Google Scholar]
- 156.Lin MT, Shubin AA, Samoilova RI, Narasimhulu KV, Baldansuren A, Gennis RB, Dikanov SA. 2011. Exploring by pulsed EPR the electronic structure of ubisemiquinone bound at the QH site of cytochrome bo3 from Escherichia coli with in vivo 13C-labeled methyl and methoxy substituents. J Biol Chem 286:10105–10114. [PubMed] 10.1074/jbc.M110.206821 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 157.Rothery RA, Houston AM, Ingledew WJ. 1987. The respiratory chain of anaerobically grown Escherichia coli: reactions with nitrite and oxygen. J Gen Microbiol 133:3247–3255. [PubMed] 10.1099/00221287-133-11-3247 [DOI] [PubMed] [Google Scholar]
- 158.Meinhardt SW, Gennis RB, Ohnishi T. 1989. EPR studies of the cytochrome-d complex of Escherichia coli. Biochim Biophys Acta 975:175–184. [PubMed] 10.1016/S0005-2728(89)80216-6 [DOI] [PubMed] [Google Scholar]
- 159.Rothery R, Ingledew WJ. 1989. The cytochromes of anaerobically grown Escherichia coli. Biochem J 262:437–443. 10.1042/bj2610437 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 160.Kahlow MA, Zuberi TM, Gennis RB, Loehr TM. 1991. Identification of a ferryl intermediate of Escherichia coli cytochrome d terminal oxidase by Resonance Raman spectroscopy. Biochemistry 30:11485–11489. [PubMed] 10.1021/bi00113a001 [DOI] [PubMed] [Google Scholar]
- 161.Hill BC, Hill JJ, Gennis RB. 1994. The room temperature reaction of carbon monoxide and oxygen with the cytochrome bd quinol oxidase from Escherichia coli. Biochemistry 33:15110–15115. [PubMed] 10.1021/bi00254a021 [DOI] [PubMed] [Google Scholar]
- 162.Ingledew WJ, Rothery RA, Gennis RB, Salerno JC. 1992. The orientation of the three haems of the in situ ubiquinol oxidase, cytochrome bd, of Escherichia coli. Biochem J 282:255–259. [PubMed] 10.1042/bj2820255 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 163.Dueweke TJ, Gennis RB. 1990. Epitopes of monoclonal antibodies which inhibit ubiquinol oxidase activity of Escherichia coli cytochrome d complex localize a functional domain. J Biol Chem 265:4273–4277. [PubMed] [PubMed] [Google Scholar]
- 164.Kranz RG, Gennis RB. 1984. Characterization of the cytochrome d terminal oxidase complex of Escherichia coli using polyclonal and monoclonal antibodies. J Biol Chem 259:7998–8003. [PubMed] [PubMed] [Google Scholar]
- 165.Lorence RM, Carter K, Gennis RB, Matsushita K, Kaback HR. 1988. Trypsin proteolysis of the cytochrome d complex of Escherichia coli selectively inhibits ubiquinol oxidase activity while not affecting N,N,N′,N′-tetramethyl-p-phenylenediamine oxidase activity. J Biol Chem 11:5271–5276. [PubMed] [Google Scholar]
- 166.Dueweke TJ, Gennis RB. 1991. Proteolysis of the cytochrome d complex with trypsin and chymotrypsin localizes a quinol oxidase domain. Biochemistry 30:3401–3406. [PubMed] 10.1021/bi00228a007 [DOI] [PubMed] [Google Scholar]
- 167.Poole RK. 1988. Bacterial cytochrome oxidases, p 231–291. In Anthony C (ed), Bacterial Energy Transduction. Academic Press, London. [Google Scholar]
- 168.Bloch DA, Borisov VB, Mogi T, Verkhovsky MI. 2009. Heme/heme redox interaction and resolution of individual optical absorption spectra of the hemes in cytochrome bd from Escherichia coli. Biochim Biophys Acta 1787:1246–1253. [PubMed] 10.1016/j.bbabio.2009.05.003 [DOI] [PubMed] [Google Scholar]
- 169.Koland JG, Miller MJ, Gennis RB. 1984. Potentiometric analysis of the purified cytochrome d terminal oxidase complex from Escherichia coli. Biochemistry 23:1051–1056. 10.1021/bi00301a003 [DOI] [Google Scholar]
- 170.Borisov V, Arutyunyan AM, Osborne JP, Gennis RB, Konstantinov AA. 1999. Magnetic circular dichroism used to examine the interaction of Escherichia coli cytochrome bd with ligands. Biochemistry 38:740–750. [PubMed] 10.1021/bi981908t [DOI] [PubMed] [Google Scholar]
- 171.Arutyunyan AM, Borisov VB, Novoderezhkin VI, Ghaim J, Zhang J, Gennis RB, Konstantinov AA. 2008. Strong excitonic interactions in the oxygen-reducing site of bd-type oxidase: the Fe-to-Fe distance between hemes d and b595 is 10 Å. Biochemistry 47:1752–1759. [PubMed] 10.1021/bi701884g [DOI] [PubMed] [Google Scholar]
- 172.Arutyunyan AM, Sakamoto J, Inadome M, Kabashima Y, Borisov VB. 2012. Optical and magneto-optical activity of cytochrome bd from Geobacillus thermodenitrificans. Biochim Biophys Acta 1817:2087–2094. [PubMed] 10.1016/j.bbabio.2012.06.009 [DOI] [PubMed] [Google Scholar]
- 173.Fang H, Lin R-J, Gennis RB. 1989. Location of heme axial ligands in the cytochrome d terminal oxidase complex of Escherichia coli determined by site-directed mutagenesis. J Biol Chem 264:8026–8032. [PubMed] [PubMed] [Google Scholar]
- 174.Spinner F, Cheesman MR, Thomson AJ, Kaysser T, Gennis RB, Peng Q, Peterson J. 1995. The haem b558 component of the cytochrome bd quinol oxidase complex from Escherichia coli has histidine-methionine axial ligation. Biochem J 308:641–644. [PubMed] 10.1042/bj3080641 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 175.Kaysser TM, Ghaim JB, Georgiou C, Gennis RB. 1995. Methionine-393 is an axial ligand of the heme b558 component of the cytochrome bd ubiquinol oxidase from Escherichia coli. Biochemistry 34:13491–13501. [PubMed] 10.1021/bi00041a029 [DOI] [PubMed] [Google Scholar]
- 176.Osborne JP, Gennis RB. 1999. Sequence analysis of cytochrome bd oxidase suggests a revised topology for subunits I. Biochim Biophys Acta 1410:32–50. [PubMed] 10.1016/S0005-2728(98)00171-6 [DOI] [PubMed] [Google Scholar]
- 177.Zhang J, Barquera B, Gennis RB. 2004. Gene fusions with β-lactamase show that subunit I of the cytochrome bd quinol oxidase from E. coli has nine transmembrane helices with the O2 reactive site near the periplasmic surface. FEBS Lett 561:58–62. [PubMed] 10.1016/S0014-5793(04)00125-5 [DOI] [PubMed] [Google Scholar]
- 178.Poole RK. 1983. Bacterial cytochrome oxidases: a structurally and functionally diverse group of electron transfer proteins. Biochim Biophys Acta 726:205–243. [PubMed] 10.1016/0304-4173(83)90006-X [DOI] [PubMed] [Google Scholar]
- 179.Vos MH, Borisov VB, Liebl U, Martin J-L, Konstantinov AA. 2000. Femtosecond resolution of ligand-heme interactions in the high-affinity quinol oxidase bd: A di-heme active site? Proc Natl Acad Sci USA 97:1554–1559. [PubMed] 10.1073/pnas.030528197 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 180.Sun J, Kahlow MA, Kaysser TM, Osborne JP, Hill JJ, Rohlfs RJ, Hille R, Gennis RB, Loehr TM. 1996. Resonance Raman spectroscopic identification of a histidine ligand of b595 and the nature of the ligation of chlorin d in the fully reduced Escherichia coli cytochrome bd oxidase. Biochemistry 35:2403–2412. [PubMed] 10.1021/bi9518252 [DOI] [PubMed] [Google Scholar]
- 181.Hill JJ, Alben JO, Gennis RB. 1993. Spectroscopic evidence for a heme-heme binuclear center in the cytochrome bd ubiquinol oxidase from Escherichia coli. Proc Natl Acad Sci USA 90:5863–5867. [PubMed] 10.1073/pnas.90.12.5863 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 182.Tsubaki M, Hori H, Mogi T, Anraku Y. 1995. Cyanide-binding site of bd-type ubiquinol oxidase from Escherichia coli. J Biol Chem 270:28565–28569. [PubMed] 10.1074/jbc.270.48.28565 [DOI] [PubMed] [Google Scholar]
- 183.Borisov VB, Gennis RB, Konstantinov AA. 1995. Interaction of cytochrome bd from Escherichia coli with hydrogen peroxide. Biochemistry (Mosc) 60:231–239. [Google Scholar]
- 184.Borisov VB, Sedelnikova SE, Poole RK, Konstantinov AA. 2001. Interaction of cytochrome bd with carbon monoxide at low and room temperatures: evidence that only a small fraction of heme b595 reacts with CO. J Biol Chem 276:22095–22099. [PubMed] 10.1074/jbc.M011542200 [DOI] [PubMed] [Google Scholar]
- 185.Borisov VB, Liebl U, Rappaport F, Martin J-L, Zhang J, Gennis RB, Konstantinov AA, Vos MH. 2002. Interactions between heme d and heme b595 in quinol oxidase bd from Escherichia coli: a photoselection study using femtosecond spectroscopy. Biochemistry 41:1654–1662. [PubMed] 10.1021/bi0158019 [DOI] [PubMed] [Google Scholar]
- 186.Rappaport F, Zhang J, Vos MH, Gennis RB, Borisov VB. 2010. Heme-heme and heme-ligand interactions in the di-heme oxygen-reducing site of cytochrome bd from Escherichia coli revealed by nanosecond absorption spectroscopy. Biochim Biophys Acta 1797:1657–1664. [PubMed] 10.1016/j.bbabio.2010.05.010 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 187.Borisov VB, Verkhovsky MI. 2013. Accommodation of CO in the di-heme active site of cytochrome bd terminal oxidase from Escherichia coli. J Inorg Biochem 118:65–67. [PubMed] 10.1016/j.jinorgbio.2012.09.016 [DOI] [PubMed] [Google Scholar]
- 188.Siletsky SA, Zaspa AA, Poole RK, Borisov VB. 2014. Microsecond time-resolved absorption spectroscopy used to study CO compounds of cytochrome bd from Escherichia coli. PLoS One 9:e95617. 10.1371/journal.pone.0095617. [PubMed] 10.1371/journal.pone.0095617 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 189.Borisov VB. 2008. Interaction of bd-type quinol oxidase from Escherichia coli and carbon monoxide: heme d binds CO with high affinity. Biochemistry (Mosc) 73:14–22. [PubMed] 10.1134/S0006297908010021 [DOI] [PubMed] [Google Scholar]
- 190.Poole RK, Williams HD. 1987. Proposal that the function of the membrane-bound cytochrome a1-like haemoprotein (cytochrome b-595) in Escherichia coli is a direct electron donation to cytochrome d. FEBS Lett 217:49–52. [PubMed] 10.1016/0014-5793(87)81240-1 [DOI] [PubMed] [Google Scholar]
- 191.Hata-Tanaka A, Matsuura K, Itoh S, Anraku Y. 1987. Electron flow and heme-heme interaction between cytochromes b-558, b-595 and d in a terminal oxidase of Escherichia coli. Biochim Biophys Acta 893:289–295. [PubMed] 10.1016/0005-2728(87)90050-8 [DOI] [PubMed] [Google Scholar]
- 192.Timkovich R, Cork MS, Gennis RB, Johnson PY. 1985. Proposed structure of heme d, a prosthetic group of bacterial terminal oxidases. J Am Chem Soc 107:6069–6075. 10.1021/ja00307a041 [DOI] [Google Scholar]
- 193.Borisov VB, Smirnova IA, Krasnosel’skaya IA, Konstantinov AA. 1994. Oxygenated cytochrome bd from Escherichia coli can be converted into the oxidized form by lipophilic electron acceptors. Biochemistry (Mosc) 59:437–443. [PubMed] [Google Scholar]
- 194.Jiang FS, Zuberi TM, Cornelius JB, Clarkson RB, Gennis RB, Belford RL. 1993. Nitrogen and proton ENDOR of cytochrome d, hemin, and metmyoglobin in frozen solutions. J Am Chem Soc 115:10293–10299. 10.1021/ja00075a052 [DOI] [Google Scholar]
- 195.Hirota S, Mogi T, Anraku Y, Gennis RB, Kitagawa T. 1995. Resonance Raman study on axial ligands of heme irons in cytochrome bd-type ubiquinol oxidase from Escherichia coli. Biospectroscopy 1:305–311. 10.1002/bspy.350010502 [DOI] [Google Scholar]
- 196.Hori H, Tsubaki M, Mogi T, Anraku Y. 1996. EPR study of NO complex of bd-type ubiquinol oxidase from Escherichia coli. J Biol Chem 271:9254–9258. [PubMed] 10.1074/jbc.271.16.9254 [DOI] [PubMed] [Google Scholar]
- 197.Pudek MR, Bragg PD. 1976. Redox potentials of the cytochromes in the respiratory chain of aerobically grown Escherichia coli. Arch Biochem Biophys 174:546–552. [PubMed] 10.1016/0003-9861(76)90382-9 [DOI] [PubMed] [Google Scholar]
- 198.Lorence RM, Miller MJ, Borochov A, Faiman-Weinberg R, Gennis RB. 1984. Effects of pH and detergent on the kinetic and electrochemical properties of the purified cytochrome d terminal oxidase complex of Escherichia coli. Biochim Biophys Acta 790:148–153. [PubMed] 10.1016/0167-4838(84)90218-8 [DOI] [PubMed] [Google Scholar]
- 199.Hosler JP, Ferguson-Miller S, Calhoun MW, Thomas JW, Hill J, Lemieux L, Ma J, Georgiou C, Fetter J, Shapleigh J, Tecklenburg MMJ, Babcock GT, Gennis RB. 1993. Insight into the active-site structure and function of cytochrome oxidase by analysis of site-directed mutants of bacterial cytochrome aa3 and cytochrome bo. J Bioenerg Biomembr 25:121–136. [PubMed] 10.1007/BF00762854 [DOI] [PubMed] [Google Scholar]
- 200.Thomas JW, Lemieux LJ, Alben JO, Gennis RB. 1993. Site-directed mutagenesis of highly conserved residues in helix VIII of subunit I of the cytochrome bo ubiquinol oxidase from Escherichia coli: an amphipathic transmembrane helix that may be important in conveying protons to the binuclear center. Biochemistry 32:11173–11180. [PubMed] 10.1021/bi00092a029 [DOI] [PubMed] [Google Scholar]
- 201.Thomas JW, Puustinen A, Alben JO, Gennis RB, Wikström M. 1993. Substitution of asparagine for aspartate-135 in subunit I of the cytochrome bo ubiquinol oxidase of Escherichia coli eliminates proton-pumping activity. Biochemistry 32:10923–10928. [PubMed] 10.1021/bi00091a048 [DOI] [PubMed] [Google Scholar]
- 202.Fetter JR, Qian J, Shapleigh J, Thomas JW, Garcia-Horsman A, Schmidt E, Hosler J, Babcock GT, Gennis RB, Ferguson-Miller S. 1995. Possible proton relay pathways in cytochrome c oxidase. Proc Natl Acad Sci USA 92:1604–1608. [PubMed] 10.1073/pnas.92.5.1604 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 203.Matsumoto Y, Murai M, Fujita D, Sakamoto K, Miyoshi H, Yoshida M, Mogi T. 2006. Mass spectrometric analysis of the ubiquinol-binding site in cytochrome bd from Escherichia coli. J Biol Chem 281:1905–1912. [PubMed] 10.1074/jbc.M508206200 [DOI] [PubMed] [Google Scholar]
- 204.Mogi T, Akimoto S, Endou S, Watanabe-Nakayama T, Mizuochi-Asai E, Miyoshi H. 2006. Probing the ubiquinol-binding site in cytochrome bd by site-directed mutagenesis. Biochemistry 45:7924–7930. [PubMed] 10.1021/bi060192w [DOI] [PubMed] [Google Scholar]
- 205.Hao W, Golding GB. 2006. Asymmetrical evolution of cytochrome bd subunits. J Mol Evol 62:132–142. [PubMed] 10.1007/s00239-005-0005-7 [DOI] [PubMed] [Google Scholar]
- 206.Zhang J, Hellwig P, Osborne JP, Gennis RB. 2004. Arginine 391 in subunit I of the cytochrome bd quinol oxidase from Escherichia coli stabilizes the reduced form of the hemes and is essential for quinol oxidase activity. J Biol Chem 279:53980–53987. [PubMed] 10.1074/jbc.M408626200 [DOI] [PubMed] [Google Scholar]
- 207.Oden KL, Gennis RB. 1991. Isolation and characterization of a new class of cytochrome d terminal oxidase mutants of Escherichia coli. J Bacteriol 173:6174–6183. [PubMed] [DOI] [PMC free article] [PubMed] [Google Scholar]
- 208.Poole RK, Kumar C, Salmon I, Chance B. 1983. The 650 nm chromophore in Escherichia coli is an ‘Oxy-’ or oxygenated compound, not the oxidized form of cytochrome oxidase d: a hypothesis. J Gen Microbiol 129:1335–1344. 10.1099/00221287-129-5-1335 [DOI] [PubMed] [Google Scholar]
- 209.Lorence RM, Gennis RB. 1989. Spectroscopic and quantitative analysis of the oxygenated and peroxy states of the purified cytochrome d complex of Escherichia coli. J Biol Chem 264:7135–7140. [PubMed] [PubMed] [Google Scholar]
- 210.Kahlow MA, Loehr TM, Zuberi TM, Gennis RB. 1993. The oxygenated complex of cytochrome d terminal oxidase: direct evidence for Fe-O2 coordination in a chlorin-containing enzyme by resonance Raman spectroscopy. J Am Chem Soc 115:5845–5846. 10.1021/ja00066a071 [DOI] [Google Scholar]
- 211.Backhed F, Ley RE, Sonnenburg JL, Peterson DA, Gordon JI. 2005. Host-bacterial mutualism in the human intestine. Science 307:1915–1920. [PubMed] 10.1126/science.1104816 [DOI] [PubMed] [Google Scholar]
- 212.Alben JO, Moh PP, Fiamingo FG, Altschuld RA. 1981. Cytochrome Oxidase (a3) Heme and Copper observed by low-temperature Fourier transform infrared spectroscopy of the CO complex. Proc Natl Acad Sci USA 78:234–237. [PubMed] 10.1073/pnas.78.1.234 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 213.Bailey JA, Tomson FL, Mecklenburg SL, MacDonald GM, Katsonouri A, Puustinen A, Gennis RB, Woodruff WH, Dyer RB. 2002. Time-resolved step-scan Fourier transform infrared spectroscopy of the CO adducts of bovine cytochrome c oxidase and of cytochrome bo3 from Escherichia coli. Biochemistry 41:2675–2683. [PubMed] 10.1021/bi010823g [DOI] [PubMed] [Google Scholar]
- 214.Woodruff WH. 1993. Coordination dynamics of heme-copper oxidases. The ligand shuttle and the control and coupling of electron transfer and proton translocation. J Bioenerg Biomemb 25:177–188. [PubMed] 10.1007/BF00762859 [DOI] [PubMed] [Google Scholar]
- 215.Yoshikawa S, Shinzawa-Itoh K, Nakashima R, Yaono R, Yamashita E, Inoue N, Yao M, Fei MJ, Libeu CP, Mizushima T, Yamaguchi H, Tomizaki T, Tsukihara T. 1998. Redox-coupled crystal structural changes in bovine heart cytochrome c oxidase. Science 280:1723–1729. [PubMed] 10.1126/science.280.5370.1723 [DOI] [PubMed] [Google Scholar]
- 216.Butler C, Forte E, Maria Scandurra F, Arese M, Giuffrè A, Greenwood C, Sarti P. 2002. Cytochrome bo3 from Escherichia coli: the binding and turnover of nitric oxide. Biochem Biophys Res Commun 296:1272–1278. [PubMed] 10.1016/S0006-291X(02)02074-0 [DOI] [PubMed] [Google Scholar]
- 217.Ingledew WJ, Horrocks J, Salerno JC. 1993. Ligand binding to the haem-copper binuclear catalytic site of cytochrome bo, a respiratory quinol oxidase from Escherichia coli. Eur J Biochem 212:657–664. [PubMed] 10.1111/j.1432-1033.1993.tb17703.x [DOI] [PubMed] [Google Scholar]
- 218.Van Gelder BF, Beinert H. 1969. Studies of the heme components of cytochrome c oxidase by EPR spectroscopy. Biochim Biophys Acta 189:1–24. [PubMed] 10.1016/0005-2728(69)90219-9 [DOI] [PubMed] [Google Scholar]
- 219.Watmough NJ, Cheesman MR, Butler CS, Little RH, Greenwood C, Thomson AJ. 1998. The dinuclear center of cytochrome bo3 from Escherichia coli. J Bioenerg Biomembr 30:55–62. [PubMed] 10.1023/A:1020507511285 [DOI] [PubMed] [Google Scholar]
- 220.Moody AJ, Mitchell R, Jeal AE, Rich PR. 1997. Comparison of the ligand-binding properties of native and copper-less cytochromes bo from Escherichia coli. Biochem J 324:743–752. [PubMed] 10.1042/bj3240743 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 221.Cheesman MR, Watmough NJ, Gennis RB, Greenwood C, Thomson AJ. 1994. Magnetic-circular-dichroism studies of Escherichia coli cytochrome bo identification of high-spin ferric, low-spin ferric and ferryl [Fe(IV)] forms of heme o. Eur J Biochem 219:595–602. [PubMed] 10.1111/j.1432-1033.1994.tb19975.x [DOI] [PubMed] [Google Scholar]
- 222.Tsubaki M, Mogi T, Hori H, Sato-Watanabe M, Anraku Y. 1996. Infrared and EPR studies on cyanide binding to the heme-copper binuclear center of cytochrome bo-type ubiquinol oxidase from Escherichia coli. J Biol Chem 271:4017–4022. [PubMed] 10.1074/jbc.271.16.9254 [DOI] [PubMed] [Google Scholar]
- 223.Pinakoulaki E, Vamvouka M, Varotsis C. 2004. Resonance Raman detection of the Fe2+-C-N modes in heme-copper oxidases: a probe of the active site. Inorg Chem 43:4907–4910. [PubMed] 10.1021/ic035216r [DOI] [PubMed] [Google Scholar]
- 224.Little RH, Cheesman MR, Thomson AJ, Greenwood C, Watmough NJ. 1996. Cytochrome bo from Escherichia coli: binding of azide to CuB. Biochemistry 35:13780–13787. [PubMed] 10.1021/bi961221d [DOI] [PubMed] [Google Scholar]
- 225.Tsubaki M. 1993. Fourier-transform infrared study of azide binding to the Fea3-CuB binuclear site of bovine heart cytochrome c oxidase: new evidence for a redox-linked conformational change at the binuclear site. Biochemistry 32:174–182. [PubMed] 10.1021/bi00052a023 [DOI] [PubMed] [Google Scholar]
- 226.Tsubaki M, Mogi T, Anraku Y, Hori H. 1993. Structure of the heme-copper binuclear center of the cytochrome bo complex of Escherichia coli: WPR and Fourier transform infrared spectroscopic studies. Biochemistry 32:6065–6072. [PubMed] 10.1021/bi00074a018 [DOI] [PubMed] [Google Scholar]
- 227.Tsubaki M, Mogi T, Hori H. 1999. Fourier-transform infrared studies on azide-binding to the binuclear center of the Escherichia coli bo-type ubiquinol oxidase. FEBS Lett 449:191–195. [PubMed] 10.1016/S0014-5793(99)00423-8 [DOI] [PubMed] [Google Scholar]
- 228.Hubbard JAM, Hughes MN, Poole RK. 1983. Nitrite, but not silver, ions induce spectral changes in Escherichia coli cytochrome d. FEBS Lett 164:241–243. [PubMed] 10.1016/0014-5793(83)80293-2 [DOI] [PubMed] [Google Scholar]
- 229.Hubbard JAM, Hughes MN, Poole RK. 1985. Reactions of some nitrogen oxyanions and nitric oxide with cytochrome oxidase d from oxygen-limited Escherichia coli K12, p 231–236. In Poole RK, Dow CS (ed), Microbial Gas Metabolism: Mechanistic, Metabolic and Biotechnological Aspects. Academic Press, London. [PubMed] [Google Scholar]
- 230.Bonner FT, Hughes MN, Poole RK, Scott RI. 1991. Kinetics of the reactions of trioxodinitrate and nitrite ions with cytochrome d in Escherichia coli. Biochim Biophys Acta 1056:133–138. [PubMed] 10.1016/S0005-2728(05)80279-8 [DOI] [PubMed] [Google Scholar]
- 231.Jünemann S. 1997. Cytochrome bd terminal oxidase. Biochim Biophys Acta 1321:107–127. [PubMed] 10.1016/S0005-2728(97)00046-7 [DOI] [PubMed] [Google Scholar]
- 232.Sarti P, Giuffrè A, Forte E, Mastronicola D, Barone MC, Brunori M. 2000. Nitric oxide and cytochrome c oxidase: mechanisms of inhibition and NO degradation. Biochem Biophys Res Commun 274:183–187. [PubMed] 10.1006/bbrc.2000.3117 [DOI] [PubMed] [Google Scholar]
- 233.Lemon DD, Calhoun MW, Gennis RB, Woodruff WH. 1993. The gateway to the active site of heme-copper oxidases. Biochemistry 32:11953–11956. [PubMed] 10.1021/bi00096a002 [DOI] [PubMed] [Google Scholar]
- 234.Pudek MR, Bragg PD. 1974. Inhibition by cyanide of the respiratory chain oxidases of Escherichia coli. Arch Biochem Biophys 164:682–693. [PubMed] 10.1016/0003-9861(74)90081-2 [DOI] [PubMed] [Google Scholar]
- 235.Krasnoselskaya I, Arutjunjan AM, Smirnova I, Gennis R, Konstantinov AA. 1993. Cyanide-reactive sites in cytochrome bd complex from E. coli. FEBS Lett 327:279–283. [PubMed] 10.1016/0014-5793(93)81004-J [DOI] [PubMed] [Google Scholar]
- 236.Poole RK, Williams HD. 1988. Formation of the 680-nm-absorbing form of the cytochrome bd oxidase complex of Escherichia coli by reaction of hydrogen peroxide with the ferric form. FEBS Lett 231:243–246. [PubMed] 10.1016/0014-5793(88)80740-3 [DOI] [PubMed] [Google Scholar]
- 237.Borisov V, Gennis R, Konstantinov AA. 1995. Peroxide complex of cytochrome bd: kinetics of generation and stability. Biochem Mol Biol Int 37:975–982. [PubMed] [PubMed] [Google Scholar]
- 238.Sun J, Osborne JP, Kahlow MA, Kaysser TM, Gennis RB, Loehr TM. 1995. Resonance Raman studies of Escherichia coli cytochrome bd oxidase. Selective enhancement of the three heme chromophores of the “as-isolated” enzyme and characterization of the cyanide adduct. Biochemistry 34:12144–12151. [PubMed] 10.1021/bi00038a007 [DOI] [PubMed] [Google Scholar]
- 239.Wikström M. 1977. Proton pump coupled to cytochrome c oxidase in mitochondria. Nature 266:271–273. [PubMed] 10.1038/266271a0 [DOI] [PubMed] [Google Scholar]
- 240.Puustinen A, Finel M, Virkki M, Wikström M. 1989. Cytochrome o (bo) is a proton pump in Paracoccus denitrificans and Escherichia coli. FEBS Lett 249:163–167. [PubMed] 10.1016/0014-5793(89)80616-7 [DOI] [PubMed] [Google Scholar]
- 241.Yap LL, Samoilova RI, Gennis RB, Dikanov SA. 2006. Characterization of the exchangeable protons in the immediate vicinity of the semiquinone radical at the QH site of the cytochrome bo3 from Escherichia coli. J Biol Chem 281:16879–16887. [PubMed] 10.1074/jbc.M602544200 [DOI] [PubMed] [Google Scholar]
- 242.Kobayashi K, Tagawa S, Mogi T. 2000. Transient formation of ubisemiquinone radical and subsequent electron transfer process in the Escherichia coli cytochrome bo. Biochemistry 39:15620–15625. [PubMed] 10.1021/bi0014094 [DOI] [PubMed] [Google Scholar]
- 243.Belevich I, Verkhovsky MI, Wikström M. 2006. Proton-coupled electron transfer drives the proton pump of cytochrome c oxidase. Nature 440:829–832. [PubMed] 10.1038/nature04619 [DOI] [PubMed] [Google Scholar]
- 244.Morgan JE, Verkhovsky MI, Puustinen A, Wikström M. 1993. Intramolecular electron transfer in cytochrome o of Escherichia coli: events following the photolysis of fully and partially reduced CO-bound forms of the bo3 and oo3 enzymes. Biochemistry 32:11413–11418. [PubMed] 10.1021/bi00093a019 [DOI] [PubMed] [Google Scholar]
- 245.Jasaitis A, Johansson MP, Wikström M, Vos MH, Verkhovsky MI. 2007. Nanosecond electron tunneling between the hemes in cytochrome bo3. Proc Natl Acad Sci USA 104:20811–20814. [PubMed] 10.1073/pnas.0709876105 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 246.Oliveberg M, Malmstrцm BG. 1991. Internal electron transfer in cytochrome c oxidase: evidence for a rapid equilibrium between cytochrome a and the bimetallic site. Biochemistry 30:7053–7057. [PubMed] 10.1021/bi00243a003 [DOI] [PubMed] [Google Scholar]
- 247.Hallen S, Brzezinski P, Malmstrom BG. 1994. Internal electron transfer in cytochrome c oxidase is coupled to the protonation of a group close to the bimetallic site. Biochemistry 33:1467–1472. [PubMed] 10.1021/bi00172a024 [DOI] [PubMed] [Google Scholar]
- 248.Moser CC, Keske JM, Warncke K, Farid RS, Dutton PL. 1992. Nature of biological electron transfer. Nature 355:796–802. [PubMed] 10.1038/355796a0 [DOI] [PubMed] [Google Scholar]
- 249.Page CC, Moser CC, Dutton PL. 2003. Mechanism for electron transfer within and between proteins. Curr Opin Chem Biol 7:551–556. [PubMed] 10.1016/j.cbpa.2003.08.005 [DOI] [PubMed] [Google Scholar]
- 250.Svensson-Ek M, Thomas JW, Gennis RB, Nilsson T, Brzezinski P. 1996. Kinetics of electron and proton transfer during the reaction of wild type and helix VI mutants of cytochrome bo3 with oxygen. Biochemistry 35:13673–13680. [PubMed] 10.1021/bi961466q [DOI] [PubMed] [Google Scholar]
- 251.Verkhovskaya ML, Garcia-Horsman A, Puustinen A, Rigaud J-L, Morgan JE, Verkhovsky MI, Wikström M. 1997. Glutamic acid 286 in subunit I of cytochrome bo3 is involved in proton translocation. Proc Natl Acad Sci USA 94:10128–10131. [PubMed] 10.1073/pnas.94.19.10128 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 252.Watmough NJ, Katsonouri A, Little RH, Osborne JP, Furlong-Nickels E, Gennis RB, Brittain T, Greenwood C. 1997. A conserved glutamic acid in helix VI of cytochrome bo3 influences a key step in oxygen reduction. Biochemistry 36:13736–13742. [PubMed] 10.1021/bi971434i [DOI] [PubMed] [Google Scholar]
- 253.Shinzawa-Itoh K, Aoyama H, Muramoto K, Terada H, Kurauchi T, Tadehara Y, Yamasaki A, Sugimura T, Kurono S, Tsujimoto K, Mizushima T, Yamashita E, Tsukihara T, Yoshikawa S. 2007. Structures and physiological roles of 13 integral lipids of bovine heart cytochrome c oxidase. EMBO J 26:1713–1725. [PubMed] 10.1038/sj.emboj.7601618 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 254.Qin L, Hiser C, Mulichak A, Garavito RM, Ferguson-Miller S. 2006. Identification of conserved lipid/detergent-binding sites in a high-resolution structure of the membrane protein cytochrome c oxidase. Proc Natl Acad Sci USA 103:16117–16122. [PubMed] 10.1073/pnas.0606149103 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 255.Morgan JE, Verkhovsky MI, Wikström M. 1994. The histidine cycle: a new model for proton translocation in the respiratory heme-copper oxidases. J Bioenerg Biomembr 26:599–608. [PubMed] 10.1007/BF00831534 [DOI] [PubMed] [Google Scholar]
- 256.Wikström M, Bogachev A, Finel M, Morgan JE, Puustinen A, Raitio M, Verkhovskaya M, Verkhovsky MI. 1994. Mechanism of proton translocation by the respiratory oxidases. The histidine cycle. Biochim Biophys Acta 1187:106–111. [PubMed] 10.1016/0005-2728(94)90093-0 [DOI] [PubMed] [Google Scholar]
- 257.Iwata S, Ostermeier C, Ludwig B, Michel H. 1995. Structure at 2.8 Å resolution of cytochrome c oxidase from Paracoccus denitrificans. Nature 376:660–669. [PubMed] 10.1038/376660a0 [DOI] [PubMed] [Google Scholar]
- 258.Konstantinov AA, Siletsky S, Mitchell D, Kaulen A, Gennis RB. 1997. The roles of the two proton input channels in cytochrome c oxidase from Rhodobacter sphaeroides probed by the effects of site-directed mutations on time-resolved electrogenic intraprotein proton transfer. Proc Natl Acad Sci USA 94:9085–9090. [PubMed] 10.1073/pnas.94.17.9085 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 259.Hosler JP, Shapleigh JP, Kim Y, Pressler M, Georgiou C, Babcock GT, Alben JO, Ferguson-Miller S, Gennis RB. 1996. Polar residues in helix VIII of subunit I of cytochrome c oxidase influence the activity and the structure of the active site. Biochemistry 35:10776–10783. [PubMed] 10.1021/bi9606511 [DOI] [PubMed] [Google Scholar]
- 260.Wikström M, Jasaitis A, Backgren C, Puustinen A, Verkhovsky MI. 2000. The role of the D- and K-pathways of proton transfer in the function of the haem-copper oxidases. Biochim Biophys Acta 1459:514–520. [PubMed] 10.1016/S0005-2728(00)00191-2 [DOI] [PubMed] [Google Scholar]
- 261.Tsukihara T, Shimokata K, Katayama Y, Shimada H, Muramoto K, Aoyama H, Mochizuki M, Shinzawa-Itoh K, Yamashita E, Yao M, Ishimura Y, Yoshikawa S. 2003. The low-spin heme of cytochrome c oxidase as the driving element of the proton-pumping process. Proc Natl Acad Sci USA 100:15304–15309. [PubMed] 10.1073/pnas.2635097100 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 262.Puustinen A, Wikström M. 1999. Proton exit from the heme-copper oxidase of Escherichia coli. Proc Natl Acad Sci USA 96:35–37. [PubMed] 10.1073/pnas.96.1.35 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 263.Verkhovsky MI, Morgan JE, Puustinen A, Wikström M. 1996. The “ferrous-oxy” intermediate in the reaction of dioxygen with fully reduced cytochromes aa3 and bo3. Biochemistry 35:16241–16246. [PubMed] 10.1021/bi961433a [DOI] [PubMed] [Google Scholar]
- 264.Proshlyakov DA, Pressler MA, Babcock GT. 1998. Dioxygen activation and bond cleavage by mixed-valence cytochrome c oxidase. Proc Natl Acad Sci USA 95:8020–8025. [PubMed] 10.1073/pnas.95.14.8020 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 265.Fabian M, Wong WW, Gennis RB, Palmer G. 1999. Mass spectrometric determination of dioxygen bond splitting in the “peroxy” intermediate of cytochrome c oxidase. Proc Natl Acad Sci USA 96:13114–13117. [PubMed] 10.1073/pnas.96.23.13114 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 266.Morgan JE, Verkhovsky MI, Puustinen A, Wikström M. 1995. Identification of a “peroxy” intermediate in cytochrome bo3 of Escherichia coli. Biochemistry 34:15633–15637. [PubMed] 10.1021/bi00048a005 [DOI] [PubMed] [Google Scholar]
- 267.Hirota S, Mogi T, Ogura T, Hirano T, Anraku Y, Kitagawa T. 1994. Observation of the Fe-O2 and FeIV=0 stretching Raman bands for dioxygen reduction intermediates of cytochrome bo isolated from Escherichia coli. FEBS Lett 352:67–70. [PubMed] 10.1016/0014-5793(94)00919-8 [DOI] [PubMed] [Google Scholar]
- 268.Wang J, Rumbley J, Ching YC, Takahashi S, Gennis RB, Rousseau DL. 1995. Reaction of cytochrome bo3 with oxygen: extra redox center(s) are present in the protein. Biochemistry 34:15504–15511. [PubMed] 10.1021/bi00047a016 [DOI] [PubMed] [Google Scholar]
- 269.Puustinen A, Verkhovsky MI, Morgan JE, Belevich NP, Wikström M. 1996. Reaction of the Escherichia coli quinol oxidase cytochrome bo3 with dioxygen: the role of a bound ubiquinone molecule. Proc Natl Acad Sci USA 93:1545–1548. [PubMed] 10.1073/pnas.93.4.1545 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 270.Orii Y, Mogi T, Kawasaki M, Anraku Y. 1994. Facilitated intramolecular electron transfer in cytochrome bo-type ubiquinol oxidase initiated upon reaction of the fully reduced enzyme with dioxygen. FEBS Lett 352:151–154. [PubMed] 10.1016/0014-5793(94)00939-2 [DOI] [PubMed] [Google Scholar]
- 271.Svensson Ek M, Brzezinski P. 1997. Oxidation of ubiquinol by cytochrome bo3 from Escherichia coli: kinetics of electron and proton transfer. Biochemistry 36:5425–5431. [PubMed] 10.1021/bi962478e [DOI] [PubMed] [Google Scholar]
- 272.Kinoshita N, Unemoto T, Kobayashi H. 1984. Proton motive force is not obligatory for growth of Escherichia coli. J Bacteriol 160:1074–1077. [PubMed] [DOI] [PMC free article] [PubMed] [Google Scholar]
- 273.Verkhovskaya ML, Verkhovsky MI, Wikström M. 1996. K+-dependent Na+ transport driven by respiration in Escherichia coli cells and membrane vesicles. Biochim Biophys Acta 1273:207–216. [PubMed] 10.1016/0005-2728(95)00142-5 [DOI] [PubMed] [Google Scholar]
- 274.van Verseveld HW, Krab K, Stouthamer AH. 1981. Proton pump coupled to cytochrome c oxidase in Paracoccus denitrificans. Biochim Biophys Acta 635:525–534. [PubMed] 10.1016/0005-2728(81)90111-0 [DOI] [PubMed] [Google Scholar]
- 275.Bloch D, Belevich I, Jasaitis A, Ribacka C, Puustinen A, Verkhovsky MI, Wikström M. 2004. The catalytic cycle of cytochrome c oxidase is not the sum of its two halves. Proc Natl Acad Sci USA 101:529–533. [PubMed] 10.1073/pnas.0306036101 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 276.Verkhovsky MI, Belevich I, Bloch DA, Wikström M. 2006. Elementary steps of proton translocation in the catalytic cycle of cytochrome oxidase. Biochim Biophys Acta 1757:401–407. [PubMed] 10.1016/j.bbabio.2006.05.026 [DOI] [PubMed] [Google Scholar]
- 277.Popovic DM, Stuchebrukhov AA. 2004. Proton pumping mechanism and catalytic cycle of cytochrome c oxidase: coulomb pump model with kinetic gating. FEBS Lett 566:126–130. [PubMed] 10.1016/j.febslet.2004.04.016 [DOI] [PubMed] [Google Scholar]
- 278.Siegbahn PEM, Blomberg MRA, Blomberg ML. 2003. Theoretical study of the energetics of proton pumping and oxygen reduction in cytochrome oxidase. J Phys Chem B 107:10946–10955. 10.1021/jp035486v [DOI] [Google Scholar]
- 279.Siletsky SA, Pawate AS, Weiss K, Gennis RB, Konstantinov AA. 2004. Transmembrane charge separation during the ferryl-oxo -> oxidized transition in a nonpumping mutant of cytochrome c oxidase. J Biol Chem 279:52558–52565. [PubMed] 10.1074/jbc.M407549200 [DOI] [PubMed] [Google Scholar]
- 280.Wikström M, Verkhovsky MI. 2007. Mechanism and energetics of proton translocation by the respiratory heme-copper oxidases. Biochim Biophys Acta 1767:1200–1214. [PubMed] 10.1016/j.bbabio.2007.06.008 [DOI] [PubMed] [Google Scholar]
- 281.Faxen K, Gilderson G, Adelroth P, Brzezinski P. 2005. A mechanistic principle for proton pumping by cytochrome c oxidase. Nature 437:286–289. [PubMed] 10.1038/nature03921 [DOI] [PubMed] [Google Scholar]
- 282.Koland JG, Miller MJ, Gennis RB. 1984. Reconstitution of the membrane-bound, ubiquinone-dependent pyruvate oxidase respiratory chain of Escherichia coli with the cytochrome d terminal oxidase. Biochemistry 23:445–453. [PubMed] 10.1021/bi00298a008 [DOI] [PubMed] [Google Scholar]
- 283.Borisov VB, Forte E, Sarti P, Giuffrè A. 2011. Catalytic intermediates of cytochrome bd terminal oxidase at steady-state: Ferryl and oxy-ferrous species dominate. Biochim Biophys Acta 1807:503–509. [PubMed] 10.1016/j.bbabio.2011.02.007 [DOI] [PubMed] [Google Scholar]
- 284.Gibson Q, Greenwood C. 1963. Reactions of cytochrome oxidase with oxygen and carbon monoxide. Biochem J 86:541–555. [PubMed] 10.1042/bj0860541 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 285.Paulus A, Rossius SG, Dijk M, de Vries S. 2012. Oxoferryl-porphyrin radical catalytic intermediate in cytochrome bd oxidases protects cells from formation of reactive oxygen species. J Biol Chem 287:8830–8838. [PubMed] 10.1074/jbc.M111.333542 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 286.Al-Attar S, de Vries S. 2013. Energy transduction by respiratory metallo-enzymes: from molecular mechanism to cell physiology. Coord Chem Rev 257:64–80. 10.1016/j.ccr.2012.05.022 [DOI] [Google Scholar]
- 287.Yang K, Borisov VB, Konstantinov AA, Gennis RB. 2008. The fully oxidized form of the cytochrome bd quinol oxidase from E. coli does not participate in the catalytic cycle: direct evidence from rapid kinetics studies. FEBS Lett 582:3705–3709. [PubMed] 10.1016/j.febslet.2008.09.038 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 288.Dassa J, Fsihi H, Marck C, Dion M, Kieffer-Bontemps M, Boquet PL. 1991. A new oxygen-regulated operon in Escherichia coli comprises the genes for a putative third cytochrome oxidase and for pH 2.5 acid phosphatase (appA). Mol Gen Genet 229:341–352. [PubMed] 10.1007/BF00267454 [DOI] [PubMed] [Google Scholar]
- 289.Atlung T, Brondsted L. 1994. Role of the transcriptional activator AppY in regulation of the cyxappA operon of Escherichia coli by anaerobiosis, phosphate starvation, and growth phase. J Bacteriol 176:5414–5422. [PubMed] [DOI] [PMC free article] [PubMed] [Google Scholar]
- 290.Brondsted L, Atlung T. 1996. Effect of growth conditions on expression of the acid phosphatase (cyx-appA) operon and the appY gene, which encodes a transcriptional activator of Escherichia coli. J Bacteriol 178:1556–1564. [PubMed] [DOI] [PMC free article] [PubMed] [Google Scholar]
- 291.Sturr MG, Krulwich TA, Hicks DB. 1996. Purification of a cytochrome bd terminal oxidase encoded by the Escherichia coli app locus from a Δcyo Δcyd strain complemented by genes from Bacillus firmus OF4. J Bacteriol 176:1742–1749. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 292.Bekker M, de Vries S, Ter Beek A, Hellingwerf KJ, de Mattos MJ. 2009. Respiration of Escherichia coli can be fully uncoupled via the nonelectrogenic terminal cytochrome bd-II oxidase. J Bacteriol 191:5510–5517. [PubMed] 10.1128/JB.00562-09 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 293.Shepherd M, Sanguinetti G, Cook GM, Poole RK. 2010. Compensations for diminished terminal oxidase activity in Escherichia coli: cytochrome bd-II-mediated respiration and glutamate metabolism. J Biol Chem 285:18464–18472. [PubMed] 10.1074/jbc.M110.118448 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 294.Sharma P, Hellingwerf KJ, Teixeira de Mattos MJ, Bekker M. 2012. Uncoupling of substrate-level phosphorylation in Escherichia coli during glucose-limited growth. Appl Environ Microbiol 78:6908–6913. [PubMed] 10.1128/AEM.01507-12 [DOI] [PMC free article] [PubMed] [Google Scholar]
