Previous Article in Journal
Advancing Precision Neurology and Wearable Electrophysiology: A Review on the Pivotal Role of Medical Physicists in Signal Processing, AI, and Prognostic Modeling
Previous Article in Special Issue
Protein Polarimetry, Perfected: Specific Rotation Measurement for the Refracto-Polarimetric Detection of Cryptic Protein Denaturation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Resistance of Nitric Oxide Dioxygenase and Cytochrome c Oxidase to Inhibition by Nitric Oxide and Other Indications of the Spintronic Control of Electron Transfer

by
Paul R. Gardner
1,2
1
Research and Development Division, Miami Valley Biotech, Dayton, OH 45402, USA
2
Division of Critical Care Medicine, Cincinnati Children’s Hospital Medical Center, Cincinnati, OH 45229, USA
Biophysica 2025, 5(3), 41; https://doi.org/10.3390/biophysica5030041
Submission received: 31 July 2025 / Revised: 3 September 2025 / Accepted: 6 September 2025 / Published: 9 September 2025
(This article belongs to the Special Issue Investigations into Protein Structure)

Abstract

Heme enzymes that bind and reduce O2 are susceptible to poisoning by NO. The high reactivity and affinity of NO for ferrous heme produces stable ferrous-NO complexes, which in theory should preclude O2 binding and turnover. However, NO inhibition is often competitive with respect to O2 and rapidly reversible, thus providing cellular and organismal survival advantages. This kinetic paradox has prompted a search for mechanisms for reversal and hence resistance. Here, I critically review proposed resistance mechanisms for NO dioxygenase (NOD) and cytochrome c oxidase (CcO), which substantiate reduction or oxidation of the tightly bound NO but nevertheless fail to provide kinetically viable solutions. A ferrous heme intermediate is clearly not available during rapid steady-state turnover. Reversible inhibition can be attributed to NO competing with O2 in transient low-affinity interactions with either the ferric heme in NOD or the ferric heme-cupric center in CcO. Toward resolution, I review the underlying principles and evidence for kinetic control of ferric heme reduction via an O2-triggered ferric heme spin crossover and an electronically-forced motion of the heme and structurally-linked protein side chains that elicit electron transfer and activate O2 in the flavohemoglobin-type NOD. For CcO, kinetics, structures, and density functional theory point to the existence of an analogous O2 and reduced oxygen intermediate-controlled electron-transfer gate with a linked proton pump function. A catalytic cycle and mechanism for CcO is finally at hand that links each of the four O2-reducing electrons to each of the four pumped protons in time and space. A novel proton-conducting tunnel and channel, electron path, and pump mechanism, most notably first hypothesized by Mårten Wikström in 1977 and pursued since, are laid out for further scrutiny. In both models, low-energy spin-orbit couplings or ‘spintronic’ interactions with O2 and NO or copper trigger the electronic motions within heme that activate electron transfer to O2, and the exergonic reactions of transient reactive oxygen intermediates ultimately drive all enzyme, electron, and proton motions.

1. Background and Introduction

1.1. NO Toxicity and Inhibition of Heme Enzymes

NO has been known to be toxic to living organisms since the synthesis and discovery of the gas by Sir Humphrey Davy well over 200 years ago [1]. NO, and the orangish-brownish NO2 gas formed in the reaction of NO with atmospheric O2, was quickly recognized as harmful and not euphoric, anesthetic or analgesic, and seemingly innocuous [2], like N2O. Nevertheless, NO serves as an important intermediate in microbial denitrification [3] and, at high nanomolar steady-state concentrations, requires detoxification by NO reductases (NORs) and NO dioxygenases (NODs) [4,5]. Mammals also synthesize NO, which is ultimately oxidized to form nitrate and nitrite under normal physiological conditions. NO is released by leukocytes and is elevated in various pathological conditions involving inflammation [6,7,8,9,10]. NO is produced primarily by inducible NO synthases (NOSs) [11], secondarily via enzymatic and non-enzymatic reduction of NO2 [12], or pharmacologically through the metabolism of vasorelaxants such as nitroglycerin and isoamyl nitrite [13]. Indeed, an understanding of the physiology of nitroglycerin, first described by the early discovery of Ascanio Sobrero and subsequent research of Constantin Hering and coworkers in the mid-19th century [14], has finally been achieved. The elucidation of the role and mechanism of NO or the ‘endothelium-derived relaxing factor’ in regulating blood pressure was recognized with the Nobel Prize in Medicine or Physiology in 1998 [15]. Similar to microbes, mammals, and apparently all life forms, metabolize and detoxify NO [5,16]. Hemoglobin (Hb) and myoglobin (Mb) are important and common scavengers of NO that catalytically dioxygenate NO to form nitrate [4,17]. When tightly-coupled to reductases, globins enzymatically dioxygenate NO [4,5,17,18]. It is noteworthy that the non-enzymatic reaction of NO with oxyHb was first reported by Sir Humphrey Davy over 200 years ago, with the visible browning of red blood upon exposure [1], and was further characterized by Doyle and Hoekstra in 1981 [19]. OxyHb and oxyMb react with NO with measured bimolecular rate constants in the range of 2 to 4 × 107 M−1 s−1 [19,20,21]. Short-lived caged intermediates in an O-atom rearrangement mechanism are hypothesized [17,20,21,22,23].
NO inhibits or inactivates enzymes either directly or indirectly through NO-derived reactive species, including ONOO, NO2, +NO, or NO [24,25]. At sufficient concentrations, NO can inactivate or activate metalloenzymes, thiol-dependent enzymes, di-iron enzymes, and tyrosine-dependent enzymes within living organisms. Iron-sulfur [9,26,27,28] and heme-containing enzymes, including cytochrome c oxidase (CcO) [29,30] and cytochrome P450s [31,32,33], were identified early on as sensitive targets for NO toxicity. Several NO-sensitive iron-sulfur and heme targets have been found to be utilized as NO sensors for diverse functional roles by various organisms. Indeed, blood pressure regulation ultimately occurs due to ferrous heme-NO formation that activates the soluble guanylate cyclase [11,34]. NO impacts mammalian iron metabolism via reactions with the cytosolic aconitase iron-sulfur center [35,36]. In some scenarios, reversible NO inhibition of CcO may serve a beneficial function by maximizing tissue O2 distribution or sensing O2 [37,38].

1.2. Cellular NO Detoxification Pathways, NODs, NORs, and Other Pathways

Similar to the evolution of NO-sensing functions, it should not be surprising that sensitive and critical targets of NO toxicity have forced the evolution of mechanisms for NO resistance, including damage repair. Furthermore, it should not be unexpected that an NO resistance mechanism would provide a chemical and genetic route for the evolution of enzymes specialized for NO metabolism and detoxification. According to the dictum eloquently stated by Max Perutz [39] and reiterated by Matti Saraste and others [40,41], evolution transforms one system to produce another through small changes based on a shared chemistry. The NO and O2 metabolizing enzymes (i.e., NODs, NORs, and O2 reductases), sensors, or O2 transport-storage proteins are perfect examples where this is made abundantly clear. A proviso to this condition is that any progress in understanding observed reactions and other phenomena requires the discernment of genuine biological functions among shared chemistries via comprehensive studies of kinetics, mechanism, and genetics in the context of physiology. The effort demands a critical examination of many, if not all, of the empirical observations, possible interpretations, a recognition of the reproducible and consistent patterns, and the contradictions. Only after extensive searching and ‘wandering’ [42] can a logical essence or understanding emerge from ‘the gardens of the mind’ [43]. Understanding the function and evolution of NO and O2 systems has been one focus and goal of my life’s work and of this review.

1.3. The Paradoxical Function(s) of flavoHbs in NO Detoxification

Evidence for an O2-dependent nitrate-producing NO metabolic enzyme in soil microbes, including Escherichia coli (abbreviated as E. coli), was reported by Ralf Conrad and coworkers in a series of publications from 1989 to 1996 [44,45,46,47]. Unfamiliar with Conrad and coworkers’ publications, my group observed an NO-inducible, O2-dependent, and cyanide-sensitive NO metabolic activity in E. coli while investigating the NO-sensitive citric acid cycle enzyme aconitase [48,49]. In cell-free extracts, the activity required NAD(P)H, O2, and FAD and generated nitrate. The NOD activity was isolated, sequenced, and identified with flavoHb [49]. The requirement for FAD for NOD activity was unexpected and troublesome, but it was advantageous for FAD affinity chromatography and rapid isolation. Similar to the structurally related cytochrome P450 reductase [50], the FAD cofactor is apparently released from the reductase domain during ammonium sulfate fractionation but readily rebinds and reconstitutes the apo-enzyme [51].
A simple enzymatic cycle was readily proposed based on the well-known reaction of NO with oxy-globins to produce nitrate (Equations (1) and (2)) [19,20,21,52]. NAD(P)H reduced the FAD in the reductase domain (Equation (3)), which reduced the ferric heme in the globin domain (Equation (4)) with one NAD(P)H consumed per two NO molecules (Equations (5) and (6)). Hausladen et al. independently demonstrated the inducible and protective NO metabolic activity of E. coli flavoHb, reported the reaction stoichiometry (Equation (6)), and briefly decided [53] (see Section 3.3.2 below) in favor of the NOD mechanism (Equations (1)–(6)) [54].
O2 + Fe2+ ⇌ Fe3+O2
Fe3+O2 + NO → Fe3+ + NO3
NAD(P)H + H+ + FAD → FADH2 + NAD(P)+
FADH2 + Fe3+ → Fe2+ + FADH + H+
FADH + Fe3+ → Fe2+ + FAD + H+
2 NO + 2 O2 + NAD(P)H → 2 NO3 + NAD(P)+ + 2 H+
Several earlier investigations had pointed to a role for flavoHbs in NO metabolism or detoxification. Numerous investigations have subsequently supported that biological role [5,55,56,57]. However, the mechanism, and thus molecular function, continues to be debated [56]. Poole and coworkers had shown that E. coli flavoHb is induced by nitrite or by bolus NO addition and postulated a role in nitrogen oxide metabolism and a NOR function [58]. Nakano also reported increased flavoHb expression in Bacillus subtilis with growth on nitrite and postulated a role in nitrite reduction [59]. Bärbel Friedrich’s lab had determined earlier that the Alcaligenes eutrophus (aka Ralstonia eutropha) flavoHb bore no detectable NOR activity but, nevertheless, was essential for NO reduction to N2O during denitrification [60]. The high NO affinity of the ferrous heme and a rapid autooxidation of the oxy complex of the yeast and A. eutrophus flavoHbs purportedly discouraged Austen Riggs, Hao Zhu, John Olson, and Yi Dou, and likely others, from an earlier foray into investigations of flavoHb NO metabolic activity [61,62] involving an ‘NO-induced oxidation’ similar to oxyHb or oxyMb [19,20]. In our eventual collaboration with Riggs, Zhu, Olson, Li, and Dou, the two flavoHbs, along with the E. coli flavoHb, were analyzed side-by-side and were found to have remarkably similar transient ligand binding kinetics and steady-state NOD kinetics [62]. A ping-pong mechanism with O2 first binding the ferrous heme and NO reacting with the complex was not entirely supported by the data. The problem of high-affinity NO binding to ferrous heme in place of O2 in the proposed NOD mechanism (Equation (1)) was now patently evident [51,62], but the contradiction or paradox was never resolved in our joint efforts. With the prohibitively high-avidity of NO for the ferrous heme now measured and published, Hausladen et al. readily argued against the NOD mechanism (Equations (1)–(6)). A more favored alternative ‘denitrosylase’ mechanism with a reverse binding order for NO and O2 was argued, and a new enzyme nomenclature was suggested [53,63] (Equations (7) and (8)).
NO + Fe2+ ⇌ Fe3+NO
Fe3+NO + O2 → Fe3+ + NO3−
Further confounding the mechanism and function of flavoHb, an ancillary anoxic NOR activity that was demonstrated by Kim et al. [64,65] but was contradicted by the observations of Cramm et al. [60] was put forth to explain several anaerobic induction and protection phenomena. Indeed, the phenomena continue to point to NO-related functions for flavoHb that are both unrelated to NO dioxygenation and occur in the absence of any demonstrable NOR activity of flavoHb within cells [66]. Anoxic NO-dependent physiological effects invariably continue to lead to the supposition of NO reduction [60,67]. However, the absence of trace and infinitesimal O2 is rarely, if ever, demonstrated. In this regard, an NOD activity consumes both NO and O2, and either O2 or NO may be important under hypoxic conditions in the absence of other more dominant O2 or NO consumption activities. NO dioxygenation by globins may remove O2 and protect O2-sensitive targets in legumes [68], in Ascaris [69], or increase NO reduction by adventitious NO/O2 reductases, including some that are readily inactivated by O2 [70]. Under microaerophilic conditions, trace O2 may inhibit the dedicated NOR while NO may inhibit the dedicated O2 reductase. An overexpressed NOD activity would partly relieve both inhibitions by contributing to NO and O2 depletion. A sulfur trioxide anion radical scavenging function has also been demonstrated for flavoHbs [71] and may explain phenomena related to oxidative stress, such as the protection of aerobic E. coli against paraquat toxicity [72] or yeast against copper toxicity [73]. Notwithstanding the numerous intriguing, yet unexplained, physiological phenomena, progress in deciphering the biological NO detoxification mechanism(s) and molecular function(s) of flavoHbs and related single domain Hbs continues to be achieved through kinetic, structural, and other biophysical investigations [18,74].
The NOD vs. denitrosylase mechanism has been debated for well over two decades [5,53,56,63], with the concept of an NO reduction-dependent denitrosylase mechanism and dual aerobic–anaerobic function sporadically relegating the concept of a limited aerobic NOD mechanism and function [56,67]. Nevertheless, progress on many fronts has illuminated an NOD mechanism with subtle, albeit unorthodox, interactions of the ferric heme with O2 and NO [18,74,75] that may settle the lingering controversy and confusion. The revised NOD mechanism depends upon spin-orbit coupling (SOC) phenomena, which I dub ‘spintronics’ in close analogy to the quantum principles recently deployed in spintronic devices, including those being designed for efficient electrochemical O2 reduction in fuel cells [76,77,78]. The proposed mechanism also employs motions similar to those demonstrated long ago in cooperative O2 binding to red blood cell Hb and allostery. In Section 3.3.4, I focus on the essential role of spintronics and motions for the NOD mechanism. It is hoped that greater knowledge of the NOD mechanism may ultimately be applied to the design of targeted antibiotics [79] and other therapies or, perhaps in the distant future, to bioengineering spintronic devices.

1.4. The Paradoxical NO Resistance of CcO and Other O2 Reductases

Observations of O2-dependent, cyanide-sensitive, and protective NO metabolism by E. coli and various mammalian cell lines and tissues originally prompted my lab to investigate the possible role of the terminal oxidases (i. e. CcO, cytochrome bo, and other O2 reductases) in NO metabolism within E. coli [48] and mammalian cells [16]. A role for the mitochondrial CcO (aka cytochrome aa3) in cellular NO metabolism had been hypothesized [80,81]. And, at the time, the role of globin-type NOD activities in NO metabolism discussed above in Section 1.3 had not yet been discovered. The exquisite NO sensitivity of cytochrome aa3 was confirmed, but a role for cytochrome aa3 in NO metabolism was not substantiated [16]. Nor did the E. coli cytochrome bo or cytochrome bd catalyze measurable NO metabolism within E. coli [48]. Cyanide was without effect on the respiration of E. coli at levels that clearly inhibited NO detoxification [48,82]. Moreover, E. coli strains deficient in either of the two terminal respiratory oxidases showed normal levels of the constitutive and inducible aerobic NO consumption activities, whereas flavoHb-deficient E. coli express essentially no constitutive or inducible aerobic NO consumption activity [49]. NO metabolism by some mammalian cells is remarkably robust [16], suggesting independent enzymes. The ratio of NO and O2 consumption rates of various cultured mammalian cell lines under normoxia ranges from 0.1 to 1.1. Similar to O2 respiration, NO consumption is cyanide-sensitive [16,80]. However, inhibition of O2 respiration and the function of the cyanide-sensitive cytochrome aa3, with rotenone, antimycin A, myxothiazole, and oligomycin did not proportionally inhibit the NO consumption activity, thus indicating little, if any, role for cytochrome aa3 in NO metabolism, contrary to what Moncada and coworkers argued [83]. Furthermore, cells showed an O2-independent NO consumption activity that equaled 10–20% of the aerobic activity, suggesting the expression of an unidentified NOR, perhaps similar to the eukaryotic P450-type NOR [84]. Purportedly, cytochrome aa3 does not function within cells as an NOR [85] as early measurements with the purified enzyme suggested [86]. More recently, Giles et al. demonstrated a lagging and CcO-independent NO metabolic activity of cardiac mitochondria with an estimated turnover of 1.3 s−1 relative to CcO [87].
NO scavenging by globin-type NODs, various NORs, or reactive oxygen species now appears as an important mechanism for the NO resistance of terminal oxidases in microbes [70,82] and mammals [16,18,87,88]. However, this limited view ignores the important issue of what happens when terminal oxidases, or other metabolic pathways (e.g., cytochrome P450s), are indeed inhibited or essentially inactivated by excess NO accompanying inflammation (e.g., septic shock).
Several O2 reductases, including the mitochondrial cytochrome aa3, bind NO with high affinity [89,90] and are very sensitive to competitive inhibition by NO [16,29,30,82,91,92,93,94,95,96]. But, as already discussed, these O2 reductases do not robustly metabolize NO within cells [16,48,87]. Other O2 reductases, such as the di-heme cytochrome bd, are more resistant to NO inhibition [96,97,98,99]. Indeed, the cytochrome cbb3 does reduce and metabolize NO in addition to O2 [100]. In the case of heme-copper oxidases, including cytochrome aa3 (CcO) and cytochrome bo3, NO inhibition is rapidly reversible [29,37,94]. Yet, for these O2 reductases, measured association and dissociation rate constants suggest that NO should bind tightly to the ferrous-heme intermediate in place of O2 and inhibit the enzymes irreversibly. This underrecognized contradiction and paradox suggests either the absence of a ferrous-heme intermediate for NO and O2 binding during normal cellular catalysis or a very rapid release of NO. A review of the investigations of mechanisms of NO inhibition and reversal provides potential explanations involving NO reduction or oxidation by the enzyme (Section 3.3). However, after analysis and pondering, none of these explanations are kinetically adequate or consistent, thus leading us back to the first explanation, namely, the paucity of the ferrous heme intermediate R in the catalytic cycle and the proposal of an unorthodox ferric-O2 intermediate (Section 4.3.4) similar to that postulated in the NOD mechanism (Section 3.3.4) [18,74,75]. Supporting this supposition, the R intermediate is only significantly observed within intact mitochondria under severe hypoxia [101] and even more so with uncouplers and impaired membrane potentials [102]. A comparative analysis of the enzyme kinetics and structures of the two O2-metabolizing heme enzymes, namely NOD and CcO, further indicates the necessity and likely operation of similar electron-transfer control motifs. For the O2 reductases, a heme Fe3+O2Cu2+ spintronic electron-transfer control also provides, in my judgement, an aesthetically satisfying mechanism for chemically, energetically, and structurally linking the countercurrent movements of electrons and protons.

2. Purpose and Scope

The NO sensitivity-resistance paradox provides a clear window for viewing the structure–function of NOD and CcO and opens up a unique perspective not provided by isolated kinetic measurements, spectroscopic intermediate studies, and static structures. A thorough and critical review of the investigations of NO binding and of the strategies that these two enzymes are thought to employ to resist a far more detrimental irreversible NO inhibition, as predicted from NO avidities, leads to the necessity of considering alternative mechanisms and introducing adequate concepts. Data and theory supporting an unusual, but apparently common, O2-ferric heme interaction and enzyme intermediate that controls heme spin state, protonation, and electron transfer and that averts ferrous heme-NO formation are discussed and related to structural models, allostery, and motions. The concept and derived model provide a unique perspective for understanding proton pumping by CcO as well.
“No paradox, no progress”—Niels Bohr (1885–1962).
“For an understanding of the phenomena, the first condition is the introduction of adequate concepts; only with the help of the correct concepts can we really know what has been observed.”—Werner Heisenberg (1901–1976).

3. NOD Resistance to NO Poisoning

The NOD function, distribution, reaction, and mechanism of various members of the globin superfamily and their associated reductases have been widely reviewed, and the reader is referred to specific sources for additional information [5,18,55,56]. In my analysis, I consider the well-studied flavoHb-type NOD as an archetype for all NODs. While the function of globins as NODs has received considerable attention, the topic of NO resistance to irreversible inhibition has surprisingly received considerably little attention despite the glaring paradox it presents. As already stated, globins with associated reductases should not function as enzymatic NODs as initially envisioned [4,49] and more reservedly conceived [53,54]. A viable mechanism to explain the widely observed NOD activities and reversible NO inhibition has only recently emerged.
FlavoHbs catalyze NO dioxygenation with exceptional efficiency. At the core of the reaction, an enzyme-bound superoxide radical couples with the NO radical to form nitrate (Equation (1)). The bimolecular rate constant of the reaction estimated from kcat/Km (NO) is 2.4 × 109 M−1 s−1, which approaches the diffusion limit and is 50–100 fold larger than the rate constant measured for HbO2 and MbO2 [19,20]. Globin-NODs serve the additional role of enzymatically catalyzing O-atom rearrangement in nitrate formation [17]. While tetrameric red blood cell HbFe2+O2, or rather HbFe3+O2, catalytically dioxygenates NO, it acts stoichiometrically and not enzymatically. Herein lies the root of the paradox. An obvious difference is that the flavoHb-type NOD enzyme has a structurally-linked reductase domain, whereas the erythrocyte metHb reductase is not coupled to the oxidation event, at least not in the tetramer. Furthermore, in order to function in O2 transport-storage, the red blood cell Hb and muscle Mb are reduced to the ferrous state regardless of the presence of O2. The interactions of various globins with specific reductase partners are likely choreographed for the structural and functional motions required for NOD activity [18].

3.1. O2 vs. NO Binding to Globins

As a rule, NO shows manyfold greater affinity for ferrous heme in globins than O2 [4,89,103,104,105]. The E. coli flavoHb-type NOD is typical in this regard [4,51,62,75,106]. The bimolecular association rate constant measured for O2, k’, is 3.3 × 107 M−1 s−1, and the first-order dissociation constant, k, is 0.2 s−1 (Equation (9)). The calculated half-time for O2 dissociation is 3.5 s, whereas catalytic turnover occurs in roughly 7 ms at 20 °C and 1.5 ms at 37 °C. The calculated equilibrium dissociation constant, KD, is 6.1 nM. Thus, by this criterion, half-saturation of the enzyme is expected at nanomolar O2. But the Km(O2) value is 20 µM at 20 °C and 120 µM at 37 °C. NO associates with the ferrous heme with a roughly similar bimolecular rate constant of 3.1 × 107 M−1 s−1 (Equation (10)). However, NO binds ~1000-fold more tightly with a dissociation rate constant of 0.0002 s−1. The calculated KD for NO is 6.5 pM, and the half-time for NO release is 58 min at 20 °C, or roughly 500,000 times longer than turnover at 20 °C. Our initially proposed enzyme mechanism [49] ignored the well-known high-affinity binding of NO to ferrous heme. Transient ligand binding kinetic measurements obtained in collaboration with John Olson’s lab [51,62] demonstrated high-affinity NO binding and produced additional glaring inconsistencies in the proposed mechanism that have long demanded an explanation. More recently, a resolution to the contradictions, inconsistencies, and discordant data and interpretations (Section 3.2 and Section 3.3) has been achieved with evidence of an O2-sensing electron transfer (ET) switch and stepwise motions in the NOD mechanism (see Section 3.3.3) [18,74,75].
Fe2+ + O2 ⇌ Fe2+O2
k’ = 3.3 × 107 M−1 s−1, k = 0.2 s−1, t1/2 = 3.5 s, KD = 6.1 nM
Fe2+ + NO ⇌ Fe2+NO
k’ = 3.1 × 107 M−1 s−1, k = 0.0002 s−1, t1/2 = 58 min, KD = 6.5 pM

3.2. NO Inhibition of NODs

Steady-state kinetic investigations of the wild-type flavoHb and engineered mutants have been most revealing of the enzyme mechanism, in particular, where contradictions between the steady-state kinetics and elementary rate constant measurements have arisen. For example, NO reacts with the flavoHb ferrous heme with a rate constant of 2.6 × 107 M−1 s−1 limited only by the rate of NO migration through the protein matrix to the iron, but the reaction of NO during steady-state catalysis is nearly 100-fold faster with a rate constant of 2.4 × 109 M−1 s−1, as calculated from the kcat/Km(NO), thus suggesting an alternative pathway for NO migration. Indeed, substitution of the conserved G8 position valine in the putative gated tunnel for NO with a more bulky leucine residue decreased the kcat/Km(NO) value and presumably NO migration three-fold, while a smaller alanine increased the value ~two-fold [75]. Clearly, a consideration of motion in the protein structure during catalytic turnover is critical for a reliable interpretation. In this case, as in others, it was imperative to assign greater weight to the steady-state kinetic results than to the various potentially artificial, suboptimal, or preconceived states used for pre-steady-state transient kinetic measurements. For example, the measured Km value of 20 µM for O2 at 20 °C is 3300-fold greater than the KD for O2, and interaction or binding of O2 paradoxically appears rate-limiting. At 37 °C, the Km (O2) value is even larger at 120 µM [51,62,75]. As stated above, the kcat/Km(NO) value is 100-fold larger than the value measured for NO association with the ferrous heme, supporting an NO migration route that opens for NO association during turnover [75]. And, most importantly, NO inhibits competitively with respect to O2 and not irreversibly as suggested by the ferrous heme affinity measurements. NO inhibits the enzyme by 50% at an NO:O2 ratio of approximately 1:100 at 37 °C and 1:20 at 20 °C. The inhibition pattern observed in 1/v vs. 1/O2 plots [51,62,75] indicates competitive substrate inhibition in a ping-pong mechanism [107]. The observed cooperativity of O2 and NO in steady-state kinetics and the effects of mutations demanded a mechanism that was better informed by structural dynamics [75].

3.3. Reversal of NO Inhibition of NODs

Most, if not all, oxy-globins catalyze NO dioxygenation, and globins tightly coupled to specific reductase partners often show enzymatic NOD activity [51,57,62,75,108,109,110]. Globins that are not tightly coupled to a reductase, such as the tetrameric red blood cell Hb, are reduced and irreversibly inhibited for subsequent NO dioxygenation with NO binding to the ferrous heme and thus prevent further O2 binding [16,62]. On the other hand, globins that are tightly coupled to a reductase, such as the monomeric cytoglobin [110], are reversibly inhibited by NO in competition with O2 [16,51,75,108,111,112]. NO inhibition is not progressive as expected for the small dissociation constants for ferrous heme-NO. The NO:O2 ratio for 50% inhibition of the flavoHb-NODs and the flavoHb-like mammalian NOD activity, presumably a Hb monomer tightly coupled to the cytochrome b5 or P450 reductase [18], is remarkably similar at approximately 1:100 at 37 °C [51,62,75].
Discerning reversible NO inhibition and sleuthing mechanisms of inhibition reversal has been hampered by non-uniform assay methods and conditional kinetic values. For example, the extensively characterized wild-type E. coli flavoHb-NOD [75] shows respective turnover numbers of 580 s−1 and 160 s−1 at temperatures of 37 °C and 20 °C. The respective Km(O2) values are 120 µM and 20 µM. In contrast, Mills et al. [108] independently reported respective Vmax and Km(O2) values of ~10 s−1 (calculated) and 47 µM for the E. coli flavoHb-NOD at 37 °C. Using kinetic equations derived for the E. coli flavoHb-NOD [75], we can estimate that the NOD activity would be inhibited >95% by the excessive and non-physiological 36 µM NO supplied at the O2 saturation of 200 µM used in the assays. Rather low NOD activities of 4.5 s−1 and 9.4 s−1 have also been reported for Aspergillus oryzae flavoHbs at room temperature in a 50 mM sodium citrate buffer at pH 5.5 [113]. A Staphylococcus aureus flavoHb-NOD turnover number of 66 s−1 was reported [114]. Frey et al. [57] reported NOD turnover numbers for various overexpressed bacterial flavoHbs in the range of 78–128 s−1, including a value of 91 s−1 for the E. coli flavoHb at room temperature with a uniformly inhibitory condition of 10 µM NO at normal O2 saturation. The method of NO delivery may confound NOD assay results, given that NO is capable of inhibiting the NOD activity reversibly. Furthermore, under potentially uncoupled reducing conditions, NO may also inhibit NODs less reversibly or irreversibly. In future investigations, standardized methods and conditions for the NOD assay [115] should be used to ensure reliable kinetic comparisons and mechanistic interpretations.

3.3.1. NO Reduction by flavoHb-NOD

Numerous phenomena have suggested a NOR activity and function for flavoHbs. Such an activity (Equations (11)–(13)) would explain the NO resistance of the flavoHb-NOD. For full reversibility, the NOR turnover number would simply need to equal the NOD turnover number, and the NOD would never appear inhibited with respect to NO consumption. Moreover, in determinations of reaction stoichiometry, the NO-inhibited NOD would yield one N2O per two NO instead of two NO3. There would be no O2 dependence of the NOD activity. For a sub-NOD turnover rate, NO inhibition would appear progressive at all NOD velocities greater than the NOR turnover rate. Measured NOR turnover numbers of various flavoHbs range from 0.02 to ~1 NO s−1 [51,62,64,107,108] or from 0.2 to 1% of the maximal NOD turnover. Clearly, flavoHb-NOR activity provides little, if any, role in the NO resistance of the NOD activity or reversal of NO inhibition.
Fe2+ + NO → Fe2+NO
Fe2+NO + e → Fe2+ + NO k = < 1 s−1
2 NO + H+ → N2O + OH
In addition, the overexpressed flavoHb shows no measurable NOR activity and no protection of the NO-sensitive aconitase within anoxic E. coli [66]. Yet, observations by multiple investigators using different organisms have suggested an O2-independent function for the inducible flavoHbs. The interested reader should consult the numerous relevant publications [49,60,64,66,108,114,116,117,118,119,120,121,122,123]. There are many possible explanations, yet no common activity or function has emerged. I have already introduced a couple of explanatory scenarios above (Section 1.3). The possibility of a two e N2-yielding N2O reductase activity (Equations (14) and (15)) requires testing. Many organisms express dedicated NORs that generate N2O [3,70,84]. The abundant flavoHb, like copper and cobalamin enzymes, may bind and reduce N2O to N2. E. coli grown anaerobically with nitrate or nitrite reduces N2O to N2 [124], but the identity of the catalyst remains unknown. Univalent reduction of N2O generates a toxic hydroxyl radical (Equation (16)) [2] and a gas-binding flavoHb with a two-electron reducing capacity is, as previously pointed out by Poole and coworkers [58], an attractive candidate.
Fe3+ + N2O → Fe3+N2O
Fe3+N2O + 2 e + 2 H+ → Fe3+ + N2 + H2O
N2O + e + H+ → N2 + OH

3.3.2. Denitrosylase Mechanism of flavoHb

Denitrosylation or O2 nitrosylation has been suggested as an alternative catalytic reaction mechanism of Ascaris Hb and flavoHb for O2 and NO metabolism and detoxification [63,69]. The basic premise of the proposed mechanism is that heme-Fe2+NO or a nitrosylthiol of cysteine is the species that reacts with freely diffusing O2 to form nitrate (Equations (17) and (18)). In the denitrosylase mechanism, formation of the heme-NO (Equation (11)) is not viewed as an irreversible inhibitory step. Rather, the ferrous-heme reduces and activates NO for reaction with O2 (Equations (17) and (18)). The deficiencies of the proposed mechanism have previously been enumerated [5,18]. Chiefly, the rate constant for the reaction of heme-Fe2+NO or heme-Fe3+NO is grossly insufficient for both the proposed catalytic mechanism and for the reversal of NO inhibition of the NOD mechanism. NgbFe2+NO, MbFe2+NO, and HbFe2+NO show rate constants of <1 M−1 s−1 [125,126,127]. Nevertheless, interest and enthusiasm for the potential relevance of the reaction mechanism persists [104] with the demonstration of a larger rate constant of 6.3 × 103 M−1 s−1 for the reaction of structurally unrelated ferrous nitrosylated nitrobindin with O2 [128]. In terms of flavoHb-NOD activity, denitrosylation (Equation (18)) with the smaller rate constant of 1 M−1 s−1 falls short by over a million-fold for reversal of NO inhibition in the steady-state under the condition of 50% NOD inhibition, while the larger rate constant is insufficient by a factor of 4000-fold [18]. The rate constant required for the full reversal of inhibition (Equation (17)) is simply that measured for NO dioxygenation (Equation (2)) or 2.4 × 109 M−1 s−1.
Fe2+NO ⇌ Fe3+NO
Fe3+NO + O2 → Fe3+ + NO3 kox = ~1 M−1 s−1

3.3.3. NO and O2 Ligand Exchange Reaction or a Kinetic Aberration

In an attempt to resolve the NO inhibition paradox presented by the initially proposed NOD mechanism (Equations (1)–(6), (9), and (10)), Das and Meuwly suggested a possible mechanism for NO and O2 exchange (Equation (19)) at the ferrous heme of Mycobacterium tuberculosis HbN-NOD [129]. However, neither the measured rate constants nor the structures support a simple ligand exchange or tunnel selectivity for NO and O2 entry [130,131,132].
Fe2+NO + O2 → Fe2+O2 + NO
Following the first data-driven confrontation with NO resistance paradox [51], we considered the possibility that the large apparent Km(O2) value determined for the NOD activity relative to the unusual 4000-fold smaller KD(O2) value measured for the reduced flavoHb represented a competition of O2 with NO that prevented NO binding and presumably irreversible NO inhibition. In this case, the true k’NOD or kcat/Km(NO) value would have to greatly exceed the measured value of 2.4 × 109 M−1 s−1 [51,75], which is near the diffusion limit for the reaction. Following the early reaction scheme design and kinetic modelling with some favorable rate constant estimates (e.g., k’NOD of 0.6 × 109 M−1 s−1 for 20 °C), a smaller but nevertheless troubling >10-fold discrepancy between the measured apparent Km(O2) value and the predicted Km(O2) value remained [51]. Those early kinetic discrepancies were finally eliminated with a revised reaction scheme (Section 3.3.4) that incorporates an O2-sensing ferric heme trigger or switch for the ET step and O2 activation [75]. The new scheme also explains reversible NO inhibition.

3.3.4. Heme-Fe3+O2 Quantum State Switch in the flavoHb-NOD Catalytic Cycle

In the ensuing years of the unresolved paradox, significant advances in structure determinations, ET pathway modeling, and gas migration pathways had been achieved by investigators studying homologous globin-NODs, and those contributions were invaluable in formulating the structural model and resolving the discrepancies in the early model. Following our exhaustive kinetic and structural investigations of the E. coli flavoHb-NOD mechanism, a revised reaction scheme (Figure 1) that incorporates an O2-sensing ferric heme trigger or switch for the ET step and O2 activation was proposed [75], which, by all accounts, and the reticence, appears to have resolved the otherwise intractable paradox. In retrospect, intimate knowledge of the O2, O2, and NO-mediated inactivation and dynamic reactivation (i.e., inhibition) of [4Fe-4S]2+ (de)hydratases [133], where ferric iron is clearly involved, should have led to the model (Figure 1) from the start.
In the NOD catalytic cycle model, ET transfer to the ferric heme is activated by weak O2 and NO SOC [18,74]. Kinetic and structural evidence support controlled ET from the FADH2 to the ferric heme (Figure 1, step 6). The existence of a spin crossover (SC)-driven switch (Figure 1, step 5) is supported primarily by theory but also by the combined effects of O2 and key side-chain substitutions on the ferric heme spin state, steady-state kinetics, and ligand binding kinetics [74,75]. The stepwise ordered motions and kinetics of the ligands and flavoHb-NOD structure during the catalytic cycle have been previously described in detail elsewhere [18,75] and are briefly summarized in the legend to Figure 1
Another aspect of the model requiring emphasis here is the role of stored energy in the proposed ET switch mechanism. The ET switch requires protein motions and an estimated +2 kcal/mol of energy [74]. The NO dioxygenation reaction (step 7) produces NO3 and −30 kcal/mol of free energy that is utilized in the ET switch mechanism. The NO3 anion is produced in the hydrophobic distal pocket and migrates to the CD-loop and is captured in an anion hole of the unfurled spring-like CD loop (step 8). Energy is presumably stored through steric repulsion forces in the loop. To trigger the ET switch, O2 must displace the distal leucine (E11) residue away from its van der Waals contact with the ferric iron. For this motion, O2 is assisted by the structurally stored spring energy of the CD loop with loop re-furling and NO3 expulsion (step 2). The protein motions required for O2 to set the trigger for the ET switch constitute the slow step in the catalytic cycle. In Figure 2, NOD cycle clocks are shown for 20 °C and 37 °C, with the approximate timing of each step of the cycle denoted for the respective 6.7 ms and 1.5 ms turnover times. In the model, transient O2 binding to the ferric heme slowly occurs following the duration of step 2, and it is during this step that the primary gate and short tunnel for rapid NO entry also opens (step 3). Reversible NO inhibition occurs when NO competes with O2 for transient binding to the ferric heme (Equations (20) and (21)).
heme-Fe3+ + NO ⇌ heme-Fe3+NO
heme-Fe3+ + O2 ⇌ heme-Fe3+O2
The respective on and off rates measured for NO and the resting ferric flavoHb heme are 4.4 × 107 M−1 s−1 and ~4000 s−1 at 20 °C, and the estimated KD(NO) value is relatively large at ~100 µM [51]. The calculated half-time for NO dissociation is 170 µs, or a 20 million-fold shorter time than measured for the ferrous heme. However, from steady-state kinetic and structural studies [75], we have determined that NO accesses the heme via a gated short tunnel during turnover (step 3) with a much larger rate constant of 1.6 × 109 M−1 s−1 at 20 °C, thus revealing a smaller and more relevant KD(NO) value of 2.5 µM. From the kinetic competition of NO with O2 and a 50% inhibition at an NO:O2 ratio of ~1:20 at 20 °C, we approximate a KD(O2) value of 50 µM for catalysis, which is somewhat larger than the 20–27 µM Km(O2) value determined for 20 °C [51,75]. The estimated KD(NO) and half-time for NO dissociation can explain both the potency and rapid reversibility of NO inhibition. Interestingly, the measured rate constant for O2 entering the 20 Å long hydrophobic tunnel of the resting reduced flavoHb and binding the ferrous heme is 3.3 × 107 M−1 s−1 [51] or ~49-fold smaller than for NO entry during turnover. If we assume a similar rate constant for O2 entry and interaction (Figure 1, step 1) and a KD(O2) value of 50 µM, the estimated off rate for O2 is 1500 s−1 at 20 °C, and the half-time for O2 dissociation is roughly 0.5 ms or 7000-fold shorter than for the ferrous heme. Through this analysis, it becomes clear why the SOC of NO with the heme-Fe3+O2 complex (step 5) is essential for the ET switch mechanism. Dissociation of O2 from the heme following reduction would allow NO to bind and inhibit irreversibly. Observed cooperativity of NO and O2 in the steady-state has also suggested an interaction between NO and O2 that triggers the ET [75].
The proposed O2 binding to ferric heme is unusual but is supported by theory and experiment [134,135,136,137]. Bonding of O2 to the bare ferric heme is thought to occur through the overlap of the filled or half-filled O2 2π* orbital with the empty iron 3d2z orbital and through back-bonding of the half-filled 3dyz orbital with the half-filled O2 2π* orbital [135,137]. In flavoHb and other globins, the same orbital overlap, electrostatics of ferric iron, and hydrogen bonding from distal residues strengthen the weak O2 interaction [74]. Calculated KD(O2) values are consistent with the calculated bond energies and measured Km(O2) values [74].
In addition to weak electrostatic bonding, ferric heme and O2 interact magnetically through SOC. Similar to spintronic O2 reduction devices, SOC likely contributes to bonding through electron density currents, and ferromagnetic coupling reportedly favors currents more than antiferromagnetic coupling [76,77,78]. We can also find analogies of the weak electrostatic bonding in the SOC of O2 and ferric iron in advanced quantum dot theory. In electrostatic coupling between quantum dots, only the charge degrees of freedom of the electrons are involved. Within each quantum dot, the spin degree of freedom is then coupled to the charge degree of freedom via SOC. SOC is expected to be first order in the electrostatic interaction and second order in the spin [138]. Hence, electrostatics and SOC will interact in the bonding of O2 to the ferric iron, with the charge degrees of freedom, electron densities, and electrostatic forces dominating, with spins playing a secondary role by influencing electron currents and densities.
Energy differences in the ferric heme spin states (s = 1/2, 3/2, and 5/2) are relatively small but show preferences in ligand binding presumably due to electrostatics and SOC-induced charge currents. O2, NO, and water bind to low-spin states (s = 1/2 and 3/2) and anions with greater charge density preferably bind with ferric iron in the high-spin state (s = 5/2) [139]. Free heme stably binds O2 in the s = 3/2 spin state [134,137]. In the E. coli flavoHb, the energy differences between the ferric heme spin states are also small, with an estimated +2 kcal/mol difference between the s = 1/2 and s = 5/2 states, with smaller energy differences between the s = 1/2 and s = 3/2 states [74]. In the proposed O2 sensing ET switch activation mechanism (Figure 1, step 5 and Figure 3), O2, with its two half-filled parallel spin 2π* orbitals, binds and spin couples with the half-filled iron 3d orbital antiferromagnetically. SOC induces SC of the iron to the s = 3/2 state, achieving more favorable antiferromagnetic spin–spin and electrostatic interactions. SOC between metal atoms and O2 increases electron density currents to O2 as evidenced by low electrode overpotentials in O2 reduction reactions [76,77,78] and would thus increase O2 bonding. In the model, due to the collision of NO (s = 1/2) with the O2 molecule, NO loses angular momentum, gains spin (s = 3/2) [140], ferromagnetically couples with O2, induces SC of the iron to the high spin s = 5/2 state to achieve spin parity in the antiferromagnetic SOC of iron with O2 and NO, and the heme propionates and via the strong heme dipole, which transfers electron density to O2. Any loss of electron density from the propionates would decrease the carboxylate group pKa values and favor ET. The effect is somewhat analogous to the Bohr effect, where increased O2 binding correlates with the loss of Hb protons and protonation of Hb decreases O2 binding [141]. The change in iron spin also causes or correlates with the ruffling and saddling of heme in general [142] and Hb heme [143,144].
In the model, the energy from NO dioxygenation is stored in the flavoHb CD loop structure (Figure 1, steps 7 and 8), which energizes the protein and the ET switch mechanism (step 5) in the ready position for the high-spin (s = 5/2) ferric heme (Figure 3) triggered motion. NO may not stay coupled with O2 during the duration of the ET switch mechanism and ET. Regardless, a reaction of internalized NO with the Fe3+O2 complex (step 7) is evidenced by a steady-state k’’NOD value of 0.9 to 1.9 × 1010 M−1 s−1 that exceeds the calculated diffusion-limit for free NO and flavoHb-Fe3+O2 [75]. Chung et al. have discussed the important role of SOC in the NOD-catalyzed reaction (step 7) and other O2 activation reactions [145]. A similar ultrafast reaction rate constant of NO with free O2 of 1.6 × 1010 M−1 s−1 has been measured in what appears to be a diffusion-restricted caged reaction, although Koppenol has argued that the constant represents a free diffusion reaction [146].
A survey of the flavoHb-NOD and other globin-NOD structures has revealed a common motif in the mechanism of ET switching [18]. An ET pathway via the heme propionate d is influenced and ‘switched’ by the charge status and electrostatic field at the carboxylate group. In the model, SC from low-spin to high-spin ferric heme (step 5) with saddling and vice versa with flattening is thought to instigate the motions in the protein that affect the charge status and electrostatic field that turn the ET switch ON and OFF [74].
Comparisons of the E. coli and Saccharomyces cerevisiae flavoHb X-ray crystallographic structures [147,148] have provided insight into the ET path [149,150] and ET switch mechanism. In the proposed ferric flavoHb-NOD switching mechanism [18,74], the S. cerevisiae flavoHb structure models the ON state and the E. coli flavoHb structure models the OFF state (Figure 4). In the OFF state conformation, the water molecule O-atom p orbitals create an electrostatically repulsive Coulomb blockade for ET between the 5.6 Å distant FAD C8 methyl group C atom and the heme d propionate carboxylate O-atom. In the ON state conformation, the blockade water molecule moves, and the positively charged lysine (F7) ammonium group binds and neutralizes the negatively charged heme d propionate carboxylate group for ET. In addition, the flavoHb nine amino acid carboxy-terminus forms a short 310 helix with the positive end of its electrostatic dipole oriented toward the neutralized carboxylate group and favoring the ET [18]. Motion of the switch to the ON state (Figure 1, step 5) is triggered by saddling of the high-spin heme and motion of the iron atom toward the proximal histidine (F8) that conveys rotational and translational forces on the F-helix. Motion is mostly driven by the energy and forces stored in the O2-activated conformation (step 2) and triggered by the NO-induced SC (step 5). The switch returns to the OFF state following ET (step 6) and NO dioxygenation (steps 7 and 8). Kinetic and mutagenesis studies support the structural model [74,75].
Figure 4. Proposed ET switching mechanism. Key structural elements of the switch mechanism are shown for the E. coli flavoHb (PDB ID: 1GVH) (Left) and S. cerevisiae flavoHb (PDB ID: 4G1V) (Right). Reacting O2 and NO molecules drawn into the structural model (Right) are shown at 40% of their van der Waals radii. Water molecule shown in aqua and all other colors follow the CPK scheme. Iron and water O-atoms shown at 50% and ~100% of their van der Waals radii, respectively. Pathway for ET shown by the green arrow. Carboxy-terminal 310-helix dipole shown as a magenta arrow pointed from the negative to the positive pole of the dipole (Right).
Figure 4. Proposed ET switching mechanism. Key structural elements of the switch mechanism are shown for the E. coli flavoHb (PDB ID: 1GVH) (Left) and S. cerevisiae flavoHb (PDB ID: 4G1V) (Right). Reacting O2 and NO molecules drawn into the structural model (Right) are shown at 40% of their van der Waals radii. Water molecule shown in aqua and all other colors follow the CPK scheme. Iron and water O-atoms shown at 50% and ~100% of their van der Waals radii, respectively. Pathway for ET shown by the green arrow. Carboxy-terminal 310-helix dipole shown as a magenta arrow pointed from the negative to the positive pole of the dipole (Right).
Biophysica 05 00041 g004
Analogies to the proposed switching mechanism can be found in the low-to-high spin state change in cytochrome P450 BM3 with the binding of the fatty acid substrate laurate [151,152]. An effect on heme redox potential has been suggested to be important (for ET) with a 100 mV more oxidizing high-spin state heme [153]. Also, with laurate binding and SC, water moves to a nearby position for a potential proton donor role in O2 bond scission [154].
A direct analogy and scientific precedent to the globin-NOD structural changes with O2 ligand binding and heme spin state changes is also found in the classic studies of the red blood cell Hb cooperativity and allostery [42,143,144,155,156]. However, in the well-studied Hb, the more familiar ferrous heme undergoes low-spin to high-spin crossovers that affect O2 binding and the ‘relaxed’ and ‘tense’ globin structures. Whether O2 affects structure or structure affects O2 binding has been debated, but it is clear that both are thermodynamically related. In the case of the globin-NODs, energy stored in the CD loop spring-like structure appears critical for the O2- and NO-triggered structural changes [18,74,75]. On close inspection, the core of NOD enzyme structural motions appears precursory to the evolution of the O2-transporting red blood cell Hb cooperativity and allostery [75,155,156]. Indeed, the long-storied history and dedicated efforts in understanding the O2 cooperativity, motions, and allostery of Hb [42] have facilitated investigations and understanding of molecular, atomic, and electron motions in the globin-NOD mechanism [18].

4. CcO Resistance to NO Poisoning

Heme-copper cytochromes aa3, bo, and ba3 are all sensitive to reversible NO inhibition, while the cytochrome bd-type oxidases show far less sensitivity [96,97,98,99,157]. Sensitivity to NO inhibition has been attributed to the slower release of NO from the binuclear heme-copper type oxidases compared to the cytochrome bd-type oxidases [157]. The cytochrome bd-type oxidases utilize heme iron for O2 or NO binding, lack a distal metal, but bear glutamate residues both distal and proximal to the heme iron that apparently increase the dissociation of NO. The rapidly reversible NO inhibition of CcO and other heme-copper oxidases prompted my foray into the literature and proposal of a mechanism for resistance similar to that described for the flavoHb-NOD (Section 3). To my knowledge, the evidence for a role of the oxidized CcO in binding O2, NO, and CO during catalysis, like the flavoHb-NOD, has been only cursorily considered and discussed before [74].

4.1. O2 vs. NO Binding to CcO

Long after Keilin and Hartree’s seminal studies of reduced cytochrome oxidase and O2 [158], Greenwood and Gibson put forth a mechanism in 1967 for O2 binding BNCred [159] and later amended the mechanism to allow potential O2 binding to Cu+B in competition with CO [90]. O2 binds to BNCred rapidly with a second-order rate constant of 0.4 to 3.5 × 108 M−1 s−1 (Equation (22)) [90,159,160,161]. NO also binds to BNCred with a roughly equivalent rate constant to that for O2 at 0.4 to 1 × 108 M−1 s−1 (Equation (23)) [90,161]. An unusually large dissociation rate constant for O2 was approximated at 5 × 104 s−1 when the rapid reduction of O2 by BNCred was accounted for [90]. The NO dissociation constant is much smaller, with reported values of 0.004 s−1 at 20 °C and 0.012 s−1 at 37 °C [93]. The calculated equilibrium dissociation constant (KD) values for O2 and NO at 20 °C are roughly 500 µM and 40 pM, respectively. These early investigations by Gibson, Greenwood, and others revealed that following O2 binding to BNCred, O2 is reduced within roughly a millisecond [90,159,160,161]. In contrast, following NO binding to BNCred, NO dissociates very slowly with a half-time of ~1.5 min. The early data and proposed CcO enzyme mechanism [159] provided a clear, albeit potentially misleading (see Section 4.3), mechanism for envisioning irreversible poisoning of CcO and respiration by NO.
In retrospect, the large KD estimated for O2, as well as the early observation of Antonini et al. of ‘pulsed’ and ‘resting’ enzyme state effects on electron-transfer rates to cytochrome a3 and O2 [162], also revealed important deficiencies in the simple model for O2 binding and reduction. From these early investigations, it was apparent that an activation step for ET was required and that the O2 affinity, as measured and calculated, was much lower than expected for ferrous heme binding. Standing alone, these two early observations demanded an explication (see Section 4.3.5) that is not provided by current models of the CcO mechanism [163,164]. Granted the enormity and complexity of CcO investigations, some observations will remain unexplained. However, these two early observations, activated ET and low O2 affinity, appear fundamental and still require clarification.
Fe2+Cu+ + O2 ⇌ Fe2+O2Cu+ ⇢ ⇢ O2 reduced + Fe3+ Cu2+
k’ = 1 × 108 M−1 s−1, k = 5 × 104 s−1
Fe2+Cu+ + NO ⇌ Fe2+NOCu+
k’ = 0.4–1.0 × 108 M−1 s−1
k = 0.0025 s−1 @ 20 °C and 0.012 s−1 @ 37 °C; t1/2 = 58 s and 277 s

4.2. NO Inhibition of CcO

In 1994, Cleeter et al. [30] demonstrated potent NO inhibition of CcO in quick freeze-thawed mitochondria using the NO-generator S-nitrosoglutathione and suggested reversible binding to the BNC copper to account for the previously reported negligible effect of NO-releasing macrophages on CcO activity following cell harvests [10,165]. Brown and Cooper demonstrated potent inhibition of respiration and CcO by authentic NO within cells and, more surprisingly, showed that NO inhibition is indeed competitive with respect to O2 and is rapidly reversible [29]. NO was observed to inhibit CcO by 50% at NO:O2 ratios as low as 1:500 in neuronal cells [29], but ratios as large as 1:40 have been reported with more oxidizing redox states of mitochondria and CcO [95]. With observations of higher NO:O2 ratios and weaker inhibition, an important additional consideration is NO removal by autooxidation and non-enzymatic or enzymatic NO detoxifying pathways [16,88]. Remarkably, in order for NO and O2 to appear competitive, NO must be released from binding with a rate constant equal to or exceeding CcO turnover or on the order of 100 s−1. Otherwise, NO would rapidly and irreversibly inhibit the CcO as indicated by the on and off rates for NO binding to reduced CcO (Section 4.1). In some reports, NO inhibition is more slowly reversed (t1/2 = ~15 s), suggesting multiple mechanisms for inhibition and/or multiple mechanism(s) for release from NO inhibition (Section 4.3) [88,92]. Inhibition reversal mechanisms involving either oxidative or reductive NO metabolism have the potential to also provide a mechanism for significant cellular or mitochondrial NO consumption via the relatively abundant CcO [80,166]. Bona fide NO metabolizing enzymes, such as the NODs, also require careful attention in the assessment of CcO-catalyzed NO metabolism and NO inhibition reversal kinetics within cells or mitochondria [16,87].

4.3. Reversal of NO Inhibition of CcO

In 1980, Gary Brudvig and Sunney Chan first reported reductive and oxidative reactions of CcO with NO, generating either N2O or nitrite and involving the BNC [86]. Following Brown and Cooper’s seminal report on the NO sensitivity of CcO, the competition of NO and O2, and rapid reversibility of NO inhibition [29], several independent groups focused their investigations on the reductive (Section 4.3.1) and oxidative (Section 4.3.2) mechanisms for NO metabolism by CcO that may account for the unexpected reversal or resistance to NO inhibition and that may also contribute to significant NO metabolism by mitochondria, cells, and tissues [80,81,85,91,92,93]. In their culminating investigation, Cooper and coworkers proposed a mixed inhibition model to explain reversibility of inhibition in which NO can either bind reduced heme and dissociate (competitive) or bind cupric ion to be oxidized (non-competitive) [94]. Indeed, Torres et al. reported NO-induced CcO Soret band increases at 441 nm during turnover and assigned the increase to a3-Fe2+NO complex formation with a Soret maximum of 444 nm [91]. Similar spectra and interpretations were reported by Sarti et al. [93]. In retrospect, it is noteworthy that the Soret maxima for the flavoHb heme-Fe2+NO and Fe3+NO complexes are found at nearly indistinguishable wavelengths of 420 nm and 419 nm [51,62], thus suggesting an ambiguity. The possibility of NO competing with O2 for binding a CcO cycle intermediate with an oxidized Fe3+Cu2+ center (Section 4.3.3) does not appear to have been considered [93,94,167,168,169,170]. Since the early investigations of reactions of NO with BNCox [86,171], subsequent investigations appear to have focused solely on the oxidative Cu2+ reaction with NO [81,166] and ignored the likelihood and significance of NO binding or transiently interacting with a3-Fe3+ (Section 4.3.2).

4.3.1. NO Reduction by CcO

In 1996, Borutaité et al. suggested a role for CcO in mitochondrial NO reduction [80] that was consistent with the in vitro observations of Brudvig and Chan [86]. Brudvig and Chan had reported N2O formation following prolonged incubation of CcO with 1 mM NO, 10 mM ascorbate, and the ET mediator p-phenylenediamine and suggested an enzymatic NOR mechanism (Equation (24)). Furthermore, the recognition of the evolution of CcO from the cytochrome bc-type NOR [3,40,41] had suggested an ancestral and residual NOR function of CcO. However, Sarti’s group subsequently reported that purified CcO has no detectable NOR activity [85], and a similar conclusion was recently reported by Balaban’s group for cardiac muscle mitochondria [87]. Clearly, the low or negligible NO reduction activity of CcO would be inadequate to rapidly reverse and prevent progressive NO inhibition of the O2 reduction activity in the steady-state. Other heme-copper oxidases including the cytochrome ba3 and cytochrome caa3 of Thermus thermophilus do catalyze the reduction of NO [172]. The Pseudomonas stutzeri cytochrome bb3 oxidase with significant homology to the cytochrome bc-type NOR also shows a notable NOR activity of approximately 2 s−1 [100].
Fe2+NOCu+NO + 2 e + 2 H+ → Fe2+Cu+ + N2O + H2O
The possible NOR function of CcO and other O2 reductases has also been important for delineating pathways for NO metabolism in cells and mitochondria especially under the condition of hypoxia or anoxia [16,80,83,87], where NOD activities or MbO2-catalyzed NO dioxygenation fail [16,87,88,173] and knowledge of NORs is lacking. The reported failure of CcO to reduce NO [85,87] contradicts the findings and interpretations of Borutaité et al. [80] and suggests the existence of other mechanisms.

4.3.2. NO Oxidation by CcO

The results of numerous investigations by two independent groups and schools of thought over a period of more than two decades have supported a CcO-catalyzed oxidation of NO to nitrite as an important mechanism for NO inhibition reversal [81,88,166,167,169,174]. In the mechanism, NO is proposed to react with cupric ion in the oxidized intermediate ‘O’ with a rate constant of 104 to 2 × 105 M−1 s−1 and to rapidly form a cuprous-nitrite complex (Equation (25)) [88,175]. Reduction of BNCox rapidly ejects the bound nitrite (Equation (26)) [81,166,174] and produces the reduced CcO for O2 binding in the catalytic cycle. The oxidation reaction is relatively slow at physiological NO concentrations, temporarily removes a fraction of the CcO from the O2 reduction cycle, and produces non-competitive inhibition in steady-state kinetic analysis primarily at higher [NO] [94]. In the proposed model, it is the fully reduced CcO intermediate R that is thought to be susceptible to competitive inhibition by NO (Equations (22) and (23)) [88,94].
Recall, however, that the calculated half-time for NO dissociation from the reduced heme is 58–173 s. The inhibition by competitive binding of NO to BNCred would be progressive and essentially irreversible in the time frame of the CcO kinetic assays barring the progressive non-competitive inhibition expected with NO oxidation [94]. There stands a glaring contradiction in the current models suggesting reversible competitive NO inhibition of reduced intermediate R with relatively small rate constants for NO dissociation [88,94,95,169]. It is noteworthy that Antunes et al. applied a 10-fold larger NO dissociation rate constant value of 0.13 s−1 and a 2.5-fold smaller NO association rate constant for an overall 25-fold smaller kNOon/kNOoff ratio in their kinetic modeling of simple reversible competitive inhibition of the reduced CcO [95]. The smaller ratio undoubtedly produced a more modest concave curvature and less progressive inhibition in the plots and simulations of -d[O2]/dt.
Fe3+H2OCu2+ + NO ⇌ Fe3+Cu+NO+ + H2O → Fe3+Cu+NO2 + H+
k’ = 104–2 × 105 M−1 s−1
Fe3+Cu+NO2 + e → Fe2+Cu+ + NO2
While it has been argued that kinetic experiments showing large >100 s delays in the full reversal of inhibition of the isolated CcO following rapid NO depletion with HbO2 support the formation of Fe2+NOCu+ [93], NO inhibition of respiration and CcO within mammalian cells is immediate, is not progressive, correlates contemporaneously with NO concentrations, and is reversed fairly rapidly [16,29]. In this case, CcO may be forced into a reduced state by unfavorable catalytic conditions (e.g., low O2 or dysregulated ET) to produce a poorly reversible NO inhibition that appears as a rapid and progressive inhibition under sub-optimal conditions [92], produces bona fide Fe2+NOCu+ spectra [92], and shows small NO dissociation rate constants [93]. Similar effects of dysregulated ET are observed in assays of globin-NODs with small NO dissociation constants with heterologous reductases [111] or a non-functional dysregulated ET switch with glutamine substituted for the F7 lysine residue (Figure 4) [75]. A greater understanding of the controls on ET in CcO and the factors influencing switching are required (see Section 4.3.3). Again, recall that the R intermediate was reportedly not detected in actively respiring mitochondria but only seen in extremely hypoxic non-respiring mitochondria [101]. The known bimolecular association rate constants may be insufficient for a reaction with a low-abundance intermediate and low ligand concentrations within the time frame of turnover. For k’(NO) = 1 × 108 M−1 s−1 and at 10 × 10–3 s per cycle, the [R] would need to be ≥ 2 µM for 50% inhibition.
It should be noted that Pearce et al. [176] and Collman et al. [177] have argued for an entirely different NO inhibition reversal mechanism involving O2 binding and reduction at the reduced Cu+, followed by the reaction of the O2 formed with Fe2+NO to form peroxynitrite. Moreover, it was further suggested that CcO catalyzes the reduction of peroxynitrite to nitrite [176,178]. Furthermore, the authors suggested that CcO significantly metabolizes NO via peroxynitrite formation and reduction within mitochondria and cells [176,179]. The extensive investigations of the reactions of NO with CcO to form nitrite have established a mechanism (Equations (25) and (26)) fully supported by enzyme kinetics [94] that does not support any additional roles for peroxynitrite or O2.
Finally, it is important to appreciate that only trivial amounts of NO are removed from CcO, mitochondria, and cells by the nitrite-generating NO inhibition reversal mechanism. CcO shows little, if any, role in the metabolism of NO by mammalian cells [16] or E. coli [48]. Recall that CcO was initially considered potentially important for NO metabolism by cells and mitochondria [16,80,83,166]. In fact, it was the supposition that initiated and supported my initial work on NO metabolism [16]. It is now well-appreciated that dedicated NODs and NORs metabolize NO efficiently and, at least partially, protect the NO-sensitive CcO [16,87] and other O2 reductases [82] from both a rapidly and slowly reversible NO inhibition.

4.3.3. Heme-Fe(III)O2 Quantum State Switch and the CcO Catalytic Cycle

The similarities in the NO resistance of the NOD and CcO functions have suggested a common enzyme strategy involving O2, NO, or CO binding or interaction with ferric heme. With want of extensive knowledge of the complex structures and mechanisms of CcO and other terminal oxidases, I hesitantly, but with further study, ardently set my sights on developing an internally consistent model for the CcO catalytic cycle. The resulting model introduces an ET control mechanism and builds upon the principles encountered or developed in investigations of the globin-type NOD structure–function, electron transfer, kinetics, and O2 chemistry (Section 3). Concurrently, in order to better explain the proton pumping mechanism first demonstrated by Mårten Wikstrom in 1977 [180], I describe a potential proton tunnel from the matrix, and I also introduce an O2 and reduced oxygen intermediate-driven countercurrent proton-coupled electron transfer (PCET) pathway. A de novo model-building approach that I, like others [163,164,168,181,182,183,184,185], have taken necessitates the subsequent pursuit of the following critical questions.
(i)
Can the model explain all empirical observations?
(ii)
Does the model agree with existing theory, or overlap with other models?
(iii)
Does the model suggest definitive experiments?
(iv)
Does the model need to be revised or rejected?
The rapid reversibility of the NO inhibition of CcO and competition with O2 suggested a CcO catalytic cycle and mechanism in which O2 binds weakly and interacts transiently with BNCox to initiate turnover (see Figure 5 and discussion below). Indeed, in 2011, Wikström and coworkers used DFT calculations to conclude that O2 binding to BNCox is “nearly isoergonic with dioxygen binding to a reduced BNC” [186]. Apparently, the authors did not grasp the implication of univalent O2 reduction (Fig. 1A in Kaila et al.) for CcO catalysis (see Figure 5) or rejected the possibility due to (i) thermodynamic considerations, (ii) a failure to observe UV–visible spectral changes (Fig. 6 in Kaila et al.), (iii) an unawareness of the role of SOC in redox reactions and the burgeoning field of spintronics, and (iv) unforeseen mechanisms for the membrane potential to affect proton tunneling, OW (or OH) dehydration, and ET (see below). In this model, NO competes with O2 for the oxidized center, but only O2 with two parallel spin (s = 1/2) 2π* electrons is able to couple with ferric heme iron (s = 1/2) and cupric ion (s = 1/2) in the center (OD) to trigger SC in the low spin a3 heme (see Figure 6 and discussion below) and to facilitate electron transfer to the ferric heme via arginine 438 and the d propionate of the ferric a3 heme (see Figure 7 and discussion below). In the full model, the ET switch mechanism constitutes the energetic proton pump in an orchestrated system with an electrostatically-driven proton influx tunnel and an exiting PCET pathway of the concerted EPT type [187], with bifurcating unidirectional proton and electron current controls (see Section 4.3.4). The beginnings of the model can be found in an earlier pondering over the evidence for O2 binding to BNCox [74].
CcO Catalytic Cycle
The proposed CcO catalytic cycle (Figure 5) is simply envisioned as four univalent O2 reduction steps, each coupled with a pumped proton and with each of those steps followed by a protonation of the reduced oxygen intermediate produced in the prior step. The reduction of O2, or a reduced oxygen intermediate, releases the free energy that is required to pump the proton against the pH gradient. The proposed cycle links each O2 or oxygen intermediate reducing electron with each proton in time and space, potentially resolving the apparent timing conundrum [188,189].
Figure 5. Proposed CcO catalytic cycle. Competitive and non-competitive inhibition pathway for NO originating at BNCox (top) with O2 binding BNCox followed by cycle intermediates. The cycle model is formulated with alternating successive exergonic reductive-proton pumping steps (red brackets) and proton-consuming steps (black). Standard free energy changes shown for each reduction step are from electrolytic O2 reduction measurements [190] or are calculated (*, see text). Iron spin states shown are, in some cases, unknown or hypothetical. In the model, iron and copper ions are shown in their oxidized states with the assumption of SOCs and electron density distributions between the metal ions and the reduced oxygen intermediates. The nomenclature of intermediates is regularized for a simplified model description where O is oxidized, R is reduced, S is superoxide, P is peroxide, and F is ferryl. The subscript A designates the anionic form, and the subscript H designates the protonated state. D stands for dioxygen, and T represents tyrosyl. Corresponding common designations for cycle intermediates [163] are given in parentheses: OW (O hydrated), SA (A), FA (PR), FH (F), OT (PM), OA (OH), and R, (R).
Figure 5. Proposed CcO catalytic cycle. Competitive and non-competitive inhibition pathway for NO originating at BNCox (top) with O2 binding BNCox followed by cycle intermediates. The cycle model is formulated with alternating successive exergonic reductive-proton pumping steps (red brackets) and proton-consuming steps (black). Standard free energy changes shown for each reduction step are from electrolytic O2 reduction measurements [190] or are calculated (*, see text). Iron spin states shown are, in some cases, unknown or hypothetical. In the model, iron and copper ions are shown in their oxidized states with the assumption of SOCs and electron density distributions between the metal ions and the reduced oxygen intermediates. The nomenclature of intermediates is regularized for a simplified model description where O is oxidized, R is reduced, S is superoxide, P is peroxide, and F is ferryl. The subscript A designates the anionic form, and the subscript H designates the protonated state. D stands for dioxygen, and T represents tyrosyl. Corresponding common designations for cycle intermediates [163] are given in parentheses: OW (O hydrated), SA (A), FA (PR), FH (F), OT (PM), OA (OH), and R, (R).
Biophysica 05 00041 g005
In the proposed catalytic cycle (Figure 5), I have renamed some of the well-known cycle intermediates with subscript A representing an anion and subscript H representing the protonated form for the sake of logic and simplicity in presentation and discussion. The series of univalent reductions of intermediates in catalytic sequence is named for the oxidized center with water (OW), oxidized center with dioxygen (OD), oxidized center with superoxide anion (SA), oxidized center with peroxide anion (PA), ‘oxidized’ center with ferryl (s = 1) [191], hydroxide anion bound to Cu2+ (FA), and the oxidized center with hydroxide anion (OA). Included as intermediates of protonation are the corresponding SH, PH, FH, and OW. The existence of the tyrosyl radical intermediate [192] was accounted for and included as an oxidation product or side-reaction of ferryl (FH), named OT, and is shown in an equilibrium with FH. As an anion, OT is proposed to serve a critical function in proton pumping (see below). For the simplified nomenclature, proposed intermediates roughly correspond to empirically characterized intermediates with traditional nomenclature [163,164,168,184,193,194] as follows: OW corresponds to hydrated O, SA corresponds to A, FA corresponds to PR, and OA corresponds to OH. There is no ferrous intermediate (R) or cuprous intermediate (EH) in the normal cycle put forth. Importantly, the resting CcO with an oxidized center is EPR silent with antiferromagnetic SOC of the ferric iron (s = 1/2) and cupric ion (s = −1/2) [171] is proposed to transiently bind O2 and be rapidly reduced. Spin states for the remaining intermediates are hypothesized based upon ferric heme spin state preferences with various ligands [139] and have not been examined in, or correlated with, the literature. Energies assigned to the reduction-coupled proton pump steps are taken from the standard free energy changes measured with electrodes [190] except for O2, which is endergonic, and would be unfeasible for proton pumping at −180 mV or +4.2 kcal/mol. For intermediate OD and O2 reduction, a rough estimate of −6 kcal/mol is assigned to the step. In Blackmore and Gibson’s investigation of O2 binding and rapid reduction, the rate constant of 35,000 s−1 for reduced oxygen intermediate formation and the relationship ∆G = -RT ln k is used to estimate the actual free energy change during O2 reduction. The favorable thermodynamics are unexpectedly due to SOC and spintronics in a3-Fe3+O2Cu2+ (see below). Without correction for temperature and concentration effects; each step in the proposed cycle (Figure 5) appears to generate sufficient free energy to pump a proton against an [H+] gradient of 2 pH units or +3 kcal/mol where ∆G = −2.3 RT ∆pH. The large excess of energy is likely expended in conformational changes and heat generation. For an in-depth discussion of free energy for the proposed catalytic cycle with O2 binding to BNCred, the reader should consult Wikström and Verkhovsky [195], Wikström, Rich, and Gennis [163], and Blomberg [196]. The proposed cycle does not hypothesize or require the additional two-electron reduction of BNCox before O2 binding. Also, differences in proposed intermediates (e.g., PA and PH) and the sequence (e.g., OT or PM) will require resolution.
Competitive and non-competitive or slowly reversed NO competition are proposed to occur with NO initially competing for binding and interaction with the ferric iron of the OW (or a dehydrated O) intermediate. The association of O2 and NO and hence migration through the protein to the heme iron is equally rapid with similar rate constants of ~1 × 108 M−1 s−1 as measured for the ferrous cytochrome a3 (Equations (22) and (23)) [90,159]. A 50% inhibition and competition of NO with O2 at a 1:500 ratio [29] suggests a much larger dissociation rate constant for O2 than NO. If the turnover rate is 100 s−1 and the slow step is the trigger for OW reduction and requires the 10 ms, then for only 50% of the time or 5 ms would OW be available for O2 collision and interaction. The remaining 5 ms an NO molecule would be bound and interacting. If we assume a half-time for NO binding of 5 ms and an association rate constant of 1 × 108 M−1 s−1, we can calculate an NO dissociation rate constant of 139 s−1. The O2 dissociation rate constant would be roughly 500 times larger or 70,000 s−1 and half-time for binding of an O2 molecule would be ~20 µs. The association lasts roughly 2 billion-fold longer than a simple collision time of 10 fs. Relative to human experience, a brief second of contact would proportionally result in a 63 year association. At 1 µM O2, we can calculate that 100 molecules of O2 will collide in a second and interact ~20 µs each time resulting in a turnover rate of 100 s−1 or 10 ms per cycle. With a supraphysiological and toxic [NO] of 2 nM, 20 molecules of NO have the opportunity to compete with O2 by binding more tightly and longer to OW in that same 10 ms. From the analysis, we see that the large respective KD values for O2 and NO of 0.7 mM and 1.4 µM can be misleading. It is not necessary that the measured Km(O2) of ~1 µM equal the KD(O2) value, as is often supposed [38,197,198]. This relation is only true when substrate (i.e., O2) binding is the rate-limiting step in catalysis. On the other hand, the Ki(NO) value is expected to approximate the calculated KD(NO) value or k’(NO)/k(NO) of ~0.12 nM (Equations (25) and (26)), and it does [37,94] when ET in CcO is apparently dysregulated in vitro, as apparently also occurs in hypoxic uncoupled mitochondria [102]. The thought experiment above suggests an in vivo Ki(NO) value closer to 1.4 µM. Blackmore and Gibson calculated a remarkably similar large O2 off rate of 50,000 s−1 in their O2 binding and reduction studies [90]; however, that large off rate was assigned to an A (Fe2+O2Cu+) intermediate and not to the putative OD intermediate (see Section 4.3.5 below). In the model (Figure 5), non-competitive NO inhibition can be explained, at least in part, by the reaction of NO with the cupric copper to form bound nitrite, which is then released with reduction (Section 4.3.2, Equations (25) and (26)) [88,94,174]. However, reduction now forms the R intermediate, and the R intermediate is also subject to irreversible or very slowly reversible inhibition (Section 4.3 and Section 4.3.2, Equation (23)).
Spintronics in BNCox Binding of O2
If we ignore the contribution of rotational and translational motions of O2 and thus entropy [199], in the more restricted confines of the active site, we can calculate a relatively weak binding energy of −4.5 kcal/mol, where ΔG0 = RT ln K, R = 1.99 cal/mol K, K is in molar units, and the KD (O2) value is 7 × 10–4 M from the approximations above. Weak binding can be attributed to the limited electrostatic forces between Fe3+ and O2 and Cu2+ and O2 and the spin-orbit interactions. The estimated short half-time for bridging of Fe3+ and Cu2+ of roughly 20 µs would still suffice for the postulated ferric heme SC and doming motion. In the Mb, the heme iron SC and doming occur in 2.4 ps [200]. However, the available CcO structures, unlike the flavoHb-NOD structures (Section 3.3.4) [18,74], reveal little [201] or no evidence for doming or other movements near the cytochrome a3 propionates, begging the question of how O2-regulated ET switching control could even occur.
The magnetic SOC forces between O2 (s = 1) and BNCox Fe3+ (s = ½) and Cu2+ (s = −½) are expected to be negligible compared to the electrostatic forces creating the weak binding energy. Nevertheless, SOC or Nature’s version of spintronics may be critical for both O2 binding and the putative ET switch mechanism. The proposed role of SOC and hence spintronics in O2 binding BNCox and the ET switching mechanism is explained by Figure 6. In the model, O2 binds to BNCox with antiferromagnetic SOC to the ferric heme iron and ferromagnetic coupling to the cupric copper. Ferromagnetic SOC greatly facilitates electron density currents between the O2 and cupric ion and increases the oxidizing potential of triplet O2. The principle is exploited in the engineering of spintronic apparati where SOC lowers the overpotentials in the electrocatalytic reduction of chemiadsorbed O2 [76,77,78]. Moreover, electron conductivity was found to be greater with ferromagnetic SOC than antiferromagnetic SOC [77]. SOC is expected to increase electron density currents between the ferric iron 3d orbital and the O2 2π* orbital and increase the electrostatic interaction and incipient O2 binding. In BNCox SOC, O2 gains oxidizing potential by sharing electron density with the cupric ion, increases spin, causes ferric heme SC from s = ½ to s = 3/2 to achieve spin and magnetic parity, and withdraws electron density from the porphyrin chiefly residing at the a3 propionate d and a carboxylate groups. The effect would be similar to the Bohr effect in Hb where increased O2 binding correlates with a deprotonation [141]. In the proposed mechanism, protonation of both propionate carboxylate groups would diminish the electron density in the a3 carboxylate groups and decrease the large porphyrin dipole and O2 binding.
Figure 6. Proposed spintronics in BNCox binding O2. O2 collision and rapid transient SOC with the ferric heme and cupric ion (grey arrows). Low spin to high spin crossover (SC) induced in the ferric heme (thick green arrow) with antiferromagnetic (AFM) SOC to O2 and ferromagnetic (FM) SOC of O2 with the cupric ion. Electron density current migration from O2 to the cupric ion and from the ferric heme to the coupled O2 and Cu2+ (magenta arrow). Alternative antiferromagnetic SOC of cupric ion with O2 (yellow highlight) with a low energy spiral spin inversion (SI) of the cupric ion (double-headed green arrow).
Figure 6. Proposed spintronics in BNCox binding O2. O2 collision and rapid transient SOC with the ferric heme and cupric ion (grey arrows). Low spin to high spin crossover (SC) induced in the ferric heme (thick green arrow) with antiferromagnetic (AFM) SOC to O2 and ferromagnetic (FM) SOC of O2 with the cupric ion. Electron density current migration from O2 to the cupric ion and from the ferric heme to the coupled O2 and Cu2+ (magenta arrow). Alternative antiferromagnetic SOC of cupric ion with O2 (yellow highlight) with a low energy spiral spin inversion (SI) of the cupric ion (double-headed green arrow).
Biophysica 05 00041 g006
In an alternative scenario (Figure 6, yellow highlight), the unpaired cupric d electron undergoes a spin inversion for antiferromagnetic SOC to one of the unpaired O2 2π* electrons concomitant with ferric heme SC. The cupric ion is unusual in that it can undergo a spiral spin inversion with a very minimal energy input of +2.9 meV or +0.069 kcal/mol in cupric oxide with the rotation by an electromagnetic force [202,203]. However, in the alternative SOC arrangement, there is less spin parity and less electron density current between O2 and Cu2+. Importantly, there is a specificity for O2 conferred in the proposed SOC. The otherwise similar NO (s = ½ or 3/2) would show unique SOCs and would not be sufficiently oxidizing to withdraw electron density from the ferric a3 heme.
As illustrated (Figure 6), the bridged SOC at BNCox may also increase the O2 electron affinity by altering the stability of the triplet state. An analogous reaction has been proposed to occur when two NO radicals simultaneously couple with the O2 diradical to form iso-N2O4 bearing a single O-O bond, which rapidly homolyzes to form two NO2 molecules [204]. In the transient interaction, two NO molecules simultaneously transfer electron density to O2 through SOC. In contrast, the ferric heme and the cupric center have far less electron density to share or transfer. Thus, the bridged coupling is not expected to produce the long peroxide-like O-O bonds observed in oxidized CcO and other heme-copper oxidases with X-ray crystallography and discussed earlier [184,186,205] and more recently observed with cryo-EM [206]. In another example, O2 forms dimers at low temperatures and can magnetically couple with a total dimer spin of S = 0, 1, or 2 [207]. The stable arrangement of the parallel spin 2π* electrons of triplet O2 can also be ‘destabilized’ by the input of high-energy light to form the more oxidizing singlet O2 molecule via SCs, but the parallel spin configuration is also apparently weakened through low-energy interactions of O2 with specific molecules [208]. Simultaneous destabilizing or coupling interactions of the 2π* electrons apparently increase the electron affinity and reactivity of O2.
In the proposed ET switching mechanism, O2 SOC (Figure 6) withdraws electron density from the propionate d carboxylate group and removes the Coulomb blockade for ET from arginine 438 guanidinium nitrogen to propionate d carboxylate group of the ferric a3 (Figure 7). Withdrawing electron density from the propionate a carboxylate group facilitates the migration of the positively charged proton from the propionate a carboxylate O-atom to the propionate d carboxylate O-atom via the bridging water. The positively charged proton and the negatively charged electron migrate simultaneously to the propionate d carboxylate O-atom for ET. The vacated propionate a carboxylate simultaneously receives or ‘pumps’ a proton from the hydrogen-bonded water 1 (or hydronium) (Figure 7) and the matrix. Hence, the barrier to ET is lowered to form a bound reduced oxygen intermediate simultaneously with the migration of the proton (from the prior pump cycle) (Figure 5). The free energy for pumping a single proton from the water channel is derived from the reduction potential (ΔG = −nFΔE) for each of the OD, SH, PH and OT cycle intermediates (Figure 5). The migrating electron and the electron density of the negatively charged reduced oxygen intermediates (SA, PA, FA) and the hydroxide form (OT) donates electron density and momentarily increases the pKa of propionate a and d carboxylates. The negatively charged propionate d carboxylate would momentarily block further ET from arginine 438 and would be ready for proton transfer from propionate a via the bridging water molecule. Protonation of the SA, PA, and FA or OA cycle intermediates and formation of the respective OD, SH, PH, and OT cycle intermediates repeat the cycle. Note the proposed importance of OT intermediate and the tyrosyl radical for donating electron density to the ‘motor’ for pumping protons. Each of the protonated intermediates, like OD, is oxidizing and electron density withdrawing. Finally, PCET, shown operating between the cytochrome a and cytochrome a3 d propionates, facilitates ET and also forms the energetic ‘motor’ in the more complete proton pumping mechanism proposed for CcO (see Section 4.3.4). Recall, arginine rapidly tautomerizes [209], and tautomerization may facilitate concerted EPT. Concerted EPT and other types of PCET eliminate large ET barriers associated with charge [187]. Here, the charge transfers are countercurrent, rapid, electrostatically choreographed, and temporally complex and will require advanced computational analysis and models to resolve [187]. Cytochrome a serves as a harbor for an electron, and water molecules and the propionates serve as harbors for protons in the scheme. The concerted motion of electrons and protons is extremely fast [187].
Figure 7. Proposed ET switch mechanism and proton pumping. (Left) OFF switch position in cycle with no ligand, water, or electron density donating reduced intermediate anions (yellow box) bound to BNCox. (Right) ON switch position with transient electron density withdrawing cycle intermediates (green box) showing the structure of the ET pathway (black arrows), proton pathway (red arrows), and the heme dipole (magenta arrow). The water molecules bridging the propionate a and d groups (WB), the proton donor water (1). Water molecule O-atoms (aqua) are shown at ~100% of their van der Waals radii. CcO structure models 2ZXW and 3ASO were used for the respective OFF and ON ET switch positions.
Figure 7. Proposed ET switch mechanism and proton pumping. (Left) OFF switch position in cycle with no ligand, water, or electron density donating reduced intermediate anions (yellow box) bound to BNCox. (Right) ON switch position with transient electron density withdrawing cycle intermediates (green box) showing the structure of the ET pathway (black arrows), proton pathway (red arrows), and the heme dipole (magenta arrow). The water molecules bridging the propionate a and d groups (WB), the proton donor water (1). Water molecule O-atoms (aqua) are shown at ~100% of their van der Waals radii. CcO structure models 2ZXW and 3ASO were used for the respective OFF and ON ET switch positions.
Biophysica 05 00041 g007

4.3.4. Tunnel and Countercurrent Charge Migration Path in Proton Pumping

Water channels and chains of water molecules have been identified for proton migration in CcO and other O2 reductases and are required for both the reduction of O2 to water and the pumping of protons from the mitochondrial matrix to the intermembrane space [210,211,212]. Moreover, catalytic cycle-linked models for the mechanism of electrogenic proton pumping, incorporating key principles [185,189,213,214,215,216], have been previously advanced. One current model suggests that electric fields drive proton pumping [201]. Another recent model proposes a specific role for CuB in proton pumping [217].
The question of O2 sensing by BNCox and the resulting ET switch mechanism (Figure 7) led to further insight into proton pumping. The D, K, and H paths for proton delivery to the active-site for O2 reduction have been previously defined [184,201,217,218,219] but a separate path for proton pumping via the proton-donating water #1 (Figure 7) (aka proton loading site [185,220]) remains poorly defined, and a route via either the D, K, or H path appears tortuous. A putative pathway for proton delivery to the proton donor water 1 at the a3 propionate a carboxylate (Figure 7) is readily observed in the bovine (PDB ID: 2ZXW) (Figure 8 top), P. denitrificans (PDB ID: 7AU6), and E. coli (PDB ID:7CUB) heme-copper oxidase structures. The tunnel runs a nearly straight course perpendicular to the plane of the membrane, beginning at or near surface prolines (respective Pro14, Pro537/Pro538, or Pro546) and travels at an ~20° angle to the 28 amino acid alpha helix bearing the respective proximal His376 and His378 for the cytochrome a3 and a iron atoms. In the static bovine CcO crystal structure, a stricture in the tunnel of ~4.7 Å is measured near the surface Thr488 and Asp486 residues. The bound internal water O-atom and the restricting amino acid carbon atoms show a negligible (~0.1 Å) overlap of their van der Waals radii. Hypothetically, protons would migrate via a water chain with the force of the strong electrostatic field created by interaction of the proton with the 28 amino acid helix dipole moment of (cos 20°) 98 D or 92 D [221] and the alkaline state of the pump cycle (Figure 7, left). Assuming an average travel distance of ~10 Å from the helix axis and ignoring potential heterogeneous shielding effects of local static dipoles and applying the field equation E = 1/4πε0 (2p/r3), we can estimate that the positively charged proton would experience an appreciable axial electrostatic field of 2.8 × 107 V/cm, which, over a ~30 Å axial travel distance, would equal a substantial free energy change of +0.084 V or −1.94 kcal/mol. If we ignore the comparable counteracting force and energy of the membrane potential [222], the proton would reach a velocity upwards of 2.9 km/s, sufficient for a CcO turnover of >5 × 1011 s−1. However, copious proton delivery may also dysregulate ET control by allowing ET in the absence of the O2 trigger (Figure 6 and Figure 7). The putative water chain or proton wire is ~53 Å long and passes from Pro14 to Thr488 and along the proximal side of cytochrome a3, the farnesyl group, and by His376 (Figure 8, bottom). Remarkably, a small number of bound waters or charged side chains are found in the putative proton tunnel in the bovine, P. denitrificans, and E. coli structures. This should not be surprising since a uniformity of the water molecules, proton-exchanging hydrogen-bonds, or unimpeded single water molecule rotations [223] in the water chain would be required for a low energy barrier for rapid proton hopping as in the classic Grotthus mechanism. Consider also the possible role of the strong membrane potential in normally loading the polar tunnel with polar water molecules. For a proton hopping distance of 3.3 Å [223] and water or hydronium rotational diameter of ~1.4 Å, we can expect a chain of 12 water molecules including water number 1 and the surface water (WS) molecule (Figure 8). If confirmed, I suggest that the pathway be named the ‘Pro’ tunnel after the conserved surface Prolines and proton pump function. Together, the putative Pro tunnel and D, K, and H pathways may constitute the proton path(s) from the matrix for the four pumped and four consumed protons in the proposed cycle (Figure 5). The Pro tunnel may also serve in water removal and ‘dry out the surroundings of the BNCox‘ in OW (aka OH) (Figure 5) to form the dehydrated O intermediate [163], which would be expected to be more accessible to O2. An electrostatically-driven water pumping or efflux mechanism that depends on the series of water dipole rotations with proton hopping [223], the alpha-helix dipole, the local membrane potential, and any cycle-dependent opening and closing motions of the stricture formed by Asp486-Thr488 can now be modelled. The Pro tunnel, or water pump, may also replenish the water required to sustain fatty acid oxidation and other matrix-localized water-consuming metabolic pathways.
A low-barrier (i.e., short gap and electrostatically favorable) pathway for electron migration from cytochrome c2+ and a feasible (i.e., short distance proton donor–acceptor) pathway for proton migration to the aqueous phase of the mitochondrial intermembrane space with countercurrent PCET of the concerted EPT-type [187] is apparent. The identified electron and proton pathways are illustrated by the extended model in Figure 9. From the cytochrome a propionate d, PCET occurs in only two short segments of the pathway to or from cytochrome c2+, namely, with tautomerization at histidine 204 and between cytochrome a propionate d and the arginine 439 guanidinium group. Between histidine 204 and the arginine 439 guanidinium group, the pathways bifurcate within the electric field created by the hydrated magnesium ion. The putative pathways are separated in space by a short 4 Å distance that may allow weak electrostatic interactions between the migrating proton and electron pairs. Two water molecules carry the proton from the arginine 439 guanidinium group to histidine 204. The protons move with the electrostatic field dipole created by Mg2+, thereby preventing the reverse migration of protons or a pump leak. In addition, the third water (Figure 9) is slightly beyond hydrogen bonding distances in at least two CcO structures (PDB IDs: 2Y69 and 5IY5), suggesting an additional control against reverse migration of protons and for preventing a possible alternative path for PCET. The third water (3) is not visible in the structure modelled in Figure 9 (PDB ID: 2ZXW) but is nevertheless present in other structures. A short 10 Å long water or hydronium ion channel is identified for proton migration from histidine 204 to a surface water molecule (number 4). The putative exit channel is bounded by the peptidyl carbonyl O-atoms of tryptophan 104, asparagine 203, and phenylalanine 206. In the model, the electrostatically paired charge migrates from cytochrome c2+ via the tryptophan 104 indole, the CuA center, histidine 204 imidazole, to the arginine 439 guanidinium group. The water or hydronium ion (number 4) at the surface [224] may facilitate ET from cytochrome c2+ to tryptophan 104. Electrostatic field dipoles (Figure 9, magenta arrows) and charge pairing neutralization synchronize and control the direction and timing of the movements consistent with the early proposals of charge neutrality by Rich et al. [225]. Ultimately, the ET switch mechanism (Figure 7) and O2 binding and SOC (Figure 6) initiate the charge movements, and the reduction of OD, SH, PH, and OT intermediates provides the driving force (Figure 5). The simple structural model postulates a direct coupling [214] of proton and electron movements with switching and gating but no significant conformational changes.

4.3.5. Toward the Resolution of Unanswered Questions with Models

I had the good fortune and privilege of being one of about 100 scientists to observe the unveiling of long-awaited structures of the bovine CcO determined by Shinya Yoshikawa and coworkers [227,228] and the Paracoccus denitrificans CcO solved by Hartmut Michel and coworkers [229] at the Bioenergetics Gordon Research Conference in the Summer of 1995 at the Proctor Academy in New Hampshire. With the 13 transparencies representing the 13 subunits of the bovine CcO ceremoniously layered upon each other, thoughts of “What could be further accomplished? … the mechanism!” passed through my mind. Ten years later, I was asked to provide comments on a review article for a special issue of the Journal of Inorganic Biochemistry on heme entitled “Cytochrome c oxidase, ligands and electrons” by Professor Maurizio Brunori and coworkers. These and other events inspired a fascination with the complexity of CcO. Out of circumstance, my wandering with CcO began with measuring its NO sensitivity (resistance) within cells [16], followed by a 20 year interim dedicated to uncovering tightly-held secrets of the globin-NODs [18,74,75], to more recently studying the CcO structures and developing a model for the NO-resistant CcO mechanism with a beneficial working knowledge of the globin-NODs. I searched many other fields for adequate concepts and answers during this long period, including ET, porphyrins, structural dynamics, electronics, quantum chemistry, and eventually the spintronics field.
In their 2005 review, Brunori, Giuffrè, and Sarti [168] made a point by “pointing out where necessary unresolved facts or questionable interpretations” lie. I will begin by attempting to resolve, at least cursorily, these and other questions [195] as a post hoc appraisal of the usefulness of the mechanistic model (Figure 5, Figure 6, Figure 7, Figure 8 and Figure 9), a model built upon little known or unresolved facts, and a profound, but little recognized, paradox (discussed in Section 4.3). As succinctly and critically stated by Brunori et al. [168], “A pre-requisite for O2 and CO binding, but not for NO, is the complete reduction of this binuclear site”. Since the work of Greenwood and Gibson [161] more than 60 years ago, all CcO mechanistic models, including those with the catalytic cycle linked to proton pumping, have been formulated with that problematic assumption. By 1975, Britton Chance and coworkers [193] very briefly evaluated O2 binding to the oxidized binuclear site but apparently just as quickly abandoned the pursuit of the unorthodox idea as well as the limited but nonetheless intriguing data showing visible wavelength spectral changes at −97 °C. Roles for transient weak ligand (e.g., O2, CO, NO, N2, and H2O) interactions with oxidized ferric heme in the gas phase [137] or in enzymes [18,74] are now appreciated.
In their review, Brunori et al. raised four longstanding unanswered questions relevant to the model that I will address. Briefly, the key issues include (1) the low O2 affinity of CcO, (2) the strong competition of NO with O2 causing CcO inhibition despite the similar binding rate constants, (3) the intermediates and their formation, (4) the slow and fast rates of ET from the reduced cytochrome a to BNCox, and (5) whether ET or PT is rate-limiting. First, (1) the low affinities measured for O2 may indeed be explained by the rapid reduction of O2 and oxidation of BNCred during the measurements, as Brunori et al. surmised. The model presented predicts a KD(O2) value of 0.7 mM for BNCox at 37 °C (Section 4.3.3), greater than the 0.32 mM estimated by Chance and coworkers [193]. Measured values for BNCred, or in more likelihood BNCox [168], of 0.14–0.5 mM [90], 0.28 mM [230], and 0.285 mM [231] at 20 °C fall in that range. Furthermore, DFT calculations, albeit tentative, support weak O2 binding to BNCox [186]. Here [168], the stated “Physiological meaningful affinities” reinforces a misunderstanding in enzyme kinetics since the O2 capture is achieved through a transient ligand binding, SOC, ET switch activation (Figure 6 and Figure 7), and rapid O2 reduction (Figure 5). (2) The data and model indicate that the competition between NO and O2 occurs at BNCox, and the competition is explained by the kinetics of binding and turnover (Section 4.3.3). CO inhibition of CcO, like the flavoHb-NOD [18,62,74], can also be explained by the competition of CO with O2 for BNCox. (3) The new model proposes O2 binding to BNCox but otherwise shares intermediates and timing with the current models (Figure 5) [163,164,191,194,232]. Albeit, in the models examined, the order and presumed role of the tyrosyl radical intermediate, PM or OT (Figure 5), differs. Otherwise, the ET switch mechanism simply demands that the reduction of O2 and oxygen intermediates precede their protonation since the anions form the motor or pump action (Figure 7). (4) The slow rate of reduction of BNCox observed in many studies may be attributed to the ET switch mechanism of the model (Figure 6 and Figure 7) (Section 4.3.3). The switch may also explain O2 activation [181,233], ‘pulsed’ and ‘resting’ states [162], slow and fast phases of O2 reduction [230,231,234], slow and fast electrogenic proton pumping phenomena in the absence of O2 [188], and the unusual oxidative and reductive phases for proton pumping [235]. As proposed, ET requires ferromagnetic SOC of O2 and Cu2+, antiferromagnetic SOC of a3-Fe3+ with O2Cu2+, SC in the ferric heme, electron withdrawal from the a3 propionates, and proton migration from propionate a to propionate d. Also, the reduction requires PCET from the reduced cytochrome a to the oxidized a3 propionate d. Anything that interferes with or dysregulates the ET switch mechanism in vitro will turn ET OFF (or ON). In addition, the proposed ET switch would also prevent the untimely reduction of the ferryl intermediate (FH), formation of highly reactive hydroxyl radical, and potentially irreversible damage to CcO. Where a slow phase has been observed, and the electron supply does not appear limiting and may be dysregulated [234] or the O2 SOC may be impaired by a hydrated (OW) and protonated BNCox, the reversibility of function should be examined and the damage, protonation, and hydration assessed. Given the observed single turnover fast to slow phase transition [234], a failed dehydrating Ow → O cycle step that occurs with a non-existent membrane potential in the isolated CcO, with possible Pro tunnel/pump dysfunction, appears most attractive. Not recognizing, or accounting for, the existence of a sensitive and rapid ET gate or switch (Figure 7) and/or a proton uptake and water expulsion tunnel (Figure 8) and OW (OH) dehydration mechanism would cause incalculable confusion. For well over 50 years, the fast-slow phase a3 reduction phenomenon [234] has been interpreted as evidence for the requirement of a 2 e- reduction of the BNC for O2 binding [159,163]. The NO sensitivity-resistance of CcO discussed in Section 4.2 and Section 4.3 and the paucity of R in respiring mitochondria [101,102] strongly suggests otherwise. According to the model calculation estimates (Section 4.3.3), with regard to O2 collision, SOC appears rate-limiting for turnover at 1 µM O2, the approximate KM(O2) value [197,198]. O2 appears efficient at triggering the ET switch mechanism during the steady-state given the estimated 1 O2 transient binding interaction per 1 turnover for a 1/2 Vmax = 100 s−1. Indeed, given the underlying physics and chemistry, the trigger process would likely occur in picoseconds. At O2 saturation with an ~3 ms turnover time, the mechanism for water dissociation from OW may involve large motions and be rate-limiting for the catalytic cycle. As proposed (Figure 5), it appears unlikely that an O2 collision alone would efficiently displace the water molecules. The process may require milliseconds and energy. For example, gating may be required. At least three different “conformations” of BNCox have been detected and studied by NO binding and EPR, and differences in O2 accessibility have been discussed [236]. If we imagine a cycle clock for CcO similar to that drawn for the flavoHb-NOD (Figure 2), extrusion of water from OW, or dehydrating OW to form the fast ‘O’ state [163], would likely be apportioned the largest time. (5) The model provides a detailed map for charge migration (Figure 7, Figure 8 and Figure 9) and explains how the necessary unidirectionality of proton migration is achieved. It also explains how the electron migration path bifurcates from the proton migration path to allow strong electrostatic field forces from the magnesium ion to influence the charge movements (Figure 9). At present, the model is not sufficiently developed to describe the complex and rapid timing of synchronized countercurrent charge movements, but proton pairing with electrons through head-on or side-to-side meetings, or a countercurrent form of PCET, such as concerted EPT [187], is expected except for the storage and release of electrons from CuA or cytochrome a. An electron would migrate from one migrating positive charge to another migrating positive charge with little or no energy barrier and act as a conductor. A description of the synchronized motions may benefit from advanced computational modelling of PCET.
Experiments and theoretical work appear to support the general model (Figure 5, Figure 6, Figure 7, Figure 8 and Figure 9). I will only mention a few here. Arg438 and Arg439 are thought to be important for proton pumping and substitution mutations impair CcO function [237,238,239]. Cytochrome a and cytochrome a3 propionates have long been considered important for ET and proton pumping [182,184,201,229,240,241]. The evidence for NO protection against cyanide, CO, or sulfide inhibition of CcO [242,243] may be explained by the competition with O2 for BNCox and by the relative association and dissociation constants. More time and space will be required to thoroughly examine the literature and make all the proper attributions.
The change to the current catalytic cycle models [163,164,168,184,206,217,232] is small, but the improvements may help resolve some unanswered and puzzling questions. Revisions and improvements to consider include
(1)
O2, NO, and CO binding to BNCox (Figure 5).
(2)
SOC and spintronics (Figure 6).
(3)
an ET switch mechanism (Figure 7)
(4)
the alpha helix dipole-driven and membrane potential-opposed Pro tunnel (Figure 8)
(5)
the ET-linked proton pump mechanism (Figure 7)
(6)
the existence of well-defined electrostatically-coordinated and concerted EPT pathways (Figure 9)
(7)
proton reverse flow and ET pathway control by the hydrated magnesium ion (Figure 9)
(8)
the proton or hydronium ion exit channel (Figure 9)
(9)
the free energy and mechanism for pumping first proton in the cycle comes from reduction of ferromagnetically coupled O2Cu2+, not O2 per se
(10)
other concepts mentioned but not listed here
The CcO model revisions and concepts that I have only tersely expounded upon can be readily tested, improved, expanded upon, and superseded with an in-depth analysis of new and existing structures and data, advanced molecular dynamics simulations, and state-of-the-art DFT calculations. As always, attention to both contradictions and agreements in the development of a comprehensive and coherent mechanistic model is essential. Physiological evidence and functional behavior should take precedence. For example, future investigations of the kinetic competition of NO, CO, cyanide, and sulfide with O2 should test the model in vitro and within intact mitochondria with thorough kinetic analyses, including binding constants. Investigators of CO inhibition will note that, in 1951, Ball et al. [244] stated that CcO “exists in both an oxidized and reduced form, while CO and O2 are competing only for the reduced form “. But recall again that the reduced form was not seen in respiring mitochondria [101]. Keilin and Hartree [158] had already recognized the contradiction between the light sensitivity of CO inhibition and the light insensitivity of the reduced CcO-CO complex as early as 1939. Keilin had stated earlier that ferric hematin bound CO but reduced hematin bound CO with higher affinity [245]. If Keilin had seen and confronted the light (in)sensitivity paradox then or if others [168,186,193] had pursued the relevance of BNCox transiently interacting with O2, we can only imagine where CcO (and NOD) investigations would stand today. Indeed, CO reduces BNCox [246] and may also transiently and competitively interact with OW (or O) to cause inhibition. SQUID-EPR [247], especially as applied to CcO [248], has the potential for measuring weak and transient O2 interactions with BNCox at the low temperatures used by Chance et al. [193] as well as at higher temperatures. While spin parity and neutrality, as predicted for ferromagnetic SOC in the model (Figure 6), may preclude detection, the net loss of O2 spin through SOC with BNCox may be apparent. In the classic experiments of Van Gelder et al. [249,250] and Brudvig et al. [236], fluoride was shown to weakly interact with BNCox by EPR or UV–visible spectroscopy. Competition between O2 and another weak, yet detectable, ligand such as fluoride or CO may be used to indirectly measure the KD(O2) value for BNCox. We can also begin to ask whether SOC with electron-withdrawing O2Cu2+ (Section 4.3.4) could produce the spectral changes recorded by Chance et al. [193]. Small differences in the Qv or β (540 nm), α (590 nm), and CT (612 nm) bands of the ferric a3 heme, similar to those recorded for the Mb ferric heme [74], may serve as sensitive indicators of a transient O2 SOC. The failure of Kaila et al. [186] to observe spectral differences in CcO BNCox with 280 µM O2 may be explained by the uptake of excess protons by the putative Pro tunnel (Section 4.3.4, Figure 8) and a hydrated BNCox (OW). In the absence of an opposing membrane potential, protonation of the a3 propionates would decrease the electron density of the heme required for O2 binding. Indeed, the CcO Soret maxima (Kaila et al., Fig. 6) appears blue-shifted at ~421 nm vs. the ~424 nm recorded by Chance et al. [251], which may occur with propionate carboxylate protonation, hydration, or an increase in high-spin iron [74]. Water bound to the BNCox (OW) may also cause a failure to observe the proposed O2 SOC. Dehydrating BNCox to an O intermediate state [163] may be essential for significant O2 interaction. A similar water and O2 competition is apparent in ferric Mb [74]. Looking upstream of the ET switch, site-directed mutagenesis of the Pro tunnel similar to prior quantitative investigations of the D, K [219], and H pathways [218] can be readily utilized to test and analyze function. The model (Figure 8) and prior investigations predict a large influence of the membrane potential on proton uptake, delivery, and dysregulation of the ET switch mechanism (Figure 7). A collapsed membrane potential should lead to R intermediate formation and a greater NO sensitivity of CcO, and it apparently does (see above) [80,101,102]. When the membrane potential of heart mitochondria is collapsed with ADP in state 4 (note that authors call this state 3), the NO sensitivity of CcO increases manyfold [80]. What is the quantitative relationship of the membrane potential to R formation? How does the membrane potential compare to the helix dipole electrical potential energy (Figure 8)? Does the membrane or an external electrical potential affect a3 propionate a and d protonation and R formation as supposed by the model? Could engineered mutations inhibit the Pro tunnel, protonation, and ET dysregulation in the isolated CcO? Could an altered Pro tunnel and protonation explain the >10-fold smaller a to a3 ET rate observed with low phospholipid CcO [252]. Does C18O2 formation via 18O2 respiration, H218O formation, and fatty acid oxidation depend on CcO and the putative Pro tunnel or (de)hydrating water pump? Looking downstream, beryllium (Be2+), with a charge to radius ratio 1.5-fold larger than Mg2+ [226], may replace Mg2+ and alter proton pumping and the CcO function, possibly accounting, in part, for respiratory toxicity [253,254]. Further downstream, the exit channel for hydronium ions and protons provides a likely route also for HSO3 entry and for the initiation of the free radical chains of sulfite oxidation by copper. Given the relatively large rate constants estimated for CcO-catalyzed sulfite oxidation (k2 = ~460 M−1 s−1) [71,255] and possible self-inflicted sulfur trioxide anion radical damage to CcO, the identified channel to the CuA site may be relevant to the susceptibility of individuals to the airway-constrictive, asthma-inducing, headache-causing, and rash-inducing effects of SO2 in pollution, fumigated fruits, and treated wines. As always, revisions to the model and new concepts will be needed to accommodate old and new observations.

5. Pathology of Irreversible NO Poisoning

NO-resistant NODs and the NO resistance mechanism of CcO have clearly evolved to avoid more harmful NO poisoning of respiratory pathways, the citric acid cycle, etc., in pathological conditions including hypoxia, stroke, ischemia–reperfusion injury, septic shock, and inflammation. Conditions which lead to dysregulated ET or a slow leaky ET in either NODs or CcO to form their reduced forms, as observed in hypoxic cardiac muscle mitochondria [101], are expected to increase NO poisoning and damage to tissues. Under the reducing conditions of hypoxia or ischemia, stoichiometric NO scavengers, such as deoxyMb or oxyMb, would show a limited capacity for sustained NO scavenging and protection in most cells. Mb-rich cardiac myocytes may be the exception. with a greater capacity for sustained NO scavenging [87]. The O2-independent NOR activities observed in cells or mitochondria [16,80] may be essential under these conditions. Numerous investigations of microbial NO toxicity and resistance have led to similar hypotheses and conclusions [70,170,256]. A greater knowledge of ET regulation and dysregulation in CcO and globin-NODs and the identification of the mammalian NORs may increase our understanding of NO pathology and resistance. Given the proposed role of impaired membrane potential in various conditions including aging and Alzheimer’s disease [222] or even extreme exercise and work, ET control dysregulation in CcO and an enhanced NO sensitivity of CcO [80] may be more generally important during fever, vaccinations, and other inflammatory conditions. For this reason, NO-resistant NODs, including Ngb [257,258], cytoglobin [259,260,261,262,263], and monomeric Hbs [264] and Mb [265] with their tightly-coupled reductases [18], may be even more critical in Alzheimer’s and other hypoxic NO stress conditions.

6. Prospective Summary

FlavoHbs and related globins with their coupled reductases should never function as NODs, but they do. The extreme NO affinity and competition with O2 for ferrous heme binding should prohibit turnover, but it does not. Recent data and theory strongly suggest that NOD activity is resistant due to an allosteric ET gating mechanism in which the ferric heme is only reduced when O2 is present. Together, NO and O2 are proposed to elicit heme SC, doming, conformational changes, and electron transfer. By the same token, CcO should be irreversibly inhibited by NO, but it is not. The only explanation consistent with the reversibility of NO inhibition is a mechanism similar to that described for the NODs in which a heme-Fe3+O2Cu2+ complex triggers the initial electron transfer to form Fe3+O2Cu2+. A subtle modification to the current mechanistic model for CcO and related terminal oxidases is suggested in which O2 and reduced oxygen intermediates affect the heme spin state, heme propionate protonation, ET, and proton pumping. A greater understanding of the mechanism of the CcO is achieved through a thorough reappraisal of the NO resistance as well as the reinterpretation of assorted phenomena and structures. Given the ubiquity of NO, there are likely other examples of heme-based ET switching. The O2-metabolizing P450s and NOSs are attractive candidates for further investigation.

Funding

This work was initiated with funding from the American Heart Association Scientist Development Grant 9730193 N, the National Institutes of Health grant R01 GM65090, and the Cincinnati Children’s Hospital Research Foundation Trustees at the Cincinnati Children’s Hospital Medical Center.

Data Availability Statement

No new data were created or analyzed in this study.

Acknowledgments

I gratefully acknowledge discussions of BNCox ligand binding with the late Mårten Wikström and Peter Rich in May of 2023. I thank the reviewers for beneficent, incisive, insightful, and challenging comments. I dedicate this work in tribute to Irwin Fridovich (1929–2019), who enthusiastically encouraged searching and wandering in the laboratory and “the gardens of the mind” and was always eager to share in the journey. The content of this work is solely the responsibility of the author and does not necessarily represent the official views of the National Institutes of Health or the other funders.

Conflicts of Interest

The author declares no conflicts of interest.

Abbreviations

NORNO reductase
NODNO dioxygenase
NOSNO synthase
HbHemoglobin
MbMyoglobin
CcOCytochrome c Oxidase
flavoHbflavohemoglobin
FADFlavin Adenine Dinucleotide
ETElectron Transfer
NgbNeuroglobin;
SOCSpin-Orbit Coupling
SCSpin Crossover
SISpin Inversion
FMFerromagnetic
AFMAntiferromagnetic
BNCBinuclear Center
BNCredthe Reduced BNC
BNCoxthe Oxidized BNC
EPRElectron Paramagnetic Resonance
DFTDensity Functional Theory
PTProton Transfer
EPTElectron–Proton Transfer
PCETProton-Coupled Electron Transfer
SQUIDSuperconducting Quantum Interference Device

References

  1. Davy, H. Humphry Davy’s Physiologische-Chemische Untersuchungen Über Das Athmen, Besonders Über Das Athmen Von Oxydirtem Stickgas (Transl.); Menerschen Buchhandlung: Lemgo, Germany, 1814; pp. 51–58. [Google Scholar]
  2. Koppenol, W.H. Thermodynamic considerations on the generation of hydroxyl radicals from nitrous oxide—No laughing matter. Free Radic. Biol. Med. 1991, 10, 85–87. [Google Scholar] [CrossRef] [PubMed]
  3. Zumft, W.G. Nitric oxide reductases of prokaryotes with emphasis on the respiratory, heme-copper oxidase type. J. Inorg. Biochem. 2005, 99, 194–215. [Google Scholar] [CrossRef]
  4. Gardner, P.R. Nitric oxide dioxygenase function and mechanism of flavohemoglobin, hemoglobin, myoglobin and their associated reductases. J. Inorg. Biochem. 2005, 99, 247–266. [Google Scholar] [CrossRef] [PubMed]
  5. Gardner, P.R. Hemoglobin: A nitric-oxide dioxygenase. Scientifica 2012, 2012, 683729. [Google Scholar] [CrossRef] [PubMed]
  6. Green, L.C.; Wagner, D.A.; Glogowski, J.; Skipper, P.L.; Wishnok, J.S.; Tannenbaum, S.R. Analysis of nitrate, nitrite, and [15N] nitrate in biological fluids. Anal. Biochem. 1982, 126, 131–138. [Google Scholar] [CrossRef] [PubMed]
  7. Hibbs, J.B., Jr.; Taintor, R.R.; Vavrin, Z. Macrophage cytotoxicity: Role for L-arginine deiminase and imino nitrogen oxidation to nitrite. Science 1987, 235, 473–476. [Google Scholar] [CrossRef]
  8. MacMicking, J.; Xie, Q.W.; Nathan, C. Nitric oxide and macrophage function. Annu. Rev. Immunol. 1997, 15, 323–350. [Google Scholar] [CrossRef]
  9. Drapier, J.C.; Hibbs, J.B., Jr. Murine cytotoxic activated macrophages inhibit aconitase in tumor cells. Inhibition involves the iron-sulfur prosthetic group and is reversible. J. Clin. Investig. 1986, 78, 790–797. [Google Scholar] [CrossRef]
  10. Stuehr, D.J.; Nathan, C.F. Nitric oxide. A macrophage product responsible for cytostasis and respiratory inhibition in tumor target cells. J. Exp. Med. 1989, 169, 1543–1555. [Google Scholar] [CrossRef]
  11. Mohanty, I.; Fazelinia, H.; Raimo, S.; Ding, H.; Spruce, L.; Symeonidou, M.; Tenopoulou, M.; Ichinose, F.; Yamashita, H.; Shimokawa, H.; et al. Proteomic surveys of the mouse heart unveil cardiovascular responses to nitric oxide/cGMP signaling deficiencies. Commun. Biol. 2025, 8, 817. [Google Scholar] [CrossRef]
  12. Dent, M.R.; DeMartino, A.W.; Tejero, J.; Gladwin, M.T. Endogenous hemoprotein-dependent signaling pathways of nitric oxide and nitrite. Inorg. Chem. 2021, 60, 15918–15940. [Google Scholar] [CrossRef]
  13. Yunker, M.H.; Szulczewski, D.; Higuchi, T. Kinetics of the degradation of isoamyl nitrite in ampuls I.: Nature of reaction products and influence of temperature, oxygen, alcohol, water, and acid on the overall rate. J. Am. Pharm. Assoc. 1958, 47, 613–620. [Google Scholar] [CrossRef]
  14. Di Carlo, F.J. Nitroglycerin revisited: Chemistry, biochemistry, interactions. Drug Metab. Rev. 1975, 4, 1–38. [Google Scholar] [CrossRef]
  15. Ignarro, L. Nitric oxide as a unique signaling molecule in the vascular system: A historical overview. J. Physiol. Pharmacol. 2002, 53, 503–514. [Google Scholar] [CrossRef] [PubMed]
  16. Gardner, P.R.; Martin, L.A.; Hall, D.; Gardner, A.M. Dioxygen-dependent metabolism of nitric oxide in mammalian cells. Free Radic. Biol. Med. 2001, 31, 191–204. [Google Scholar] [CrossRef] [PubMed]
  17. Gardner, P.R.; Gardner, A.M.; Brashear, W.T.; Suzuki, T.; Hvitved, A.N.; Setchell, K.D.; Olson, J.S. Hemoglobins dioxygenate nitric oxide with high fidelity. J. Inorg. Biochem. 2006, 100, 542–550. [Google Scholar] [CrossRef] [PubMed]
  18. Gardner, P.R. Ordered motions in the nitric-oxide dioxygenase mechanism of flavohemoglobin and assorted globins with their tightly coupled reductases. In Protein Reviews; Atassi, M.Z., Ed.; Springer International Publishing: Cham, Switzerland, 2023; Volume 23, pp. 45–96. [Google Scholar]
  19. Doyle, M.P.; Hoekstra, J.W. Oxidation of nitrogen oxides by bound dioxygen in hemoproteins. J. Inorg. Biochem. 1981, 14, 351–358. [Google Scholar] [CrossRef]
  20. Eich, R.F.; Li, T.S.; Lemon, D.D.; Doherty, D.H.; Curry, S.R.; Aitken, J.F.; Mathews, A.J.; Johnson, K.A.; Smith, R.D.; Phillips, G.N.; et al. Mechanism of NO-induced oxidation of myoglobin and hemoglobin. Biochemistry 1996, 35, 6976–6983. [Google Scholar] [CrossRef]
  21. Wade, R.S.; Castro, C.E. Reactions of oxymyoglobin with NO, NO2, and NO2 under argon and in air. Chem. Res. Toxicol. 1996, 9, 1382–1390. [Google Scholar] [CrossRef]
  22. Yukl, E.T.; de Vries, S.; Moënne-Loccoz, P. The millisecond intermediate in the reaction of nitric oxide with oxymyoglobin is an iron(III)—nitrato complex, not a peroxynitrite. J. Am. Chem. Soc. 2009, 131, 7234–7235. [Google Scholar] [CrossRef]
  23. Schopfer, M.P.; Mondal, B.; Lee, D.-H.; Sarjeant, A.A.N.; Karlin, K.D. Heme/O2/•NO nitric oxide dioxygenase (NOD) reactivity: Phenolic nitration via a putative heme-peroxynitrite intermediate. J. Am. Chem. Soc. 2009, 131, 11304–11305. [Google Scholar] [CrossRef]
  24. Möller, M.N.; Rios, N.; Trujillo, M.; Radi, R.; Denicola, A.; Alvarez, B. Detection and quantification of nitric oxide–derived oxidants in biological systems. J. Biol. Chem. 2019, 294, 14776–14802. [Google Scholar] [CrossRef] [PubMed]
  25. Ford, P.C.; Miranda, K.M. The solution chemistry of nitric oxide and other reactive nitrogen species. Nitric Oxide 2020, 103, 31–46. [Google Scholar] [CrossRef] [PubMed]
  26. Lancaster, J.R., Jr.; Hibbs, J.B., Jr. EPR demonstration of iron-nitrosyl complex formation by cytotoxic activated macrophages. Proc. Natl. Acad. Sci. USA 1990, 87, 1223–1227. [Google Scholar] [CrossRef] [PubMed]
  27. Drapier, J.-C.; Hibbs, J.B., Jr. Aconitases: A class of metalloproteins highly sensitive to nitric oxide synthesis. Methods Enzymol. 1996, 269, 26–36. [Google Scholar] [CrossRef]
  28. Gardner, P.R.; Costantino, G.; Szabo, C.; Salzman, A.L. Nitric oxide sensitivity of the aconitases. J. Biol. Chem. 1997, 272, 25071–25076. [Google Scholar] [CrossRef]
  29. Brown, G.C.; Cooper, C.E. Nanomolar concentrations of nitric oxide reversibly inhibit synaptosomal cytochrome oxidase respiration by competing with oxygen at cytochrome oxidase. FEBS Lett. 1994, 356, 295–298. [Google Scholar] [CrossRef]
  30. Cleeter, M.W.J.; Cooper, J.M.; Darley-Usmar, V.M.; Moncada, S.; Schapira, A.H.V. Reversible inhibition of cytochrome c oxidase, the terminal enzyme of the mitochondrial respiratory chain, by nitric oxide: Implications for neurodegenerative diseases. FEBS Lett. 1994, 345, 50–54. [Google Scholar] [CrossRef]
  31. Kim, Y.-M.; Bergonia, H.A.; Müller, C.; Pitt, B.R.; Watkins, W.D.; Lancaster, J.R., Jr. Loss and degradation of enzyme-bound heme induced by cellular nitric oxide synthesis. J. Biol. Chem. 1995, 270, 5710–5713. [Google Scholar] [CrossRef]
  32. Takemura, S.; Minamiyama, Y.; Imaoka, S.; Funae, Y.; Hirohashi, K.; Inoue, M.; Kinoshita, H. Hepatic cytochrome P450 is directly inactivated by nitric oxide, not by inflammatory cytokines, in the early phase of endotoxemia. J. Hepatol. 1999, 30, 1035–1044. [Google Scholar] [CrossRef]
  33. Minamiyama, Y.; Takemura, S.; Imaoka, S.; Funae, Y.; Tanimoto, Y.; Inoue, M. Irreversible inhibition of cytochrome P450 by nitric oxide. J. Pharmacol. Exp. Therap. 1997, 283, 1479–1485. [Google Scholar] [CrossRef]
  34. Arnold, W.P.; Mittal, C.K.; Katsuki, S.; Murad, F. Nitric oxide activates guanylate cyclase and increases guanosine 3′: 5′-cyclic monophosphate levels in various tissue preparations. Proc. Natl. Acad. Sci. USA 1977, 74, 3203–3207. [Google Scholar] [CrossRef] [PubMed]
  35. Bouton, C.; Raveau, M.; Drapier, J.-C. Modulation of iron regulatory protein functions. Further insights into the role of nitrogen- and oxygen-derived reactive species. J. Biol. Chem. 1996, 271, 2300–2306. [Google Scholar] [CrossRef] [PubMed]
  36. Kennedy, M.C.; Antholine, W.E.; Beinert, H. An EPR investigation of the products of the reaction of cytosolic and mitochondrial aconitases with nitric oxide. J. Biol. Chem. 1997, 272, 20340–20347. [Google Scholar] [CrossRef]
  37. Cooper, C.E.; Davies, N.A. Effects of nitric oxide and peroxynitrite on the cytochrome oxidase Km for oxygen: Implications for mitochondrial pathology. Biochim. Biophys. Acta (BBA)-Bioenerg. 2000, 1459, 390–396. [Google Scholar] [CrossRef]
  38. Brown, G.C. Nitric oxide regulates mitochondrial respiration and cell functions by inhibiting cytochrome oxidase. FEBS Lett. 1995, 369, 136–139. [Google Scholar] [CrossRef]
  39. Perutz, M.F. Mechanisms regulating the reactions of human hemoglobin with oxygen and carbon monoxide. Annu. Rev. Biochem. 1990, 52, 1–25. [Google Scholar] [CrossRef]
  40. Saraste, M.; Castresana, J. Cytochrome oxidase evolved by tinkering with denitrification enzymes. FEBS Lett. 1994, 341, 1–4. [Google Scholar] [CrossRef]
  41. Castresana, J.; Lübben, M.; Saraste, M.; Higgins, D.G. Evolution of cytochrome oxidase, an enzyme older than atmospheric oxygen. EMBO J. 1994, 13, 2516–2525. [Google Scholar] [CrossRef]
  42. Brunori, M. Wandering about allostery. Biol. Direct 2024, 19, 64. [Google Scholar] [CrossRef]
  43. Prebble, J.; Weber, B. Wandering in the Gardens of the Mind: Peter Mitchell and the Making of Glynn; Oxford University Press: New York, NY, USA, 2003; pp. 1–324. [Google Scholar]
  44. Rudolph, J.; Koschorreck, M.; Conrad, R. Oxidative and reductive microbial consumption of nitric oxide in a heathland soil. Soil Biol. Biochem. 1996, 28, 1389–1396. [Google Scholar] [CrossRef]
  45. Koschorreck, M.; Moore, E.; Conrad, R. Oxidation of nitric oxide by a new heterotrophic Pseudomonas sp. Arch. Microbiol. 1996, 166, 23–31. [Google Scholar] [CrossRef]
  46. Remde, A.; Slemr, F.; Conrad, R. Microbial production and uptake of nitric oxide in soil. FEMS Microbio. Ecol. 1989, 5, 221–230. [Google Scholar] [CrossRef]
  47. Schäfer, F.; Ralf, C. Metabolism of nitric oxide by Pseudomonas stutzeri in culture and in soil. FEMS Microbio. Ecol. 1993, 11, 119–127. [Google Scholar] [CrossRef]
  48. Gardner, P.R.; Costantino, G.; Salzman, A.L. Constitutive and adaptive detoxification of nitric oxide in Escherichia coli. Role of nitric-oxide dioxygenase in the protection of aconitase. J. Biol. Chem. 1998, 273, 26528–26533. [Google Scholar] [CrossRef]
  49. Gardner, P.R.; Gardner, A.M.; Martin, L.A.; Salzman, A.L. Nitric oxide dioxygenase: An enzymic function for flavohemoglobin. Proc. Natl. Acad. Sci. USA 1998, 95, 10378–10383. [Google Scholar] [CrossRef]
  50. Masters, B.S.S.; Kamin, H.; Gibson, Q.H.; Williams, C.H., Jr. Studies on the mechanism of microsomal triphosphopyridine nucleotide-cytochrome c reductase. J. Biol. Chem. 1965, 240, 921–931. [Google Scholar] [CrossRef]
  51. Gardner, A.M.; Martin, L.A.; Gardner, P.R.; Dou, Y.; Olson, J.S. Steady-state and transient kinetics of Escherichia coli nitric-oxide dioxygenase (flavohemoglobin). The B10 tyrosine hydroxyl is essential for dioxygen binding and catalysis. J. Biol. Chem. 2000, 275, 12581–12589. [Google Scholar] [CrossRef]
  52. Westfelt, U.N.; Benthin, G.; Lundin, S.; Stenqvist, O.; Wennmalm, Å. Conversion of inhaled nitric oxide to nitrate in man. Br. J. Pharmacol. 1995, 114, 1621–1624. [Google Scholar] [CrossRef]
  53. Hausladen, A.; Stamler, J.S. Is the flavohemoglobin a nitric oxide dioxygenase? Free Radic. Biol. Med. 2012, 53, 1209–1210. [Google Scholar] [CrossRef]
  54. Hausladen, A.; Gow, A.J.; Stamler, J.S. Nitrosative stress: Metabolic pathway involving the flavohemoglobin. Proc. Natl. Acad. Sci. USA 1998, 95, 14100–14105. [Google Scholar] [CrossRef] [PubMed]
  55. Forrester, M.T.; Foster, M.W. Protection from nitrosative stress: A central role for microbial flavohemoglobin. Free Radic. Biol. Med. 2012, 52, 1620–1633. [Google Scholar] [CrossRef] [PubMed]
  56. Poole, R.K. Flavohaemoglobin: The pre-eminent nitric oxide-detoxifying machine of microorganisms. F1000Research 2020, 9, 7. [Google Scholar] [CrossRef] [PubMed]
  57. Frey, A.D.; Farrés, J.; Bollinger, C.J.T.; Kallio, P.T. Bacterial hemoglobins and flavohemoglobins for alleviation of nitrosative stress in Escherichia coli. Appl. Environ. Microbiol. 2002, 68, 4835–4840. [Google Scholar] [CrossRef]
  58. Poole, R.K.; Anjum, M.F.; Membrillo-Hernandez, J.; Kim, S.O.; Hughes, M.N.; Stewart, V. Nitric oxide, nitrite, and Fnr regulation of hmp (flavohemoglobin) gene expression in Escherichia coli K-12. J. Bacteriol. 1996, 178, 5487–5492. [Google Scholar] [CrossRef]
  59. LaCelle, M.; Kumano, M.; Kurita, K.; Yamane, K.; Zuber, P.; Nakano, M.M. Oxygen-controlled regulation of the flavohemoglobin gene in Bacillus subtilis. J. Bacteriol. 1996, 178, 3803–3808. [Google Scholar] [CrossRef]
  60. Cramm, R.; Siddiqui, R.A.; Friedrich, B. Primary sequence and evidence for a physiological function of the flavohemoprotein of Alcaligenes eutrophus. J. Biol. Chem. 1994, 269, 7349–7354. [Google Scholar] [CrossRef]
  61. Zhu, H.; Riggs, A.F. Yeast flavohemoglobin is an ancient protein related to globins and a reductase family. Proc. Natl. Acad. Sci. USA 1992, 89, 5015–5019. [Google Scholar] [CrossRef]
  62. Gardner, P.R.; Gardner, A.M.; Martin, L.A.; Dou, Y.; Li, T.; Olson, J.S.; Zhu, H.; Riggs, A.F. Nitric-oxide dioxygenase activity and function of flavohemoglobins. Sensitivity to nitric oxide and carbon monoxide inhibition. J. Biol. Chem. 2000, 275, 31581–31587. [Google Scholar] [CrossRef]
  63. Hausladen, A.; Gow, A.; Stamler, J.S. Flavohemoglobin denitrosylase catalyzes the reaction of a nitroxyl equivalent with molecular oxygen. Proc. Natl. Acad. Sci. USA 2001, 98, 10108–10112. [Google Scholar] [CrossRef]
  64. Kim, S.O.; Orii, Y.; Lloyd, D.; Hughes, M.N.; Poole, R.K. Anoxic function for the Escherichia coli flavohaemoglobin (Hmp): Reversible binding of nitric oxide and reduction to nitrous oxide. FEBS Lett. 1999, 445, 389–394. [Google Scholar] [CrossRef]
  65. Poole, R.K.; Hughes, M.N. New functions for the ancient globin family: Bacterial responses to nitric oxide and nitrosative stress. Mol. Microbiol. 2000, 36, 775–783. [Google Scholar] [CrossRef] [PubMed]
  66. Gardner, A.M.; Gardner, P.R. Flavohemoglobin detoxifies nitric oxide in aerobic, but not anaerobic, Escherichia coli. Evidence for a novel inducible anaerobic nitric oxide-scavenging activity. J. Biol. Chem. 2002, 277, 8166–8171. [Google Scholar] [CrossRef] [PubMed]
  67. Wang, R.Z.; Lonergan, Z.R.; Wilbert, S.A.; Eiler, J.M.; Newman, D.K. Widespread detoxifying NO reductases impart a distinct isotopic fingerprint on N2O under anoxia. Proc. Natl. Acad. Sci. USA 2024, 121, e2319960121. [Google Scholar] [CrossRef] [PubMed]
  68. Puppo, A.; Pauly, N.; Boscari, A.; Mandon, K.; Brouquisse, R. Hydrogen peroxide and nitric oxide: Key regulators of the legume—Rhizobium and mycorrhizal symbioses. Antioxid. Redox. Signal. 2013, 18, 2202–2219. [Google Scholar] [CrossRef]
  69. Minning, D.M.; Gow, A.J.; Bonaventura, J.; Braun, R.; Dewhirst, M.; Goldberg, D.E.; Stamler, J.S. Ascaris haemoglobin is a nitric oxide-activated ‘deoxygenase’. Nature 1999, 401, 497–502. [Google Scholar] [CrossRef]
  70. Gardner, A.M.; Helmick, R.A.; Gardner, P.R. Flavorubredoxin, an inducible catalyst for nitric oxide reduction and detoxification in Escherichia coli. J. Biol. Chem. 2002, 277, 8172–8177. [Google Scholar] [CrossRef]
  71. Gardner, P.R.; Gardner, D.P.; Gardner, A.P. Globins scavenge sulfur trioxide anion radical. J. Biol. Chem. 2015, 290, 27204–27214. [Google Scholar] [CrossRef]
  72. Membrillo-Hernandez, J.; Coopamah, M.D.; Anjum, M.F.; Stevanin, T.M.; Kelly, A.; Hughes, M.N.; Poole, R.K. The flavohemoglobin of Escherichia coli confers resistance to a nitrosating agent, a “nitric oxide releaser,” and paraquat and is essential for transcriptional responses to oxidative stress. J. Biol. Chem. 1999, 274, 748–754. [Google Scholar] [CrossRef]
  73. Buisson, N.; Labbe-Bois, R. Flavohemoglobin expression and function in Saccharomyces cerevisiae. No relationship with respiration and complex response to oxidative stress. J. Biol. Chem. 1998, 273, 9527–9533. [Google Scholar] [CrossRef]
  74. Gardner, A.M.; Gardner, P.R. Dioxygen and glucose force motion of the electron-transfer switch in the iron(III) flavohemoglobin-type nitric oxide dioxygenase. J. Inorg. Biochem. 2023, 245, 112257. [Google Scholar] [CrossRef]
  75. Gardner, A.M.; Gardner, P.R. Allostery in the nitric oxide dioxygenase mechanism of flavohemoglobin. J. Biol. Chem. 2021, 296, 100186. [Google Scholar] [CrossRef]
  76. Nazmutdinov, R.R.; Santos, E.; Schmickler, W. Spin effects in oxygen electrocatalysis: A discussion. Electrochem. Commun. 2013, 33, 14–17. [Google Scholar] [CrossRef]
  77. Gracia, J. Itinerant spins and bond lengths in oxide electrocatalysts for oxygen evolution and reduction reactions. J. Phys. Chem. C 2019, 123, 9967–9972. [Google Scholar] [CrossRef]
  78. Sun, Y.; Sun, S.; Yang, H.; Xi, S.; Gracia, J.; Xu, Z.J. Spin-related electron transfer and orbital interactions in oxygen electrocatalysis. Adv. Mater. 2020, 32, 2003297. [Google Scholar] [CrossRef] [PubMed]
  79. Pech-Santiago, E.O.; Argüello-García, R.; Arce-Cruz, G.; Angeles, E.; Ortega-Pierres, G. Giardia duodenalis flavohemoglobin is a target of 5-nitroheterocycle and benzimidazole compounds acting as enzymatic inhibitors or subversive substrates. Free Radic. Biol. Med. 2025, 227, 355–366. [Google Scholar] [CrossRef] [PubMed]
  80. Borutaité, V.; Brown, G.C. Rapid reduction of nitric oxide by mitochondria, and reversible inhibition of mitochondrial respiration by nitric oxide. Biochem. J. 1996, 315, 295–299. [Google Scholar] [CrossRef]
  81. Cooper, C.E.; Torres, J.; Sharpe, M.A.; Wilson, M.T. Nitric oxide ejects electrons from the binuclear centre of cytochrome c oxidase by reacting with oxidised copper: A general mechanism for the interaction of copper proteins with nitric oxide? FEBS Lett. 1997, 414, 281–284. [Google Scholar] [CrossRef]
  82. Stevanin, T.M.; Ioannidis, N.; Mills, C.E.; Kim, S.O.; Hughes, M.N.; Poole, R.K. Flavohemoglobin Hmp affords inducible protection for Escherichia coli respiration, catalyzed by cytochromes bo’ or bd, from nitric oxide. J. Biol. Chem. 2000, 275, 35868–35875. [Google Scholar] [CrossRef]
  83. Palacios-Callender, M.; Hollis, V.; Mitchison, M.; Frakich, N.; Unitt, D.; Moncada, S. Cytochrome c oxidase regulates endogenous nitric oxide availability in respiring cells: A possible explanation for hypoxic vasodilation. Proc. Natl. Acad. Sci. USA 2007, 104, 18508–18513. [Google Scholar] [CrossRef]
  84. Daiber, A.; Shoun, H.; Ullrich, V. Nitric oxide reductase (P450nor) from Fusarium oxysporum. J. Inorg. Biochem. 2005, 99, 185–193. [Google Scholar] [CrossRef] [PubMed]
  85. Stubauer, G.; Giuffrè, A.; Brunori, M.; Sarti, P. Cytochrome c oxidase does not catalyze the anaerobic reduction of NO. Biochem. Biophys. Res. Commun. 1998, 245, 459–465. [Google Scholar] [CrossRef] [PubMed]
  86. Brudvig, G.W.; Stevens, T.H.; Chan, S.I. Reactions of nitric oxide with cytochrome c oxidase. Biochemistry 1980, 19, 5275–5285. [Google Scholar] [CrossRef] [PubMed]
  87. Giles, A.V.; Edwards, L.; Covian, R.; Lucotte, B.M.; Balaban, R.S. Cardiac nitric oxide scavenging: Role of myoglobin and mitochondria. J. Physiol. 2024, 602, 73–91. [Google Scholar] [CrossRef]
  88. Giuffrè, A.; Forte, E.; Brunori, M.; Sarti, P. Nitric oxide, cytochrome c oxidase and myoglobin: Competition and reaction pathways. FEBS Lett. 2005, 579, 2528–2532. [Google Scholar] [CrossRef]
  89. Moore, E.G.; Gibson, Q.H. Cooperativity in the dissociation of nitric oxide from hemoglobin. J. Biol. Chem. 1976, 251, 2788–2794. [Google Scholar] [CrossRef]
  90. Blackmore, R.S.; Greenwood, C.; Gibson, Q.H. Studies of the primary oxygen intermediate in the reaction of fully reduced cytochrome oxidase. J. Biol. Chem. 1991, 266, 19245–19249. [Google Scholar] [CrossRef]
  91. Torres, J.; Darley-Usmar, V.; Wilson, M.T. Inhibition of cytochrome c oxidase in turnover by nitric oxide: Mechanism and implications for control of respiration. Biochem. J. 1995, 312, 169–173. [Google Scholar] [CrossRef]
  92. Giuffrè, A.; Sarti, P.; D’Itri, E.; Buse, G.; Soulimane, T.; Brunori, M. On the mechanism of inhibition of cytochrome c oxidase by nitric oxide. J. Biol. Chem. 1996, 271, 33404–33408. [Google Scholar] [CrossRef]
  93. Sarti, P.; Giuffré, A.; Forte, E.; Mastronicola, D.; Barone, M.C.; Brunori, M. Nitric oxide and cytochrome c oxidase: Mechanisms of inhibition and NO degradation. Biochem. Biophys. Res. Commun. 2000, 274, 183–187. [Google Scholar] [CrossRef]
  94. Mason, M.G.; Nicholls, P.; Wilson, M.T.; Cooper, C.E. Nitric oxide inhibition of respiration involves both competitive (heme) and noncompetitive (copper) binding to cytochrome c oxidase. Proc. Natl. Acad. Sci. USA 2006, 103, 708–713. [Google Scholar] [CrossRef]
  95. Antunes, F.; Boveris, A.; Cadenas, E. On the mechanism and biology of cytochrome oxidase inhibition by nitric oxide. Proc. Natl. Acad. Sci. USA 2004, 101, 16774–16779. [Google Scholar] [CrossRef]
  96. Mason, M.G.; Shepherd, M.; Nicholls, P.; Dobbin, P.S.; Dodsworth, K.S.; Poole, R.K.; Cooper, C.E. Cytochrome bd confers nitric oxide resistance to Escherichia coli. Nat. Chem. Biol. 2009, 5, 94–96. [Google Scholar] [CrossRef] [PubMed]
  97. Borisov, V.B.; Forte, E.; Sarti, P.; Brunori, M.; Konstantinov, A.A.; Giuffrè, A. Redox control of fast ligand dissociation from Escherichia coli cytochrome bd. Biochem. Biophys. Res. Commun. 2007, 355, 97–102. [Google Scholar] [CrossRef] [PubMed]
  98. Giuffre, A.; Borisov, V.B.; Mastronicola, D.; Sarti, P.; Forte, E. Cytochrome bd oxidase and nitric oxide: From reaction mechanisms to bacterial physiology. FEBS Lett. 2012, 586, 622–629. [Google Scholar] [CrossRef] [PubMed]
  99. Giuffrè, A.; Borisov, V.B.; Arese, M.; Sarti, P.; Forte, E. Cytochrome bd oxidase and bacterial tolerance to oxidative and nitrosative stress. Biochim. Biophys. Acta (BBA)-Bioenerg. 2014, 1837, 1178–1187. [Google Scholar] [CrossRef]
  100. Forte, E.; Urbani, A.; Saraste, M.; Sarti, P.; Brunori, M.; Giuffrè, A. The cytochrome cbb3 from Pseudomonas stutzeri displays nitric oxide reductase activity. Eur. J. Biochem. 2001, 268, 6486–6491. [Google Scholar] [CrossRef]
  101. Covian, R.; Edwards, L.O.; Lucotte, B.M.; Balaban, R.S. Spectroscopic identification of the catalytic intermediates of cytochrome c oxidase in respiring heart mitochondria. Biochim. Biophys. Acta (BBA)-Bioenerg. 2023, 1864, 148934. [Google Scholar] [CrossRef]
  102. Covian, R.; Edwards, L.O.; Balaban, R.S. Effect of the mitochondrial membrane potential on the absorbance of the reduced form of cytochrome c oxidase. Biochim. Biophys. Acta (BBA)-Bioenerg. 2024, 1865, 149048. [Google Scholar] [CrossRef]
  103. Gibson, Q.; Roughton, F. The kinetics and equilibria of the reactions of nitric oxide with sheep haemoglobin. J. Physiol. 1957, 136, 507–526. [Google Scholar] [CrossRef]
  104. Foley, E.L.; Hvitved, A.N.; Eich, R.F.; Olson, J.S. Mechanisms of nitric oxide reactions with globins using mammalian myoglobin as a model system. J. Inorg. Biochem. 2022, 233, 111839. [Google Scholar] [CrossRef] [PubMed]
  105. Hille, R.; Palmer, G.; Olson, J.S. Chain equivalence in reaction of nitric oxide with hemoglobin. J. Biol. Chem. 1977, 252, 403–405. [Google Scholar] [CrossRef] [PubMed]
  106. Frey, A.D.; Kallio, P.T. Bacterial hemoglobins and flavohemoglobins: Versatile proteins and their impact on microbiology and biotechnology. FEMS Microbiol. Rev. 2003, 27, 525–545. [Google Scholar] [CrossRef] [PubMed]
  107. Segel, I.H. Enzyme Kinetics: Behavior and Analysis of Rapid Equilibrium and Steady-State Enzyme Systems; John Wiley & Sons, Inc.: New York, NY, USA, 1975; pp. 826–830. [Google Scholar]
  108. Mills, C.E.; Sedelnikova, S.; Soballe, B.; Hughes, M.N.; Poole, R.K. Escherichia coli flavohaemoglobin (Hmp) with equistoichiometric FAD and haem contents has a low affinity for dioxygen in the absence or presence of nitric oxide. Biochem. J. 2001, 353, 207–213. [Google Scholar] [CrossRef]
  109. Hill, S.; Decorso, I.; Nezamololama, N.; Babaei, Z.; Rafferty, S.P. Catalytic differences between flavohemoglobins of Giardia intestinalis and E. coli. Pathogens 2024, 13, 480. [Google Scholar] [CrossRef]
  110. Ilangovan, G.; Khaleel, S.A.; Kundu, T.; Hemann, C.; El-Mahdy, M.A.; Zweier, J.L. Defining the reducing system of the NO dioxygenase cytoglobin in vascular smooth muscle cells and its critical role in regulating cellular NO decay. J. Biol. Chem. 2021, 296, 100196. [Google Scholar] [CrossRef]
  111. Gardner, A.M.; Cook, M.R.; Gardner, P.R. Nitric-oxide dioxygenase function of human cytoglobin with cellular reductants and in rat hepatocytes. J. Biol. Chem. 2010, 285, 23850–23857. [Google Scholar] [CrossRef]
  112. Liu, X.; Follmer, D.; Zweier, J.R.; Huang, X.; Hemann, C.; Liu, K.; Druhan, L.J.; Zweier, J.L. Characterization of the function of cytoglobin as an oxygen-dependent regulator of nitric-oxide concentration. Biochemistry 2012, 51, 5072–5082. [Google Scholar] [CrossRef]
  113. Zhou, S.; Fushinobu, S.; Nakanishi, Y.; Kim, S.-W.; Wakagi, T.; Shoun, H. Cloning and characterization of two flavohemoglobins from Aspergillus oryzae. Biochem. Biophys. Res. Commun. 2009, 381, 7–11. [Google Scholar] [CrossRef]
  114. Gonçalves, V.L.; Nobre, L.S.; Vicente, J.B.; Teixeira, M.; Saraiva, L.M. Flavohemoglobin requires microaerophilic conditions for nitrosative protection of Staphylococcus aureus. FEBS Lett. 2006, 580, 1817–1821. [Google Scholar] [CrossRef]
  115. Gardner, P.R. Assay and characterization of the NO dioxygenase activity of flavohemoglobins. Methods Enzymol. 2008, 436, 217–237. [Google Scholar] [CrossRef]
  116. Liu, L.; Zeng, M.; Hausladen, A.; Heitman, J.; Stamler, J.S. Protection from nitrosative stress by yeast flavohemoglobin. Proc. Natl. Acad. Sci. USA 2000, 97, 4672–4676. [Google Scholar] [CrossRef]
  117. Crawford, M.J.; Goldberg, D.E. Role for the Salmonella flavohemoglobin in protection from nitric oxide. J. Biol. Chem. 1998, 273, 12543–12547. [Google Scholar] [CrossRef]
  118. Justino, M.C.; Vicente, J.B.; Teixeira, M.; Saraiva, L.M. New genes implicated in the protection of anaerobically grown Escherichia coli against nitric oxide. J. Biol. Chem. 2005, 280, 2636–2643. [Google Scholar] [CrossRef]
  119. Nakano, M.M. Essential role of flavohemoglobin in long-term anaerobic survival of Bacillus subtilis. J. Bacteriol. 2006, 188, 6415–6418. [Google Scholar] [CrossRef] [PubMed]
  120. Platzen, L.; Koch-Koerfges, A.; Weil, B.; Brocker, M.; Bott, M. Role of flavohaemoprotein Hmp and nitrate reductase NarGHJI of Corynebacterium glutamicum for coping with nitrite and nitrosative stress. FEMS Microbiol. Lett. 2014, 350, 239–248. [Google Scholar] [CrossRef] [PubMed]
  121. Arai, H.; Hayashi, M.; Kuroi, A.; Ishii, M.; Igarashi, Y. Transcriptional regulation of the flavohemoglobin gene for aerobic nitric oxide detoxification by the second nitric oxide-responsive regulator of Pseudomonas aeruginosa. J. Bacteriol. 2005, 187, 3960–3968. [Google Scholar] [CrossRef] [PubMed]
  122. Dalsing, B.L.; Truchon, A.N.; Gonzalez-Orta, E.T.; Milling, A.S.; Allen, C. Ralstonia solanacearum uses inorganic nitrogen metabolism for virulence, ATP production, and detoxification in the oxygen-limited host xylem environment. mBio 2015, 6, e02471. [Google Scholar] [CrossRef]
  123. Takaya, N.; Shoun, H. Genetic engineering using fungal flavohemoglobin for constructing Pseudomonas stutzeri strain emitting less nitrous oxide. J. Biosci. Bioeng. 2002, 94, 282–284. [Google Scholar] [CrossRef]
  124. Kaldorf, M.; Linne von Berg, K.H.; Meier, U.; Servos, U.; Bothe, H. The reduction of nitrous oxide to dinitrogen by Escherichia coli. Arch. Microbiol. 1993, 160, 432–439. [Google Scholar] [CrossRef]
  125. Herold, S.; Fago, A.; Weber, R.E.; Dewilde, S.; Moens, L. Reactivity studies of the Fe(III) and Fe(II)NO forms of human neuroglobin reveal a potential role against oxidative stress. J. Biol. Chem. 2004, 279, 22841–22847. [Google Scholar] [CrossRef] [PubMed]
  126. Arnold, E.V.; Bohle, D.S. Isolation and oxygenation reactions of nitrosylmyoglobins. Methods Enzymol. 1996, 269, 41–55. [Google Scholar] [CrossRef] [PubMed]
  127. Herold, S.; Röck, G. Mechanistic studies of the oxygen-mediated oxidation of nitrosylhemoglobin. Biochemistry 2005, 44, 6223–6231. [Google Scholar] [CrossRef] [PubMed]
  128. De Simone, G.; di Masi, A.; Fattibene, P.; Ciaccio, C.; Platas-Iglesias, C.; Coletta, M.; Pesce, A.; Ascenzi, P. Oxygen-mediated oxidation of ferrous nitrosylated nitrobindins. J. Inorg. Biochem. 2021, 224, 111579. [Google Scholar] [CrossRef]
  129. Das, A.K.; Meuwly, M. Kinetic analysis and structural interpretation of competitive ligand binding for NO dioxygenation in truncated hemoglobin N. Angew. Chem. Int. Ed. 2018, 57, 3509–3513. [Google Scholar] [CrossRef]
  130. Crespo, A.; Martí, M.A.; Kalko, S.G.; Morreale, A.; Orozco, M.; Gelpi, J.L.; Luque, F.J.; Estrin, D.A. Theoretical study of the truncated hemoglobin HbN: Exploring the molecular basis of the NO detoxification mechanism. J. Am. Chem. Soc. 2005, 127, 4433–4444. [Google Scholar] [CrossRef]
  131. Bidon-Chanal, A.; Marti, M.A.; Crespo, A.; Milani, M.; Orozco, M.; Bolognesi, M.; Luque, F.J.; Estrin, D.A. Ligand-induced dynamical regulation of NO conversion in Mycobacterium tuberculosis truncated hemoglobin-N. Proteins 2006, 64, 457–464. [Google Scholar] [CrossRef]
  132. Estarellas, C.; Capece, L.; Seira, C.; Bidon-Chanal, A.; Estrin, D.A.; Luque, F.J. Structural plasticity in globins: Role of protein dynamics in defining ligand migration pathways. Adv. Protein Chem. Struct. Biol. 2016, 105, 59–80. [Google Scholar] [CrossRef]
  133. Gardner, P.R. Aconitase: Sensitive target and measure of superoxide. Methods Enzymol. 2002, 349, 9–23. [Google Scholar] [CrossRef]
  134. Ferrand, L.; Soorkia, S.; Grégoire, G.; Broquier, M.; Soep, B.; Shafizadeh, N. Bonding of heme FeIII with dioxygen: Observation and characterization of an incipient bond. Phys. Chem. Chem. Phys. 2015, 17, 25693–25699. [Google Scholar] [CrossRef]
  135. Shafizadeh, N.; Soorkia, S.; Grégoire, G.; Broquier, M.; Crestoni, M.-E.; Soep, B. Dioxygen binding to protonated heme in the gas phase, an intermediate between ferric and ferrous heme. CHEM-EUR J. 2017, 23, 13493–13500. [Google Scholar] [CrossRef]
  136. Dillinger, S.; Lang, J.; Niedner-Schatteburg, G. Cryo IR spectroscopy of [hemin]+ complexes in isolation. J. Phys. Chem. A 2017, 121, 7191–7196. [Google Scholar] [CrossRef] [PubMed]
  137. Shafizadeh, N.; Crestoni, M.E.; de la Lande, A.; Soep, B. Heme ligation in the gas phase. Int. Rev. Phys. Chem. 2021, 40, 365–404. [Google Scholar] [CrossRef]
  138. Trifunovic, L.; Dial, O.; Trif, M.; Wootton, J.R.; Abebe, R.; Yacoby, A.; Loss, D. Long-distance spin-spin coupling via floating gates. Phys. Rev. X 2012, 2, 011006. [Google Scholar] [CrossRef]
  139. Scheidt, W.; Gouterman, M. Ligands, spin state, and geometry in hemes and related metalloporphyrins. In Iron Porphyrins; Lever, A.B.P., Gray, H.B., Eds.; Addison-Wesley Publishing Company: Reading, MA, USA, 1983; Volume 1, pp. 90–139. [Google Scholar]
  140. Khudyakov, I.V.; Minaev, B.F. Molecular terms of dioxygen and nitric oxide. Physchem 2021, 1, 121–132. [Google Scholar] [CrossRef]
  141. Benesch, R.E.; Edalji, R.; Benesch, R. Reciprocal interaction of hemoglobin with oxygen and protons. The influence of allosteric polyanions. Biochemistry 1977, 16, 2594–2597. [Google Scholar] [CrossRef]
  142. Lehnert, N.; Kim, E.; Dong, H.T.; Harland, J.B.; Hunt, A.P.; Manickas, E.C.; Oakley, K.M.; Pham, J.; Reed, G.C.; Alfaro, V.S. The biologically relevant coordination chemistry of iron and nitric oxide: Electronic structure and reactivity. Chem. Rev. 2021, 121, 14682–14905. [Google Scholar] [CrossRef]
  143. Kepp, K.P. Heme: From quantum spin crossover to oxygen manager of life. Coord. Chem. Rev. 2017, 344, 363–374. [Google Scholar] [CrossRef]
  144. Nagatomo, S.; Nagai, M.; Kitagawa, T. Structural origin of cooperativity in human hemoglobin: A view from different roles of α and β subunits in the α2β2 tetramer. Biophys. Rev. 2022, 14, 483–498. [Google Scholar] [CrossRef]
  145. Chung, L.W.; Li, X.; Hirao, H.; Morokuma, K. Comparative reactivity of ferric-superoxo and ferryl-oxo species in heme and non-heme complexes. J. Am. Chem. Soc. 2011, 133, 20076–20079. [Google Scholar] [CrossRef]
  146. Bounds, P.L.; Koppenol, W.H. Peroxynitrite: A tale of two radicals. Redox Biochem. Chem. 2024, 10, 100038. [Google Scholar] [CrossRef]
  147. Ilari, A.; Bonamore, A.; Farina, A.; Johnson, K.A.; Boffi, A. The X-ray structure of ferric Escherichia coli flavohemoglobin reveals an unexpected geometry of the distal heme pocket. J. Biol. Chem. 2002, 277, 23725–23732. [Google Scholar] [CrossRef] [PubMed]
  148. El Hammi, E.; Warkentin, E.; Demmer, U.; Marzouki, N.M.; Ermler, U.; Baciou, L. Active site analysis of yeast flavohemoglobin based on its structure with a small ligand or econazole. FEBS J. 2012, 279, 4565–4575. [Google Scholar] [CrossRef] [PubMed]
  149. El Hammi, E.; Houee-Levin, C.; Rezac, J.; Levy, B.; Demachy, I.; Baciou, L.; de la Lande, A. New insights into the mechanism of electron transfer within flavohemoglobins: Tunnelling pathways, packing density, thermodynamic and kinetic analyses. Phys. Chem. Chem. Phys. 2012, 14, 13872–13880. [Google Scholar] [CrossRef]
  150. Ermler, U.; Siddiqui, R.A.; Cramm, R.; Friedrich, B. Crystal structure of the flavohemoglobin from Alcaligenes eutrophus at 1.75 Å resolution. EMBO J. 1995, 14, 6067–6077. [Google Scholar] [CrossRef]
  151. Narhi, L.O.; Fulco, A.J. Characterization of a catalytically self-sufficient 119,000-dalton cytochrome P-450 monooxygenase induced by barbiturates in Bacillus megaterium. J. Biol. Chem. 1986, 261, 7160–7169. [Google Scholar] [CrossRef]
  152. Yeom, H.; Sligar, S.G.; Li, H.; Poulos, T.L.; Fulco, A.J. The role of Thr268 in oxygen activation of cytochrome P450BM-3. Biochemistry 1995, 34, 14733–14740. [Google Scholar] [CrossRef]
  153. Mowat, C.G.; Gazur, B.; Campbell, L.P.; Chapman, S.K. Flavin-containing heme enzymes. Arch. Biochem. Biophys. 2010, 493, 37–52. [Google Scholar] [CrossRef]
  154. Haines, D.C.; Tomchick, D.R.; Machius, M.; Peterson, J.A. Pivotal role of water in the mechanism of P450BM-3. Biochemistry 2001, 40, 13456–13465. [Google Scholar] [CrossRef]
  155. Dickerson, R.E.; Geis, I. Hemoglobin: Structure, Function, Evolution, and Pathology; Benjamin-Cummings Publishing Company: Menlo Park, CA, USA, 1983; pp. 1–176. [Google Scholar]
  156. Brunori, M.; Miele, A.E. Modulation of Allosteric Control and Evolution of Hemoglobin. Biomolecules 2023, 13, 572. [Google Scholar] [CrossRef]
  157. Borisov, V.B.; Forte, E.; Giuffre, A.; Konstantinov, A.A.; Sarti, P. 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. 2009, 103, 1185–1187. [Google Scholar] [CrossRef]
  158. Keilin, D.; Hartree, E.F. Cytochrome and cytochrome oxidase. Proc. R. Soc. Lond. B-Biol. Sci. 1939, 127, 167–191. [Google Scholar] [CrossRef]
  159. Greenwood, C.; Gibson, Q.H. The reaction of reduced cytochrome c oxidase with oxygen. J. Biol. Chem. 1967, 242, 1782–1787. [Google Scholar] [CrossRef] [PubMed]
  160. Greenwood, C.; Wilson, M.T.; Brunori, M. Studies on partially reduced mammalian cytochrome oxidase. Reactions with carbon monoxide and oxygen. Biochem. J. 1974, 137, 205–215. [Google Scholar] [CrossRef] [PubMed]
  161. Gibson, Q.; Greenwood, C. Reactions of cytochrome oxidase with oxygen and carbon monoxide. Biochem. J. 1963, 86, 541–555. [Google Scholar] [CrossRef]
  162. Antonini, E.; Brunori, M.; Colosimo, A.; Greenwood, C.; Wilson, M.T. Oxygen “pulsed” cytochrome c oxidase: Functional properties and catalytic relevance. Proc. Natl. Acad. Sci. USA 1977, 74, 3128–3132. [Google Scholar] [CrossRef]
  163. Wikström, M.; Gennis, R.B.; Rich, P.R. Structures of the intermediates in the catalytic cycle of mitochondrial cytochrome c oxidase. Biochim. Biophys. Acta (BBA)-Bioenerg. 2023, 1864, 148933. [Google Scholar] [CrossRef]
  164. Noodleman, L.; Götz, A.W.; Han Du, W.-G.; Hunsicker-Wang, L. Reaction pathways, proton transfer, and proton pumping in ba3 class cytochrome c oxidase: Perspectives from DFT quantum chemistry and molecular dynamics. Front. Chem. 2023, 11, 1186022. [Google Scholar] [CrossRef]
  165. Granger, D.L.; Lehninger, A.L. Sites of inhibition of mitochondrial electron transport in macrophage-injured neoplastic cells. J. Cell Biol. 1982, 95, 527–535. [Google Scholar] [CrossRef]
  166. Giuffrè, A.; Barone, M.C.; Mastronicola, D.; D’Itri, E.; Sarti, P.; Brunori, M. Reaction of nitric oxide with the turnover intermediates of cytochrome c oxidase:  reaction pathway and functional effects. Biochemistry 2000, 39, 15446–15453. [Google Scholar] [CrossRef]
  167. Sarti, P.; Giuffrè, A.; Barone, M.C.; Forte, E.; Mastronicola, D.; Brunori, M. Nitric oxide and cytochrome oxidase: Reaction mechanisms from the enzyme to the cell. Free Radic. Biol. Med. 2003, 34, 509–520. [Google Scholar] [CrossRef]
  168. Brunori, M.; Giuffrè, A.; Sarti, P. Cytochrome c oxidase, ligands and electrons. J. Inorg. Biochem. 2005, 99, 324–336. [Google Scholar] [CrossRef] [PubMed]
  169. Sarti, P.; Forte, E.; Mastronicola, D.; Giuffrè, A.; Arese, M. Cytochrome c oxidase and nitric oxide in action: Molecular mechanisms and pathophysiological implications. Biochim. Biochim. Biophys. Acta (BBA)-Bioenerg. 2012, 1817, 610–619. [Google Scholar] [CrossRef] [PubMed]
  170. Borisov, V.B.; Forte, E. Bioenergetics and reactive nitrogen species in bacteria. Int. J. Mol. Sci. 2022, 23, 7321. [Google Scholar] [CrossRef] [PubMed]
  171. Stevens, T.H.; Brudvig, G.W.; Bocian, D.F.; Chan, S.I. Structure of cytochrome a3-Cua3 couple in cytochrome c oxidase as revealed by nitric oxide binding studies. Proc. Natl. Acad. Sci. USA 1979, 76, 3320–3324. [Google Scholar] [CrossRef]
  172. Giuffrè, A.; Stubauer, G.; Sarti, P.; Brunori, M.; Zumft, W.G.; Buse, G.; Soulimane, T. The heme-copper oxidases of Thermus thermophilus catalyze the reduction of nitric oxide: Evolutionary implications. Proc. Natl. Acad. Sci. USA 1999, 96, 14718–14723. [Google Scholar] [CrossRef]
  173. Brunori, M. Nitric oxide moves myoglobin center stage. Trends Biochem. Sci. 2001, 26, 209–210. [Google Scholar] [CrossRef]
  174. Torres, J.; Sharpe, M.A.; Rosquist, A.; Cooper, C.E.; Wilson, M.T. Cytochrome c oxidase rapidly metabolises nitric oxide to nitrite. FEBS Lett. 2000, 475, 263–266. [Google Scholar] [CrossRef]
  175. Giuffrè, A.; Stubauer, G.; Brunori, M.; Sarti, P.; Torres, J.; Wilson, M.T. Chloride bound to oxidized cytochrome c oxidase controls the reaction with nitric oxide. J. Biol. Chem. 1998, 273, 32475–32478. [Google Scholar] [CrossRef]
  176. Pearce, L.L.; Kanai, A.J.; Birder, L.A.; Pitt, B.R.; Peterson, J. The catabolic fate of nitric oxide: The nitric oxide oxidase and peroxynitrite reductase activities of cytochrome oxidase. J. Biol. Chem. 2002, 277, 13556–13562. [Google Scholar] [CrossRef]
  177. Collman, J.P.; Dey, A.; Decreau, R.A.; Yang, Y.; Hosseini, A.; Solomon, E.I.; Eberspacher, T.A. Interaction of nitric oxide with a functional model of cytochrome c oxidase. Proc. Natl. Acad. Sci. USA 2008, 105, 9892–9896. [Google Scholar] [CrossRef] [PubMed]
  178. Pearce, L.L.; Pitt, B.R.; Peterson, J. The peroxynitrite reductase activity of cytochrome c oxidase involves a two-electron redox reaction at the heme a3-CuB site. J. Biol. Chem. 1999, 274, 35763–35767. [Google Scholar] [CrossRef] [PubMed]
  179. Peterson, J.; Kanai, A.J.; Pearce, L.L. A mitochondrial role for catabolism of nitric oxide in cardiomyocytes not involving oxymyoglobin. Am. J. Physiol. Heart Circ. Physiol. 2004, 286, H55–H58. [Google Scholar] [CrossRef] [PubMed]
  180. Wikström, M.K.F. Proton pump coupled to cytochrome c oxidase in mitochondria. Nature 1977, 266, 271–273. [Google Scholar] [CrossRef]
  181. Wilson, M.T.; Peterson, J.; Antonini, E.; Brunori, M.; Colosimo, A.; Wyman, J. A plausible two-state model for cytochrome c oxidase. Proc. Natl. Acad. Sci. USA 1981, 78, 7115–7118. [Google Scholar] [CrossRef]
  182. Babcock, G.T.; Varotsis, C. Discrete steps in dioxygen activation—The cytochrome oxidase/O2 reaction. J. Bioenerg. Biomembr. 1993, 25, 71–80. [Google Scholar] [CrossRef]
  183. Chance, B.; Saronio, C.; Leigh, J.S., Jr. Functional intermediates in reaction of cytochrome oxidase with oxygen. Proc. Natl. Acad. Sci. USA 1975, 72, 1635–1640. [Google Scholar] [CrossRef]
  184. Yoshikawa, S.; Shimada, A. Reaction mechanism of cytochrome c oxidase. Chem. Rev. 2015, 115, 1936–1989. [Google Scholar] [CrossRef]
  185. Popović, D.M.; Stuchebrukhov, A.A. Coupled electron and proton transfer reactions during the O→E transition in bovine cytochrome c oxidase. Biochim. Biophys. Acta (BBA)-Bioenerg. 2012, 1817, 506–517. [Google Scholar] [CrossRef]
  186. Kaila, V.R.I.; Oksanen, E.; Goldman, A.; Bloch, D.A.; Verkhovsky, M.I.; Sundholm, D.; Wikström, M. A combined quantum chemical and crystallographic study on the oxidized binuclear center of cytochrome c oxidase. Biochim. Biophys. Acta (BBA)-Bioenerg. 2011, 1807, 769–778. [Google Scholar] [CrossRef]
  187. Hammes-Schiffer, S. Proton-coupled electron transfer: Moving together and charging forward. J. Am. Chem. Soc. 2015, 137, 8860–8871. [Google Scholar] [CrossRef] [PubMed]
  188. Bloch, D.; Belevich, I.; Jasaitis, A.; Ribacka, C.; Puustinen, A.; Verkhovsky, M.I.; Wikström, M. The catalytic cycle of cytochrome c oxidase is not the sum of its two halves. Proc. Natl. Acad. Sci. USA 2004, 101, 529–533. [Google Scholar] [CrossRef] [PubMed]
  189. Michel, H. Cytochrome c oxidase:  catalytic cycle and mechanisms of proton pumping-a discussion. Biochemistry 1999, 38, 15129–15140. [Google Scholar] [CrossRef] [PubMed]
  190. Koppenol, W.H.; Stanbury, D.M.; Bounds, P.L. Electrode potentials of partially reduced oxygen species, from dioxygen to water. Free Radic. Biol. Med. 2010, 49, 317–322. [Google Scholar] [CrossRef]
  191. Jose, A.; Schaefer, A.W.; Roveda, A.C.; Transue, W.J.; Choi, S.K.; Ding, Z.; Gennis, R.B.; Solomon, E.I. The three-spin intermediate at the O–O cleavage and proton-pumping junction in heme–Cu oxidases. Science 2021, 373, 1225–1229. [Google Scholar] [CrossRef]
  192. Proshlyakov, D.A.; Pressler, M.A.; DeMaso, C.; Leykam, J.F.; DeWitt, D.L.; Babcock, G.T. Oxygen activation and reduction in respiration: Involvement of redox-active tyrosine 244. Science 2000, 290, 1588–1591. [Google Scholar] [CrossRef]
  193. Chance, B.; Saronio, C.; Leigh, J.S., Jr. Functional intermediates in the reaction of membrane-bound cytochrome oxidase with oxygen. J. Biol. Chem. 1975, 250, 9226–9237. [Google Scholar] [CrossRef]
  194. Rousseau, D.L.; Ishigami, I.; Yeh, S.-R. Structural and functional mechanisms of cytochrome c oxidase. J. Inorg. Biochem. 2025, 262, 112730. [Google Scholar] [CrossRef]
  195. Wikström, M.; Sharma, V. Cytochrome c Oxidase—Remaining Questions About the Catalytic Mechanism. In Oxygen Production and Reduction in Artificial and Natural Systems; Barber, J., Ruban, A.V., Nixon, P.J., Eds.; World Scientific: Singapore, 2019; Volume 2019, pp. 135–145. [Google Scholar]
  196. Blomberg, M.R. Activation of O2 and NO in heme-copper oxidases–mechanistic insights from computational modelling. Chem. Soc. Rev. 2020, 49, 7301–7330. [Google Scholar] [CrossRef]
  197. Petersen, L.C.; Nicholls, P.; Degn, H. The effect of energization on the apparent Michaelis–Menten constant for oxygen in mitochondrial respiration. Biochem. J. 1974, 142, 247–252. [Google Scholar] [CrossRef]
  198. Krab, K.; Kempe, H.; Wikström, M. Explaining the enigmatic KM for oxygen in cytochrome c oxidase: A kinetic model. Biochim. Biophys. Acta (BBA)-Bioenerg. 2011, 1807, 348–358. [Google Scholar] [CrossRef]
  199. Williams, D.H.; Westwell, M.S. Aspects of weak interactions. Chem. Soc. Rev. 1998, 27, 57–64. [Google Scholar] [CrossRef]
  200. Falahati, K.; Tamura, H.; Burghardt, I.; Huix-Rotllant, M. Ultrafast carbon monoxide photolysis and heme spin-crossover in myoglobin via nonadiabatic quantum dynamics. Nat. Commun. 2018, 9, 4502. [Google Scholar] [CrossRef] [PubMed]
  201. Saura, P.; Riepl, D.; Frey, D.M.; Wikström, M.; Kaila, V.R.I. Electric fields control water-gated proton transfer in cytochrome c oxidase. Proc. Natl. Acad. Sci. USA 2022, 119, e2207761119. [Google Scholar] [CrossRef] [PubMed]
  202. Seki, S.; Kurumaji, T.; Ishiwata, S.; Matsui, H.; Murakawa, H.; Tokunaga, Y.; Kaneko, Y.; Hasegawa, T.; Tokura, Y. Cupric chloride CuCl2 as an S = 1/2 chain multiferroic. Phys. Rev. B 2010, 82, 064424. [Google Scholar] [CrossRef]
  203. Jones, S.P.P.; Gaw, S.M.; Doig, K.I.; Prabhakaran, D.; Hétroy Wheeler, E.M.; Boothroyd, A.T.; Lloyd-Hughes, J. High-temperature electromagnons in the magnetically induced multiferroic cupric oxide driven by intersublattice exchange. Nat. Commun. 2014, 5, 3787. [Google Scholar] [CrossRef]
  204. Zakharov, I.; Minaev, B. DFT calculations of the intermediate and transition state in the oxidation of NO by oxygen in the gas phase. Theor. Exp. Chem. 2011, 47, 93–100. [Google Scholar] [CrossRef]
  205. Yoshikawa, S.; Shinzawa-Itoh, K.; Nakashima, R.; Yaono, R.; Yamashita, E.; Inoue, N.; Yao, M.; Fei, M.J.; Libeu, C.P.; Mizushima, T.; et al. Redox-coupled crystal structural changes in bovine heart cytochrome c oxidase. Science 1998, 280, 1723–1729. [Google Scholar] [CrossRef]
  206. Kolbe, F.; Safarian, S.; Piórek, Ż.; Welsch, S.; Müller, H.; Michel, H. Cryo-EM structures of intermediates suggest an alternative catalytic reaction cycle for cytochrome c oxidase. Nat. Commun. 2021, 12, 6903. [Google Scholar] [CrossRef]
  207. van der Avoird, A.; Brocks, G. The O2–O2 dimer: Magnetic coupling and spectrum. J. Chem. Phys. 1987, 87, 5346–5360. [Google Scholar] [CrossRef]
  208. Minaev, B. Electronic mechanisms of activation of molecular oxygen. Russian Chem. Rev. 2007, 76, 988–1010. [Google Scholar] [CrossRef]
  209. Fitch, C.A.; Platzer, G.; Okon, M.; Garcia-Moreno E., B.; McIntosh, L.P. Arginine: Its pKa value revisited. Protein Sci. 2015, 24, 752–761. [Google Scholar] [CrossRef] [PubMed]
  210. Ferguson-Miller, S.; Hiser, C.; Liu, J. Gating and regulation of the cytochrome c oxidase proton pump. Biochim. Biophys. Acta (BBA)-Bioenerg. 2012, 1817, 489–494. [Google Scholar] [CrossRef]
  211. Olkhova, E.; Hutter, M.C.; Lill, M.A.; Helms, V.; Michel, H. Dynamic water networks in cytochrome c oxidase from Paracoccus denitrificans investigated by molecular dynamics simulations. Biophys. J. 2004, 86, 1873–1889. [Google Scholar] [CrossRef] [PubMed]
  212. Gorriz, R.F.; Volkenandt, S.; Imhof, P. Protonation-state dependence of hydration and interactions in the two proton-conducting channels of cytochrome c oxidase. Int. J. Mol. Sci. 2023, 24, 10464. [Google Scholar] [CrossRef]
  213. Krab, K.; Wikström, M. Principles of coupling between electron transfer and proton translocation with special reference to proton-translocation mechanisms in cytochrome oxidase. Biochim. Biophys. Acta (BBA)-Rev. Bioenerg. 1987, 895, 25–39. [Google Scholar] [CrossRef]
  214. Malmström, B.G. Redox loops and proton pumps. FEBS Lett. 1988, 231, 268–269. [Google Scholar] [CrossRef]
  215. Chan, S.I.; Li, P.M. Cytochrome c oxidase: Understanding nature’s design of a proton pump. Biochemistry 1990, 29, 1–12. [Google Scholar] [CrossRef]
  216. Kim, Y.C.; Hummer, G. Proton-pumping mechanism of cytochrome c oxidase: A kinetic master-equation approach. Biochim. Biophys. Acta (BBA)-Rev. Bioenerg. 2012, 1817, 526–536. [Google Scholar] [CrossRef]
  217. Shimada, A.; Hara, F.; Shinzawa-Itoh, K.; Kanehisa, N.; Yamashita, E.; Muramoto, K.; Tsukihara, T.; Yoshikawa, S. Critical roles of the CuB site in efficient proton pumping as revealed by crystal structures of mammalian cytochrome c oxidase catalytic intermediates. J. Biol. Chem. 2021, 297, 100967. [Google Scholar] [CrossRef]
  218. Shimokata, K.; Katayama, Y.; Murayama, H.; Suematsu, M.; Tsukihara, T.; Muramoto, K.; Aoyama, H.; Yoshikawa, S.; Shimada, H. The proton pumping pathway of bovine heart cytochrome c oxidase. Proc. Natl. Acad. Sci. USA 2007, 104, 4200–4205. [Google Scholar] [CrossRef]
  219. Brzezinski, P.; Johansson, A.-L. Variable proton-pumping stoichiometry in structural variants of cytochrome c oxidase. Biochim. Biophys. Acta (BBA)-Rev. Bioenerg. 2010, 1797, 710–723. [Google Scholar] [CrossRef]
  220. Sugitani, R.; Medvedev, E.S.; Stuchebrukhov, A.A. Theoretical and computational analysis of the membrane potential generated by cytochrome c oxidase upon single electron injection into the enzyme. Biochim. Biophys. Acta (BBA)-Rev. Bioenerg. 2008, 1777, 1129–1139. [Google Scholar] [CrossRef] [PubMed]
  221. Hol, W.G. The role of the α-helix dipole in protein function and structure. Prog. Biophys. Mol. Biol. 1985, 45, 149–195. [Google Scholar] [CrossRef] [PubMed]
  222. Zorova, L.D.; Popkov, V.A.; Plotnikov, E.Y.; Silachev, D.N.; Pevzner, I.B.; Jankauskas, S.S.; Babenko, V.A.; Zorov, S.D.; Balakireva, A.V.; Juhaszova, M.; et al. Mitochondrial membrane potential. Anal. Biochem. 2018, 552, 50–59. [Google Scholar] [CrossRef] [PubMed]
  223. Agmon, N. Tetrahedral displacement:  The molecular mechanism behind the Debye relaxation in water. J. Phys. Chem. 1996, 100, 1072–1080. [Google Scholar] [CrossRef]
  224. de la Lande, A.; Babcock, N.S.; Rezác, J.; Sanders, B.C.; Salahub, D.R. Surface residues dynamically organize water bridges to enhance electron transfer between proteins. Proc. Natl. Acad. Sci. USA 2010, 107, 11799–11804. [Google Scholar] [CrossRef]
  225. Rich, P.R.; Meunier, B.; Mitchell, R.; John Moody, A. Coupling of charge and proton movement in cytochrome c oxidase. Biochim. Biophys. Acta (BBA)-Rev. Bioenerg. 1996, 1275, 91–95. [Google Scholar] [CrossRef]
  226. Smith, M.; Li, Z.; Landry, L.; Merz, K.M., Jr.; Li, P. Consequences of overfitting the van der Waals radii of ions. J. Chem. Theory Comput. 2023, 19, 2064–2074. [Google Scholar] [CrossRef]
  227. Tsukihara, T.; Aoyama, H.; Yamashita, E.; Tomizaki, T.; Yamaguchi, H.; Shinzawa-Itoh, K.; Nakashima, R.; Yaono, R.; Yoshikawa, S. Structures of metal sites of oxidized bovine heart cytochrome c oxidase at 2.8 Å. Science 1995, 269, 1069–1074. [Google Scholar] [CrossRef]
  228. Tsukihara, T.; Aoyama, H.; Yamashita, E.; Tomizaki, T.; Yamaguchi, H.; Shinzawa-Itoh, K.; Nakashima, R.; Yaono, R.; Yoshikawa, S. The whole structure of the 13-subunit oxidized cytochrome c oxidase at 2.8 Å. Science 1996, 272, 1136–1144. [Google Scholar] [CrossRef] [PubMed]
  229. Iwata, S.; Ostermeier, C.; Ludwig, B.; Michel, H. Structure at 2.8 Å resolution of cytochrome c oxidase from Paracoccus denitrificans. Nature 1995, 376, 660–669. [Google Scholar] [CrossRef] [PubMed]
  230. Verkhovsky, M.I.; Morgan, J.E.; Wikström, M. Oxygen binding and activation: Early steps in the reaction of oxygen with cytochrome c oxidase. Biochemistry 1994, 33, 3079–3086. [Google Scholar] [CrossRef] [PubMed]
  231. Hill, B.C.; Greenwood, C. The reaction of fully reduced cytochrome c oxidase with oxygen studied by flow-flash spectrophotometry at room temperature. Evidence for new pathways of electron transfer. Biochem. J. 1984, 218, 913–921. [Google Scholar] [CrossRef]
  232. Ishigami, I.; Russi, S.; Cohen, A.; Yeh, S.-R.; Rousseau, D.L. Temperature-dependent structural transition following X-ray-induced metal center reduction in oxidized cytochrome c oxidase. J. Biol. Chem. 2022, 298, 101799. [Google Scholar] [CrossRef]
  233. Hill, B.C.; Greenwood, C. Kinetic evidence for the re-definition of electron transfer pathways from cytochrome c to O2 within cytochrome oxidase. FEBS Lett. 1984, 166, 362–366. [Google Scholar] [CrossRef]
  234. Antonini, E.; Brunori, M.; Greenwood, C.; Malmström, B.G. Catalytic mechanism of cytochrome oxidase. Nature 1970, 228, 936–937. [Google Scholar] [CrossRef]
  235. Verkhovsky, M.I.; Jasaitis, A.; Verkhovskaya, M.L.; Morgan, J.E.; Wikström, M. Proton translocation by cytochrome c oxidase. Nature 1999, 400, 480–483. [Google Scholar] [CrossRef]
  236. Brudvig, G.W.; Stevens, T.H.; Morse, R.H.; Chan, S.I. Conformations of oxidized cytochrome c oxidase. Biochemistry 1981, 20, 3912–3921. [Google Scholar] [CrossRef]
  237. Mills, D.A.; Geren, L.; Hiser, C.; Schmidt, B.; Durham, B.; Millett, F.; Ferguson-Miller, S. An arginine to lysine mutation in the vicinity of the heme propionates affects the redox potentials of the hemes and associated electron and proton transfer in cytochrome c oxidase. Biochemistry 2005, 44, 10457–10465. [Google Scholar] [CrossRef]
  238. Puustinen, A.; Wikström, M. Proton exit from the heme-copper oxidase of Escherichia coli. Proc. Natl. Acad. Sci. USA 1999, 96, 35–37. [Google Scholar] [CrossRef]
  239. Qian, J.; Mills, D.A.; Geren, L.; Wang, K.; Hoganson, C.W.; Schmidt, B.; Hiser, C.; Babcock, G.T.; Durham, B.; Millett, F.; et al. Role of the conserved arginine pair in proton and electron transfer in cytochrome c oxidase. Biochemistry 2004, 43, 5748–5756. [Google Scholar] [CrossRef]
  240. Papa, S.; Capitanio, N.; Capitanio, G. A cooperative model for proton pumping in cytochrome c oxidase. Biochim. Biophys. Acta (BBA)-Bioenerg. 2004, 1655, 353–364. [Google Scholar] [CrossRef]
  241. Brändén, G.; Brändén, M.; Schmidt, B.; Mills, D.A.; Ferguson-Miller, S.; Brzezinski, P. The protonation state of a heme propionate controls electron transfer in cytochrome c oxidase. Biochemistry 2005, 44, 10466–10474. [Google Scholar] [CrossRef] [PubMed]
  242. Pearce, L.L.; Lopez Manzano, E.; Martinez-Bosch, S.; Peterson, J. Antagonism of nitric oxide toward the inhibition of cytochrome c oxidase by carbon monoxide and cyanide. Chem. Res. Toxicol. 2008, 21, 2073–2081. [Google Scholar] [CrossRef] [PubMed]
  243. Frawley, K.L.; Cronican, A.A.; Pearce, L.L.; Peterson, J. Sulfide toxicity and its modulation by nitric oxide in bovine pulmonary artery endothelial cells. Chem. Res. Toxicol. 2017, 30, 2100–2109. [Google Scholar] [CrossRef] [PubMed]
  244. Ball, E.G.; Strittmatter, C.F.; Cooper, O. The reaction of cytochrome oxidase with carbon monoxide. J. Biol. Chem. 1951, 193, 635–647. [Google Scholar] [CrossRef]
  245. Keilin, D. Influence of carbon monoxide and light on indophenol oxidase of yeast cells. Nature 1927, 119, 670–671. [Google Scholar] [CrossRef]
  246. Brzezinski, P.; Malmström, B.G. The reduction of cytochrome c oxidase by carbon monoxide. FEBS Lett. 1985, 187, 111–114. [Google Scholar] [CrossRef]
  247. Day, E.P.; Kent, T.A.; Lindahl, P.A.; Münck, E.; Orme-Johnson, W.H.; Roder, H.; Roy, A. SQUID measurement of metalloprotein magnetization. New methods applied to the nitrogenase proteins. Biophys. J. 1987, 52, 837–853. [Google Scholar] [CrossRef]
  248. Peterson, J.; Day, E.P.; Pearce, L.L.; Wilson, M.T. Measurement of the spin concentration of metalloprotein samples from saturation-magnetization data with particular reference to cytochrome c oxidase. Biochem. J. 1995, 305, 871–878. [Google Scholar] [CrossRef]
  249. Van Gelder, B.F.; Beinert, H. Studies of the heme components of cytochrome c oxidase by EPR spectroscopy. Biochim. Biophys. Acta (BBA)-Bioenerg. 1969, 189, 1–24. [Google Scholar] [CrossRef]
  250. Van Gelder, B.F.; Slater, E.C. Titration of cytochrome c oxidase with NADH and phenazine methosulphate. Biochim. Biophys. Acta (BBA)-Spec. Sect. Enzymol. Subj. 1963, 73, 663–665. [Google Scholar] [CrossRef]
  251. Powers, L.; Chance, B.; Ching, Y.; Angiolillo, P. Structural features and the reaction mechanism of cytochrome oxidase: Iron and copper X-ray absorption fine structure. Biophys. J. 1981, 34, 465–498. [Google Scholar] [CrossRef] [PubMed]
  252. Yu, C.; Yu, L.; King, T.E. Studies on cytochrome oxidase. Interactions of the cytochrome oxidase protein with phospholipids and cytochrome c. J. Biol. Chem. 1975, 250, 1383–1392. [Google Scholar] [CrossRef] [PubMed]
  253. Bruce, R.A.; Lovejoy, F.W., Jr.; Yu, P.N.G.; Pearson, R.; McDowell, M. Further observations on the pathological physiology of chronic pulmonary granulomatosis associated with beryllium workers. Am. Rev. Tuberc. 1950, 62, 29–44. [Google Scholar] [CrossRef]
  254. Messer, R.L.W.; Doeller, J.E.; Kraus, D.W.; Lucas, L.C. An investigation of fibroblast mitochondria enzyme activity and respiration in response to metallic ions released from dental alloys. J. Biomed. Mater. Res. 2000, 50, 598–604. [Google Scholar] [CrossRef]
  255. Fridovich, I.; Handler, P. Detection of free radicals generated during enzymic oxidations by the initiation of sulfite oxidation. J. Biol. Chem. 1961, 236, 1836–1840. [Google Scholar] [CrossRef]
  256. Shepherd, M.; Giordano, D.; Verde, C.; Poole, R.K. The evolution of nitric oxide function: From reactivity in the prebiotic earth to examples of biological roles and therapeutic applications. Antioxidants 2022, 11, 1222. [Google Scholar] [CrossRef]
  257. Brunori, M.; Giuffre, A.; Nienhaus, K.; Nienhaus, G.U.; Scandurra, F.M.; Vallone, B. Neuroglobin, nitric oxide, and oxygen: Functional pathways and conformational changes. Proc. Natl. Acad. Sci. USA 2005, 102, 8483–8488. [Google Scholar] [CrossRef]
  258. Khan, A.A.; Mao, X.O.; Banwait, S.; Jin, K.; Greenberg, D.A. Neuroglobin attenuates beta-amyloid neurotoxicity in vitro and transgenic Alzheimer phenotype in vivo. Proc. Natl. Acad. Sci. USA 2007, 104, 19114–19119. [Google Scholar] [CrossRef] [PubMed]
  259. Kawada, N.; Kristensen, D.B.; Asahina, K.; Nakatani, K.; Minamiyama, Y.; Seki, S.; Yoshizato, K. Characterization of a stellate cell activation-associated protein (STAP) with peroxidase activity found in rat hepatic stellate cells. J. Biol. Chem. 2001, 276, 25318–25323. [Google Scholar] [CrossRef]
  260. Nakatani, K.; Okuyama, H.; Shimahara, Y.; Saeki, S.; Kim, D.H.; Nakajima, Y.; Seki, S.; Kawada, N.; Yoshizato, K. Cytoglobin/STAP, its unique localization in splanchnic fibroblast-like cells and function in organ fibrogenesis. Lab. Investig. 2004, 84, 91–101. [Google Scholar] [CrossRef] [PubMed]
  261. Jourd’heuil, F.L.; Xu, H.; Reilly, T.; McKellar, K.; El Alaoui, C.; Steppich, J.; Liu, Y.F.; Zhao, W.; Ginnan, R.; Conti, D.; et al. The hemoglobin homolog cytoglobin in smooth muscle inhibits apoptosis and regulates vascular remodeling. Arterioscler. Thromb. Vasc. Biol. 2017, 37, 1944–1955. [Google Scholar] [CrossRef] [PubMed]
  262. Liu, X.; El-Mahdy, M.A.; Boslett, J.; Varadharaj, S.; Hemann, C.; Abdelghany, T.M.; Ismail, R.S.; Little, S.C.; Zhou, D.; Thuy, L.T.; et al. Cytoglobin regulates blood pressure and vascular tone through nitric oxide metabolism in the vascular wall. Nat. Commun. 2017, 8, 14807. [Google Scholar] [CrossRef]
  263. Ukeri, J.; Wilson, M.T.; Reeder, B.J. Modulating nitric oxide dioxygenase and nitrite reductase of cytoglobin through point mutations. Antioxidants 2022, 11, 1816. [Google Scholar] [CrossRef]
  264. Straub, A.C.; Lohman, A.W.; Billaud, M.; Johnstone, S.R.; Dwyer, S.T.; Lee, M.Y.; Bortz, P.S.; Best, A.K.; Columbus, L.; Gaston, B.; et al. Endothelial cell expression of haemoglobin α regulates nitric oxide signalling. Nature 2012, 491, 473–477. [Google Scholar] [CrossRef]
  265. Flögel, U.; Merx, M.W.; Gödecke, A.; Decking, U.K.; Schrader, J. Myoglobin: A scavenger of bioactive NO. Proc. Natl. Acad. Sci. USA 2001, 98, 735–740. [Google Scholar] [CrossRef]
Figure 1. Proposed ferric NOD catalytic cycle. O2 enters and traverses a long hydrophobic tunnel (step 1). O2 displaces the distal leucine E11 residue and binds to the ferric heme; the CD loop furls and expels captured nitrate; the motion opens a short tunnel for NO entry and relaxes the reductase domain (step 2) and energizes the heme-F-helix for the O2-sensing ET switch trigger motion. NO enters and traverses the short tunnel (step 3). NADH binds and reduces FAD (step 4). NO interacts with ferric O2 and induces SC in the ferric heme, causing saddling of the heme and F-helix motion, and the lysine (F7) and carboxy tail ET switch structure is thrown ON (step 5). An electron migrates from FADH2 or FADH to the ferric O2 complex (step 6). NO reacts with the ferric O2 to form a peroxynitrite intermediate, which isomerizes to nitrate (step 7). Nitrate disrupts the hydrophobic reaction pocket structure and unfurls the spring-like CD-loop, which captures the nitrate in an anion hole (step 8). Steps 2 and 8 involve large motions driven by unfurling and furling of the CD-loop that affect other steps as indicated.
Figure 1. Proposed ferric NOD catalytic cycle. O2 enters and traverses a long hydrophobic tunnel (step 1). O2 displaces the distal leucine E11 residue and binds to the ferric heme; the CD loop furls and expels captured nitrate; the motion opens a short tunnel for NO entry and relaxes the reductase domain (step 2) and energizes the heme-F-helix for the O2-sensing ET switch trigger motion. NO enters and traverses the short tunnel (step 3). NADH binds and reduces FAD (step 4). NO interacts with ferric O2 and induces SC in the ferric heme, causing saddling of the heme and F-helix motion, and the lysine (F7) and carboxy tail ET switch structure is thrown ON (step 5). An electron migrates from FADH2 or FADH to the ferric O2 complex (step 6). NO reacts with the ferric O2 to form a peroxynitrite intermediate, which isomerizes to nitrate (step 7). Nitrate disrupts the hydrophobic reaction pocket structure and unfurls the spring-like CD-loop, which captures the nitrate in an anion hole (step 8). Steps 2 and 8 involve large motions driven by unfurling and furling of the CD-loop that affect other steps as indicated.
Biophysica 05 00041 g001
Figure 2. NOD catalytic cycle clock. Steps 1 to 8 in the catalytic cycle are apportioned according to measured rate constants and approximate fractional times in steady-state turnover in clock-like diagrams with an equivalent circumference distance and time relation for both the 20 °C and 37 °C cycles.
Figure 2. NOD catalytic cycle clock. Steps 1 to 8 in the catalytic cycle are apportioned according to measured rate constants and approximate fractional times in steady-state turnover in clock-like diagrams with an equivalent circumference distance and time relation for both the 20 °C and 37 °C cycles.
Biophysica 05 00041 g002
Figure 3. Proposed ferric heme spintronics with O2 and NO SOCs. Antiferromagnetic (AFM) and ferromagnetic (FM) SOC of the ferric heme 3d orbitals with 2π* O2 and 2π*, 2π NO orbitals (grey arrows). Low spin (s = 1/2) to intermediate spin (s = 3/2) to high spin- (s = 5/2) crossover (SC) induced in the ferric heme with O2 and NO SOC (thick green arrows). Enhanced electron density current in ferromagnetic and antiferromagnetic SOC (magenta arrows).
Figure 3. Proposed ferric heme spintronics with O2 and NO SOCs. Antiferromagnetic (AFM) and ferromagnetic (FM) SOC of the ferric heme 3d orbitals with 2π* O2 and 2π*, 2π NO orbitals (grey arrows). Low spin (s = 1/2) to intermediate spin (s = 3/2) to high spin- (s = 5/2) crossover (SC) induced in the ferric heme with O2 and NO SOC (thick green arrows). Enhanced electron density current in ferromagnetic and antiferromagnetic SOC (magenta arrows).
Biophysica 05 00041 g003
Figure 8. Proposed proton-conducting pump influx tunnel in the bovine CcO structure. (top) Matrix view of 53 Å long tunnel from bound surface water (WS) at Pro14 connecting to the putative proton donor water 1 with tunnel lining amino acids and cytochrome a3 in the bovine CcO structure (PDB ID: 2ZXW) shown. (bottom) Rotated side view of the polar water tunnel showing the pathway from the surface water (WS) to the proton donor water (1) through the tunnel structure with the helix dipole shown following along ~30 Å of the helix axis at ~20° to the tunnel (magenta arrow, pointed from the positive pole toward the negative pole). Measured distances are internuclear. Water molecule O-atoms (aqua) shown at ~100% of the rotational radii or ~1. 4 Å, and the hypothetical missing tunnel chain waters represented by the 12 equidistant space-filling O-atoms (aqua). Dotted balls (black on aqua) represent additional waters found in the corresponding positions in the bovine and P. denitrificans structures (PDB IDs: 3ASO, 3X2Q, 7AU6, and 7ATE).
Figure 8. Proposed proton-conducting pump influx tunnel in the bovine CcO structure. (top) Matrix view of 53 Å long tunnel from bound surface water (WS) at Pro14 connecting to the putative proton donor water 1 with tunnel lining amino acids and cytochrome a3 in the bovine CcO structure (PDB ID: 2ZXW) shown. (bottom) Rotated side view of the polar water tunnel showing the pathway from the surface water (WS) to the proton donor water (1) through the tunnel structure with the helix dipole shown following along ~30 Å of the helix axis at ~20° to the tunnel (magenta arrow, pointed from the positive pole toward the negative pole). Measured distances are internuclear. Water molecule O-atoms (aqua) shown at ~100% of the rotational radii or ~1. 4 Å, and the hypothetical missing tunnel chain waters represented by the 12 equidistant space-filling O-atoms (aqua). Dotted balls (black on aqua) represent additional waters found in the corresponding positions in the bovine and P. denitrificans structures (PDB IDs: 3ASO, 3X2Q, 7AU6, and 7ATE).
Biophysica 05 00041 g008
Figure 9. Proposed ET and PT coupled pathways in CcO. Propionate bridging water molecules (WB), the proximal proton donor water (number 1), and all relevant water molecule O-atoms (aqua) shown at ~30% of their van der Waals radii in space-filling models. Proton pathway water/hydronium ions are numbered 1–4. Magnesium (green), cupric ions (brown), cysteine sulfurs (yellow), iron atoms (orangish-brown), and the peroxide O-atoms (red) in position represent the O2 molecule shown with space-filling at ~50% of their van der Waals radii. Positions and directions of proton (H+) (red arrows) and electron (e) (black arrows) migration pathways and electrostatic field dipoles (magenta arrows, pointed from the negative pole towards the positive pole) are diagrammed along specific atoms, bonds, and water molecules. The model is based upon the CcO structure coordinates (PDB ID: 2ZXW) with the proton pathway water number 3 added based upon the CcO structure models PDB IDs 5IY5 and 2Y69. Respective Mg2+, Fe3+, Cu2+, and water O-atom van der Waals radii values of 1.12 Å, 1.38 Å, 1.50 Å, and 1.4 Å [226] were applied to the correction of significantly oversized Mg and Fe ions, water O-atoms, and a relatively undersized cupric ion generated by the Jmol 14.2.15 program (Free Software Foundation, Inc., Boston, MA, USA).
Figure 9. Proposed ET and PT coupled pathways in CcO. Propionate bridging water molecules (WB), the proximal proton donor water (number 1), and all relevant water molecule O-atoms (aqua) shown at ~30% of their van der Waals radii in space-filling models. Proton pathway water/hydronium ions are numbered 1–4. Magnesium (green), cupric ions (brown), cysteine sulfurs (yellow), iron atoms (orangish-brown), and the peroxide O-atoms (red) in position represent the O2 molecule shown with space-filling at ~50% of their van der Waals radii. Positions and directions of proton (H+) (red arrows) and electron (e) (black arrows) migration pathways and electrostatic field dipoles (magenta arrows, pointed from the negative pole towards the positive pole) are diagrammed along specific atoms, bonds, and water molecules. The model is based upon the CcO structure coordinates (PDB ID: 2ZXW) with the proton pathway water number 3 added based upon the CcO structure models PDB IDs 5IY5 and 2Y69. Respective Mg2+, Fe3+, Cu2+, and water O-atom van der Waals radii values of 1.12 Å, 1.38 Å, 1.50 Å, and 1.4 Å [226] were applied to the correction of significantly oversized Mg and Fe ions, water O-atoms, and a relatively undersized cupric ion generated by the Jmol 14.2.15 program (Free Software Foundation, Inc., Boston, MA, USA).
Biophysica 05 00041 g009
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gardner, P.R. Resistance of Nitric Oxide Dioxygenase and Cytochrome c Oxidase to Inhibition by Nitric Oxide and Other Indications of the Spintronic Control of Electron Transfer. Biophysica 2025, 5, 41. https://doi.org/10.3390/biophysica5030041

AMA Style

Gardner PR. Resistance of Nitric Oxide Dioxygenase and Cytochrome c Oxidase to Inhibition by Nitric Oxide and Other Indications of the Spintronic Control of Electron Transfer. Biophysica. 2025; 5(3):41. https://doi.org/10.3390/biophysica5030041

Chicago/Turabian Style

Gardner, Paul R. 2025. "Resistance of Nitric Oxide Dioxygenase and Cytochrome c Oxidase to Inhibition by Nitric Oxide and Other Indications of the Spintronic Control of Electron Transfer" Biophysica 5, no. 3: 41. https://doi.org/10.3390/biophysica5030041

APA Style

Gardner, P. R. (2025). Resistance of Nitric Oxide Dioxygenase and Cytochrome c Oxidase to Inhibition by Nitric Oxide and Other Indications of the Spintronic Control of Electron Transfer. Biophysica, 5(3), 41. https://doi.org/10.3390/biophysica5030041

Article Metrics

Back to TopTop