Abstract
The electron transport chain (ETC) is a major currency converter that exchanges the chemical energy of fuel oxidation to proton motive force and subsequently, ATP, generation, using O2 as a terminal electron acceptor. Discussed herein, two new studies reveal that the mammalian ETC is forked. Hypoxia or H2S exposure promotes the use of fumarate as an alternate terminal electron acceptor. The fumarate/succinate and CoQH2/CoQ redox couples are nearly iso-potential, revealing that complex II is poised for facile reverse electron transfer, which is sensitive to CoQH2 and fumarate concentrations. The gas regulators, H2S and •NO, modulate O2 affinity and/or inhibit the electron transfer rate at complex IV. Their induction under hypoxia suggests a mechanism for how traffic at the ETC fork can be regulated.
Keywords: electron transport chain, fumarate, complex II, electron acceptor, hydrogen sulfide, hypoxia, nitric oxide
The electron transport chain is a major currency converter
Oxidation-reduction (or redox) reactions (see Glossary) constitute life’s currency, harnessing energy released from the thermodynamically favorable movement of low-potential donor electrons to high-potential acceptors. The electron transport chain (ETC) is a major currency converter in the cellular marketplace of metabolically generated low-potential reductants, namely NADH, FADH2 and CoQH2. At the redox donor end of the ETC, proton-coupled electron transfer generates a proton motive force across the inner mitochondrial membrane that is dispensed as ATP via oxidative phosphorylation and stored as potential energy to power cellular functions such as membrane transport. At the redox acceptor end, O2, a high-potential oxidant, is churned to water in four one-electron steps. The cellular profit from glucose via ETC-fueled ATP generation can be ~18-fold greater than from aerobic glycolysis [1].
The high ATP exchange rate of oxidative phosphorylation is not, however, the sole determinant of cellular metabolic choice. It is well known that various cell types such as pluripotent stem cells [2], proliferating vascular endothelial cells [3], and malignant cells [4] are highly glycolytic. In these instances, the lucrative ATP yield from oxidative phosphorylation is sacrificed for repurposing mitochondria as anabolic hubs, supporting macromolecular precursor synthesis [4]. The complex regulatory mechanisms governing the extent to which the aerobic glycolysis versus the oxidative phosphorylation currency exchange counter is utilized, continue to be elucidated.
O2 availability represents a plausible mechanism for substrate-level regulation of the ETC, governing the relative usage of glycolysis versus oxidative phosphorylation. Tissues and cell types are exposed to variable O2 tensions, ranging from a low of 0.4% in the colonic lumen to 13% in regions of lung [5]. Even within an organ, O2 tensions can vary widely, e.g., from 1.9% in the renal medulla to 9% in the cortex [5]. However, the potential for O2 concentration per se to substantially influence ETC flux is questionable since the KM(O2) for cytochrome c oxidase is very low. Depending on the rate of electron throughput, KM(O2) is estimated to range from <0.2 to 1 μM, corresponding to <0.02–0.1% O2 tension [6]. Since the O2 tension in most tissues exceeds the KM(O2) for cytochrome c oxidase, alternative cellular strategies must exist to regulate O2 affinity and/or proton-coupled electron transfer rate through the ETC. The concentration of electron donors, the proton motive force and ADP levels are other factors that influence the rate of electron transfer to cytochrome c oxidase. In this review, we discuss how the gas regulators, i.e., hydrogen sulfide (H2S), nitric oxide (•NO), and O2 interact to impact the ETC exchange rate by complex II reversal, using fumarate as an alternate terminal electron acceptor [7, 8]. We also discuss how the redox properties of complex II make it uniquely poised for bidirectional currency transactions in the ETC that is sensitive to fluctuations in the metabolic marketplace.
ETC architecture
The ETC is a marvel of cellular engineering and one that is perhaps unrivaled in complexity. Some 69 subunits encoded by the nuclear and mitochondrial genomes come together for the assembly of the ETC core, comprising complexes I, III, and IV (Fig. 1A). Complex I is an NADH:ubiquinone oxidoreductase, moving electrons from NADH to ubiquinone or coenzyme Q10 (CoQ) as protons are translocated from the matrix to the inner mitochondrial membrane space. Complex II, comprising 4 subunits, is a component of the Krebs cycle (succinate dehydrogenase) and transfers electrons from FADH2 to CoQ as succinate is oxidized to fumarate. Other mitochondrial dehydrogenases (dihydroorotate dehdydrogenase, glycerol 3-phosphate dehydrogenase, and proline dehdydrogenase) and oxidoreductases (sulfide quinone oxidoreductase (SQOR) and the electron transfer flavoprotein quinone oxidoreductase (ETF-QO)) also exchange redox currency at the ETC, using CoQ as an acceptor, and move electrons from organic (e.g., proline, dihydroorotate) or inorganic (H2S) donors. Electrons from CoQH2 (or ubiquinol) are funneled to complex III (cytochrome bc1 complex) and, via cytochrome c, to complex IV or (cytochrome c oxidase) where O2 serves as the terminal electron acceptor. Complexes III and IV also pump protons as they transfer electrons. An ~1160 mV potential difference between the NADH/NAD+ (E°′= −340 mV) and 0.5 H2O/O2 (+820 mV) redox couples translates to a maximal transmembrane protonmotive force of 180–220 mV [9]. Proton reentry to the mitochondrial matrix through complex V (ATP synthase) drives ATP synthesis. The ETC leaks both protons and electrons, with the latter generating reactive oxygen species, which can have beneficial signaling roles at low, but are damaging at high concentrations [10].
Figure 1. The mammalian ETC is forked.
A. Various electron donors support electron flux through the ETC via proton-coupled electron transfer. CoQ is an intermediate electron carrier in the membrane and accepts electrons from complexes I and II and donates electrons to complex III. CoQ also accepts electrons from various housekeeping enzymes like dihydroorotate dehydrogenase (DHOD), proline dehydrogenase (PRODH), glycerol 3-phosphate dehydrogenase (G3PDH), electron transfer flavoprotein ubiquinol oxidoreductase (ETF-QO), and sulfide quinone reductase (SQOR). Proton translocation occurs at complexes I, III and IV, building up a proton motive force. B. Modulation of O2 affinity and/or electron transfer activity at complex IV by •NO, H2S or O2 restriction due to hypoxia can inhibit O2-dependent forward electron transfer through the ETC, leading to a reductive shift in the NADH/NAD+ and CoQH2/CoQ pools (red up arrows). Under these conditions, complex II can reverse direction, using fumarate as a terminal electron acceptor. For simplicity, the establishment of redox cycles between complex II operating in reverse and DHOD and SQOR are shown here, sustaining de novo pyrimidine synthesis and H2S oxidation, respectively. In principle, these redox cycles can be established between other redox couples that use CoQ as an electron acceptor. C. The bifurcation in the mammalian ETC allows electron flow to either complex II or complex IV where fumarate and O2, respectively are utilized as terminal electron acceptors. Only complex I is shown as the electron donor for simplicity although any CoQ-dependent enzyme can use the forked path.
The ETC is forked and can use fumarate as a terminal electron acceptor
Metabolic reprogramming is an adaptive strategy that increases network diversity in response to extracellular cues (e.g., O2 or nutrient availability) or cellular needs (e.g., proliferation). Classic examples include diverting pyruvate to lactate thereby disconnecting glycolysis from oxidative phosphorylation (i.e., the Warburg effect [11]), and reversing direction within the Krebs cycle (e.g., reductive carboxylation [12]). Two recent studies have illuminated the capacity for reprogramming within the mammalian ETC, leading to electron shunting to fumarate either constitutively or under stress, or when complex III or IV is inhibited [7, 8] (Fig. 1B). Fumarate is the two-electron oxidation product of succinate in a reaction catalyzed by succinate dehydrogenase that concomitantly reduces FAD to FADH2. Inhibition of complex III or IV leads to a reductive shift upstream in the ETC as recycling of CoQH2 to CoQ is slowed, creating an electron acceptor insufficiency with potentially widespread metabolic ramifications. Succinate accumulation is a signature of cells exposed to H2S [7] or hypoxia, as well as of cells lacking complex III or complex IV activity [8]. Succinate also accumulates in other physiological states such as in the ischemic heart [13] and in post exercise muscle [14].
Electron shunting to fumarate from complex I and DHOD (Fig. 1B) is observed when O2-dependent electron flux is impaired [8]. By using fumarate in lieu of O2, cells sustain de novo pyrimidine biosynthesis, supporting proliferation. In cell culture, fumarate reduction increases progressively as O2 concentration decreases from 20% to 1%, plateauing at ~3%. Elegant in vivo isotope tracing studies reveal that tissues exhibit considerable heterogeneity in their capacity to reduce fumarate [8]. Thus, while some tissues (liver, brain, and kidney) display fumarate reduction constitutively and at levels that exceed succinate oxidation, others use this option scarcely, if at all (e.g., lung, heart, white adipose tissue, gastrocnemius muscle, pancreas, thymus). Exercise on the other hand, increases fumarate reduction in heart and white adipose tissue. Hypoxia (1%) increases fumarate reduction in most tissues, ex vivo [8].
The variable extent to which tissues use fumarate as a terminal electron acceptor is not simply correlated with either fumarate levels or the CoQH2/CoQ ratio [8] and suggests that other strategies are at play for regulating electron flow at the fork in the ETC (Fig. 1B,C). One such regulator is H2S, an endogenously produced gas, which at low concentrations, stimulates respiration but at high concentrations, inhibits complex IV [15]. The rapidity of the cellular response to H2S [16] is consistent with the hydrophobicity of H2S and its tendency to concentrate in membrane versus aqueous solvent [17]. While the pH dependence for inhibition of cytochrome c oxidase by H2S has not been reported, its reactivity with other heme proteins is enhanced at lower pH, implying that the protonated rather than the anionic form of sulfide interacts with their metal centers [18–20].
H2S induces a reductive shift in the NADH/NAD+ pool that is accompanied by increased glucose consumption and lactate generation [16, 21]. H2S also increases lipid biogenesis, which is fueled by glutamine-dependent reductive carboxylation [22], hallmarks of reductive stress. A shift in the mitochondrial but not cytoplasmic NADH/NAD+ pool triggers these H2S-driven metabolic changes and is attenuated by dissipation of the mitochondrial NADH pool using the water-forming NADH oxidase LbNOX [23]. The paradoxical observation that H2S clearance continues even at very low O2, first hinted at the existence of a complex IV-independent pathway for its oxidation [24] and led to the discovery that complex II reversal is a mechanism for prioritizing H2S clearance when respiration is inhibited [7]. In H2S treated cells, the malate-aspartate shuttle and the purine nucleotide cycle furnish fumarate to support reverse complex II activity [7], which are also the sources of fumarate leading to ischemic succinate accumulation [13].
Fumarate reduction has been previously reported under conditions where complex IV activity is greatly diminished or absent. For instance, fumarate reduction is used as an adaptive strategy in retina that is exposed to a chronically hypoxic environment, which leads to downregulation of a complex IV subunit [25]. In an ex vivo study, fumarate reduction was described as a component of a malate-succinate shuttle for the transfer of electron equivalents (as succinate) across an O2 gradient, i.e., from O2-poor retina to O2-rich retinal pigment epithelium-choroid [25]. Reverse complex II activity has also been described in ischemia when complex IV activity is precluded by the lack of O2 [13]. The relative contributions of the Krebs cycle versus reverse complex II activity to ischemic succinate accumulation is controversial [13, 26].
H2S and NO•: Gas regulators of the ETC
H2S and •NO are metabolically generated gases derived from amino acids and function as signaling molecules that elicit varied physiological effects [27, 28]. H2S is synthesized from cysteine and/or homocysteine by the action of cystathionine β-synthase, γ-cystathionase, or mercaptopyruvate sulfurtransferase [29] while arginine is the source of •NO synthesized by nitric oxide synthases [30] (Fig. 2A,B). A third gas signaling molecule, CO, is a synthesized from heme by heme oxygenases, but unlike •NO and H2S, its effects on the ETC are relatively poorly characterized [31]. •NO and H2S are reversible inhibitors of complex IV. Adding complexity to their interaction however, •NO and H2S also serve as substrates and are oxidized by complex IV to nitrate [32] and a putative hydrodisulfide species [33], respectively. As a competitive inhibitor, •NO increases the KMapp for O2 for cytochrome c oxidase. As O2 concentrations vary from 10 to 145 μM, the IC50 value for •NO increases from ~20 to 270 nM, which is within the range of estimated tissue •NO concentrations (10–450 nM) [34]. The IC50 value is comparable to the ED50 (10–20 nM) for soluble guanylate cyclase, a major •NO receptor that signals via the second messenger cGMP [35, 36].
Figure 2. Scheme showing biogenesis of •NO and H2S.
A. H2S is synthesized by three enzymes of which cystathionine β-synthase (CBS) and γ-cystathionase (CSE) function canonically in the transsulfuration pathway. At cellular concentrations of reactants, the transsulfuation reactions (blue arrows) are preferentially catalyzed. Depending on cysteine availability, the enzymes can switch to H2S biogenesis as shown. CBS and CSE catalyze multiple H2S-generating reactions but for simplicity, only the major reactions at physiologically relevant concentrations of substrate are shown. Mercaptopyruvate sulfur transferase (MPST) converts 3-mercaptopyruvate to pyruvate and H2S in the presence of a reductant (not shown). B. Arginine is the source of •NO catalyzed by nitric oxide synthases (NOS) in the presence of reducing equivalents, which are not shown here for simplicity. Hcy and Cys denote homocysteine and cysteine, respectively.
An H2S sensor analogous to soluble guanylate cyclase has not been identified. However, SQOR could potentially function in this capacity, generating CoQH2 and reactive sulfur species as potential second messengers [37]. SQOR is a membrane anchored flavoprotein with an unusual cysteine trisulfide redox cofactor that oxidizes H2S to glutathione persulfide (GSSH), concomitantly reducing CoQ to CoQH2 (Fig. 3) [38]. Electrons from H2S oxidation can enter the ETC at the level of complex III, making sulfide an inorganic substrate for ATP generation [39]. GSSH is further oxidized to sulfite via the action of ETHE1, a persulfide dioxygenase [40], or can donate its sulfane sulfur for thiosulfate synthesis catalyzed by thiosulfate sulfurtransferase (or rhodanese) [41]. The reported Ki for the inhibition of cytochrome c oxidase by H2S is 200 nM [42]. SQOR shields complex IV from respiratory poisoning by H2S and its expression levels are correlated with the cellular sensitivity to this gas [16]. The steady-state concentration of H2S ranges between 10–50 nM in tissues where reliable estimates exist and is considerably higher (~1 μM) in aorta [24, 43]. Some cell types, like the epithelial cells lining the large intestine, are exposed to high concentrations of exogenous H2S (0.2–2.4 mM) derived from microbial metabolism [44, 45] and express high levels of the sulfide oxidation pathway enzymes for its detoxification [16].
Figure 3. Sulfide oxidation connects H2S to the ETC.
H2S is oxidized via the mitochondrial sulfide oxidation pathway comprising SQOR in the mitochondrial inner membrane, which converts H2S to glutathione persulfide (GSSH) while reducing CoQ to CoQH2. GSSH is further oxidized to sulfite by ETHE1, a persulfide dioxygenase, or serves as a sulfane sulfur donor, forming thiosulfate in a reaction catalyzed by thiosulfate sulfur transferase (TST, also known as rhodanese). In some cells, thiosulfate is further oxidized to sulfate in a reaction catalyzed by sulfite oxidase (not shown).
By increasing the KM for O2 by •NO or by decreasing the rate of O2 reduction by H2S [42, 46], both gases can dial down ETC flux, increase local O2 levels, and potentially modulate the O2 threshold for hypoxia sensing. •NO induces metabolic hypoxia, a paradoxical state in which O2 increases due to cytochrome c oxidase inhibition and leads to destabilization of HIF1α under hypoxic conditions [47]. Under these conditions, O2 can be diverted to other enzymes whose activities might otherwise be limited under hypoxia. For example, •NO-enhanced O2 availability (that is independent of cGMP), stimulates luciferase activity and is important for bioluminescent communication during firefly courtship [48].
HIF and ATF4 regulation of H2S and •NO
•NO and H2S biogenesis are upregulated by hypoxia [49], nutrient stress, and the unfolded protein response [50, 51] pathway (Fig. 4). A molecular intersection between •NO (and CO) and H2S signaling occurs at cystathionine β-synthase, the first enzyme in the transsulfuration pathway, which in turn switches the substrate preference of γ-cystathionase, the second enzyme in the pathway, from cystathionine to cysteine [52]. While cystathionine β-synthase and γ-cystathionase are both promiscuous [53], kinetic control renders catalysis of cystathionine and cysteine generation, respectively as their dominant reactions at physiologically relevant substrate concentrations (Fig. 2A). The activity of cystathionine β-synthase is inhibited when •NO or CO binds to the regulatory ferrous heme cofactor [54, 55]. With reduced competition from cystathionine under these conditions, the metabolic track preference of γ-cystathionase flips from cystathionine to H2S biogenesis [52] (Fig. 4). Reprogramming of the transsulfuration pathway upregulates H2S synthesis under endoplasmic reticulum stress conditions [52] and could also be induced in response to other triggers such as hypoxia.
Figure 4. Regulation of gas modulator synthesis.
CSE is regulated by the transcription factor, ATF4 which is activated by amino acid starvation or during the unfolded protein response, leading to increased H2S synthesis (red arrow). Under hypoxia, NOS and heme oxygenase are induced leading to •NO and CO biosynthesis, respectively, which inhibit CBS, a hemeprotein, leading to reduced cystathionine. Under these conditions, the transsulfuration pathway enzymes switch metabolite tracks from the canonical cystathionine and cysteine producing reactions (white arrows) to H2S biogenesis primarily by CSE (red arrow). Thus, the three cellular triggers shown here have the potential to upregulate synthesis of the ETC gas regulators, H2S and •NO. Cyst denotes cystathionine.
Complex II is tuned for reversible electron transfer
A sizeable potential difference (ΔE°′) exists between CoQH2/CoQ and the CoQ-dependent mitochondrial oxidoreductases except for complex II and ETF-QO (Table 1), and provides the thermodynamic driving force for CoQ reduction. The large ΔE°′ also precludes ready reversal of electron transfer from CoQH2 at these redox junctions. In contrast, the CoQH2/CoQ and fumarate/succinate couples are almost iso-potential, rendering the direction of electron transfer catalyzed by complex II sensitive to the mitochondrial CoQH2/CoQ and fumarate/succinate ratios. Consistent with the predicted facile reversibility of succinate dehydrogenase, is its ability to support E. coli growth on fumarate as a terminal electron acceptor under anaerobic conditions, in a strain that lacks fumarate reductase [56].
Table 1.
Redox potentials for reactions coupling to the mitochondrial CoQH2/CoQ pool
Enzyme | Redox Couple | E°′ | Reference |
---|---|---|---|
DHOD | Dihydroorotate/Orotate | −252 V | [62] |
PRODH | Proline/Δ1-pyrroline-5-carboxylate | −123 mV | [63] |
ETF-QO | 1FADsq/FADox | +28 mV | [64] |
SQOR | 2H2S/HSSH | −230 mV | [65] |
G3PDH | Glycerol 3-phosphate/dihydroxyacetone phosphate | −190 mV | [66] |
Complex II | Fumarate/succinate | +30 mV | |
Complex I | 3NADH/NAD+ | −340 mV | [67] |
CoQH2/CoQ | +40 to +60 mV | [68] |
The one-electron redox potential for FAD in ETF-QO is reported as CoQ reduction by ETF-QO reportedly occurs in two one-electron transfer steps.
The two-electron redox potential reported for the H2S/HSSH couple is listed. The reaction catalyzed by SQOR leads to the synthesis of the alkyl persulfide, GSSH.
Value for pH 7.8
Conditions that increase CoQH2 levels (e.g. reductive stress triggered by H2S, hypoxia, complex III or complex IV dysfunction) and fumarate (via metabolic reprogramming), favor complex II reversal as seen in human cells exposed to H2S [7] or hypoxia [8] or inhibited at complex III or IV [8]. Consistent with this prediction, dissipation of the mitochondrial CoQH2 pool by ectopic expression of the alternate oxidase suppresses fumarate reduction in cells with impaired complex III or IV activity [8]. In principle, allosteric regulation of complex II, for example via posttranslational modification, could also influence its propensity to reverse the direction of electron flow but such a mechanism, if it exists, remains to be identified. In ETF-QO, the one-electron redox potential for the proximal electron donor to CoQ, i.e., the flavin semiquinone/flavin couple, is high (Table 1). ETF-QO is a component of a multiprotein electron wire comprising ETF and a family of acyl-CoA dehydrogenases. Unlike complex II, reverse electron transfer through this system is unlikely to be facile.
We posit that the concentration of gas regulators, H2S and •NO relative to O2, modulates complex IV activity, underlying the variable capacity of tissues to siphon off electrons to fumarate under steady-state conditions or during an acute response. The physiological relevance of H2S-induced electron rerouting from CoQH2 to fumarate is supported by the substantial decrease in sulfide oxidation in mice with a targeted deletion of complex II in intestinal epithelial cells, which are exposed to the gas from gut microbes [7].
Prokaryotes adapted to growth in O2 limiting or anaerobic environments, and eukaryotes like parasitic helminths and freshwater snails that are facultative anaerobes, have long been known to use fumarate as an alternate electron. In these organisms, fumarate reduction is correlated with the use of lower potential quinone derivatives such menaquinone and rhodoquinone [57]. In mammals, the establishment of an overall redox cycle between CoQ and fumarate versus O2 as a terminal electron acceptor, shaves off a sizeable thermodynamic advantage (ΔΔE°′ ~770 mV). Nevertheless, the redox couples remain highly favorable for substrate oxidation and fumarate reduction e.g., ΔE°′ is ~ +302 mV for DHOD and ~+280 mV for SQOR, corresponding to ΔG°′ values of −13.9 kcal/mol and −12.9 kcal/mol, respectively. Similarly, the ΔE°′ for coupling NADH oxidation by complex I to fumarate reduction by complex II is ~ +390 mV, corresponding to a ΔG°′ of −18 kcal/mol, rendering electron flow to either O2 or fumarate highly favorable (Fig. 1C). Importantly, the CoQ-fumarate-dependent redox cycle provides an avenue for the continued operation of CoQ-dependent reactions when forward electron transfer to complex IV is impeded and aerobic glycolysis assumes a more significant role in ATP generation.
Concluding Remarks
Two recent studies have revealed a homeostatic role for complex II reversal in tissue physiology (e.g., liver, kidney, brain, and colon) under conditions that lead to a reductive shift in the CoQH2/CoQ redox pool [7, 8]. Based on the similar potentials of the CoQH2/CoQ and fumarate/succinate redox couples, complex II appears to be uniquely poised to bifurcate the ETC, increasing its functional versatility. The gas regulators, •NO and H2S, can serve as valves for tuning O2 affinity and/or inhibiting the electron transfer rate in complex IV, which in turn, influences partitioning upstream at the ETC fork (Figs. 1B,C).
The contribution of fumarate reduction to metabolic physiology awaits fuller elucidation and might be particularly relevant in the context of pathologies that are marked by reductive stress such as diabetes, cancer, ischemia and obesity. H2S is cytoprotective when co-administered during reperfusion [58], consistent with its ability to inhibit complex IV and drive complex II reversal, thereby reducing reactive oxygen species generation that is initiated by oxidation of succinate, which accumulates during ischemia. Dissipation of the NADH pool by LbNOX improves glucose tolerance in mice on a high fat diet, suggesting the therapeutic potential of a general strategy for targeting reductive stress [59]. Hereditary mutations in complex II subunits are associated with familial paraganglioma and Leigh’s syndrome [60]. Considering the bidirectional capacity of complex II in the cellular milieu, the metabolic vulnerabilities resulting from the loss of fumarate reduction in addition to succinate oxidation will have to be considered to understand the underlying disease mechanisms. Interestingly, among the known risk factors for nonfamilial paraganglioma are chronic hypoxia, including high altitude living and cardiorespiratory diseases [60, 61]. Future studies will unravel the molecular details of how electron flow is regulated at the ETC fork and the bioenergetic and signaling contexts for and the consequences of fumarate versus O2 utilization (see Outstanding Questions).
OUTSTANDING QUESTIONS.
Why is fumarate a significant electron acceptor in some but not other tissues?
Is the varying use of fumarate by the ETC correlated with varying levels of H2S-, •NO-induced or other type of reductive stress?
What is the primary source of fumarate in tissues that utilize it as a significant electron acceptor?
What is the metabolic fate of succinate formed via fumarate reduction?
What are the regulatory mechanisms that coordinate electron flux at the ETC fork?
Does hypoxia promote utilization of fumarate as an electron acceptor by enhancing H2S and/or •NO biogenesis?
How is H2S synthesis regulated in normal and in disease states?
Does the dependence on fumarate as an electron acceptor change in disease states such as diabetes, obesity and cancer?
How much of the CoQ pool is free in the mitochondrial membrane?
How does the composition and proportion of respiratory supercomplexes influence electron flux at the ETC fork?
Does reverse electron transfer through complex II support intercellular redox cycles in tissues other than retina?
What is the therapeutic potential for modulating fumarate reduction? For example, could enhancing fumarate reduction during the reperfusion phase decrease ischemic injury?
HIGHLIGHTS.
Two recent studies reported the discovery that fumarate can serve as a terminal electron acceptor in mammalian tissues, revealing that the electron transport pathway is forked.
The similar redox potentials of the fumarate/succinate and reduced/oxidized coenzyme Q couples, indicate that complex II is poised to operate either in the direction of succinate oxidation or fumarate reduction.
Interaction between the gas regulators, i.e., O2, hydrogen sulfide and nitric oxide, could modulate flux at the electron transport chain fork, siphoning electrons to complex II or complex III.
It is critically important to understand the underlying mechanisms for tissue and cell-specific variations in the use of fumarate as a terminal electron acceptor in normal and disease states.
Acknowledgements
This work was supported in part by grants from the National Institutes of Health (GM130183 to RB and the American Heart Association (826245 to RK).
Glossary
- Cytochrome c
A hemeprotein that transfers electrons between complex III and complex IV in the electron transport chain.
- Coenzyme Q10 (CoQ)
A ubiquinone derivative with ten isoprene units attached to the benzoquinone head group. CoQ is a redox active membrane cofactor and can exist in three oxidation states: fully reduced (CoQH2), radical semiquinone (CoQH•), and fully oxidized (CoQ).
- Dihydroorotate dehydrogenase (DHOD)
A mitochondrial enzyme that catalyzes CoQ-dependent oxidation of dihydroorotate to orotate, in the de novo pyrimidine biosynthetic pathway.
- Electron transport chain
Comprises a series of protein complexes I–IV that transfer electrons while concomitantly generating a proton gradient across the mitochondrial inner membrane. In mammals, O2 and fumarate serve as terminal electron acceptors.
- Electron transfer flavoprotein quinone oxidoreductase (ETF-QO)
Transfers electrons to CoQ from ETF, which in turn is reduced by a family of acyl-CoA dehydrogenases that are involved in the β-oxidation of fatty acids.
- Fumarate
A C4-dicarboxylate intermediate in the Krebs cycle and a terminal electron acceptor for the ETC.
- Glycolysis
A cytoplasmic pathway for glucose oxidation to pyruvate that generates ATP and NADH.
- Hypoxia
A state in which the tissue O2 tension is restricted and leads to stabilization of hypoxia inducible factors (HIFs), which initiate a cascade of signaling events.
- K M
The substrate concentration at which half maximal activity of an enzyme is seen (also referred to as the Michaelis constant).
- Malate-aspartate shuttle
An intercompartmental metabolic shuttle that allows the transfer of cytosolic reducing equivalents (from NADH) as malate, which crosses the mitochondrial inner membrane via a reversible α–ketoglutarate/malate carrier. Mitochondrial aspartate is exchanged for cytosolic glutamate through the aspartate-glutamate carrier.
- Mitochondrial glycerol 3-phosphate dehydrogenase (G3PDH)
Catalyzes the CoQ-dependent oxidation of glycerol 3-phosphate to dihydroxyacetone phosphate. Together with cytoplasmic G3PDH, which catalyzes the NADH-dependent reverse reaction, constitutes the glycerophosphate shuttle, linking carbohydrate and lipid metabolism. The shuttle also serves to transfer reducing equivalents from cytosolic NADH to the mitochondrial ETC.
- NAD+
Nicotinamide adenine dinucleotide is a redox active coenzyme used by many enzymes. NADH is the reduced form.
- Proton motive force (PMF)
Refers to the transmembrane proton gradient. In the mitochondrion, proton coupled electron transfers by complexes I, III and IV generate the proton motive force. The movement of protons back to the mitochondrial matrix via complex V powers ATP synthesis in a process referred to as oxidative phosphorylation. The PMF is also used to power other cellular functions such as membrane transport.
- Purine nucleotide cycle
A sequence of reactions in which aspartate and inosine monophosphate (IMP) are converted to adenylosuccinate, which is cleaved to fumarate and adenosine monophosphate (AMP). Deamination of AMP releases ammonia and IMP and completes the cycle.
- Proline dehydrogenase (PRODH)
A mitochondrial enzyme that catalyzes the CoQ-dependent oxidation of L-proline to Δ1-pyrroline-5-carboxylate.
- Redox reaction
Oxidation-reduction (redox) reaction involving an electron donor (reductant) and an electron acceptor (oxidant).
- Reductive carboxylation
Conversion of α-ketoglutarate to isocitrate catalyzed by isocitrate dehydrogenase. This reaction is the reverse of oxidative decarboxylation catalyzed by the same enzyme in the Krebs cycle.
- Sulfide quinone oxidoreductase (SQOR)
Is a mitochondrial inner membrane enzyme that catalyzes the oxidation of H2S to glutathione persulfide (GSSH) as it reduces CoQ to CoQH2.
Footnotes
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Declaration of interests
No interests are declared.
References
- 1.Semenza GL (2007) Oxygen-dependent regulation of mitochondrial respiration by hypoxia-inducible factor 1. Biochem J 405 (1), 1–9. [DOI] [PubMed] [Google Scholar]
- 2.Zhang J et al. (2011) UCP2 regulates energy metabolism and differentiation potential of human pluripotent stem cells. EMBO J 30 (24), 4860–73. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Quintero M et al. (2006) Mitochondria as signaling organelles in the vascular endothelium. Proc Natl Acad Sci U S A 103 (14), 5379–84. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Ward PS and Thompson CB (2012) Metabolic reprogramming: a cancer hallmark even Warburg did not anticipate. Cancer Cell 21 (3), 297–308. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Ast T and Mootha VK (2019) Oxygen and mammalian cell culture: are we repeating the experiment of Dr. Ox? Nat Metab 1 (9), 858–860. [DOI] [PubMed] [Google Scholar]
- 6.Massari S et al. (1996) The variation of Km for oxygen of cytochrome oxidase with turnover under de-energized and energized conditions. Biochem Soc Trans 24 (3), 464S. [DOI] [PubMed] [Google Scholar]
- 7.Kumar R et al. (2021) A redox cycle with complex II prioritizes sulfide quinone oxidoreductase dependent H2S oxidation. J Biol Chem, 101435. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Spinelli JB et al. (2021) Fumarate is a terminal electron acceptor in the mammalian electron transport chain. Science 374 (6572), 1227–1237. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Brand MD and Nicholls DG (2011) Assessing mitochondrial dysfunction in cells. Biochem J 435 (2), 297–312. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Schieber M and Chandel NS (2014) ROS function in redox signaling and oxidative stress. Curr Biol 24 (10), R453–62. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Warburg O et al. (1924) On the metabolism of carcinoma cells. Biochem. Z 152, 309–344. [Google Scholar]
- 12.Metallo CM et al. (2011) Reductive glutamine metabolism by IDH1 mediates lipogenesis under hypoxia. Nature 481 (7381), 380–4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Chouchani ET et al. (2014) Ischaemic accumulation of succinate controls reperfusion injury through mitochondrial ROS. Nature 515 (7527), 431–435. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Reddy A et al. (2020) pH-Gated Succinate Secretion Regulates Muscle Remodeling in Response to Exercise. Cell 183 (1), 62–75 e17. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Nicholls P and Kim JK (1982) Sulphide as an inhibitor and electron donor for the cytochrome c oxidase system. Can J Biochem 60 (6), 613–23. [DOI] [PubMed] [Google Scholar]
- 16.Libiad M et al. (2019) Hydrogen sulfide perturbs mitochondrial bioenergetics and triggers metabolic reprogramming in colon cells. J Biol Chem 294 (32), 12077–12090. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Cuevasanta E et al. (2012) Solubility and permeation of hydrogen sulfide in lipid membranes. PloS One 7 (4), e34562. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Vitvitsky V et al. (2015) Sulfide oxidation by a noncanonical pathway in red blood cells generates thiosulfate and polysulfides. J Biol Chem 290, 8310–20. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Bostelaar T et al. (2016) Hydrogen Sulfide Oxidation by Myoglobin. J Am Chem Soc 138 (27), 8476–88. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Ruetz M et al. (2017) A Distal Ligand Mutes the Interaction of Hydrogen Sulfide with Human Neuroglobin. J Biol Chem 292, 6512–6528. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Vitvitsky V et al. (2021) The mitochondrial NADH pool is involved in hydrogen sulfide signaling and stimulation of aerobic glycolysis. J Biol Chem, 100736. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Carballal S et al. (2021) Hydrogen sulfide stimulates lipid biogenesis from glutamine that is dependent on the mitochondrial NAD(P)H pool. J Biol Chem, 100950. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Titov DV et al. (2016) Complementation of mitochondrial electron transport chain by manipulation of the NAD+/NADH ratio. Science 352 (6282), 231–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Vitvitsky V et al. (2012) High turnover rates for hydrogen sulfide allow for rapid regulation of its tissue concentrations. Antioxid Red Signal 17 (1), 22–31. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Bisbach CM et al. (2020) Succinate Can Shuttle Reducing Power from the Hypoxic Retina to the O2-Rich Pigment Epithelium. Cell Rep 31 (5), 107606. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Zhang J et al. (2018) Accumulation of Succinate in Cardiac Ischemia Primarily Occurs via Canonical Krebs Cycle Activity. Cell Rep 23 (9), 2617–2628. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Filipovic MR et al. (2018) Chemical Biology of H2S Signaling through Persulfidation. Chem Rev 118 (3), 1253–1337. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Martinez-Ruiz A et al. (2011) Nitric oxide signaling: classical, less classical, and nonclassical mechanisms. Free Radic Biol Med 51 (1), 17–29. [DOI] [PubMed] [Google Scholar]
- 29.Singh S and Banerjee R (2011) PLP-dependent H2S biogenesis. Biochim Biophys Acta 1814, 1518–1527. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Stuehr DJ (1999) Mammalian nitric oxide synthases. Biochim Biophys Acta 1411 (2–3), 217–30. [DOI] [PubMed] [Google Scholar]
- 31.Alonso JR et al. (2003) Carbon monoxide specifically inhibits cytochrome c oxidase of human mitochondrial respiratory chain. Pharmacol Toxicol 93 (3), 142–6. [DOI] [PubMed] [Google Scholar]
- 32.Torres J et al. (2000) Cytochrome c oxidase rapidly metabolises nitric oxide to nitrite. FEBS Lett 475 (3), 263–6. [DOI] [PubMed] [Google Scholar]
- 33.Nicholls P et al. (2013) Sulfide inhibition of and metabolism by cytochrome c oxidase. Biochem Soc Trans 41 (5), 1312–6. [DOI] [PubMed] [Google Scholar]
- 34.Moncada S and Erusalimsky JD (2002) Does nitric oxide modulate mitochondrial energy generation and apoptosis? Nat Rev Mol Cell Biol 3 (3), 214–20. [DOI] [PubMed] [Google Scholar]
- 35.Carter TD et al. (1997) Potency and kinetics of nitric oxide-mediated vascular smooth muscle relaxation determined with flash photolysis of ruthenium nitrosyl chlorides. Br J Pharmacol 122 (6), 971–3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Bellamy TC et al. (2000) Rapid desensitization of the nitric oxide receptor, soluble guanylyl cyclase, underlies diversity of cellular cGMP responses. Proc Natl Acad Sci U S A 97 (6), 2928–33. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Mishanina TV et al. (2015) Biogenesis of reactive sulfur species for signaling by hydrogen sulfide oxidation pathways. Nat Chem Biol 11 (7), 457–64. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Landry AP et al. (2019) A catalytic trisulfide in human sulfide quinone oxidoreductase catalyzes coenzyme A persulfide synthesis and inhibits butyrate oxidation. Cell Chem Biol 26 (11), 1515–1525 e4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Goubern M et al. (2007) Sulfide, the first inorganic substrate for human cells. FASEB J 21 (8), 1699–706. [DOI] [PubMed] [Google Scholar]
- 40.Kabil O et al. (2018) Mechanism-based inhibition of human persulfide dioxygenase by gamma-glutamyl-homocysteinyl-glycine. J Biol Chem 293 (32), 12429–12439. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Libiad M et al. (2015) Polymorphic variants of human rhodanese exhibit differences in thermal stability and sulfur transfer kinetics. J Biol Chem 290 (39), 23579–88. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Petersen LC (1977) The effect of inhibitors on the oxygen kinetics of cytochrome c oxidase. Biochim Biophys Acta 460 (2), 299–307. [DOI] [PubMed] [Google Scholar]
- 43.Levitt MD et al. (2011) Free and Acid-labile Hydrogen Sulfide Concentrations in Mouse Tissues: Anomalously High Free Hydrogen Sulfide in Aortic Tissue. Antioxid Redox Signal 15, 373–8. [DOI] [PubMed] [Google Scholar]
- 44.Macfarlane GT et al. (1992) Comparison of fermentation reactions in different regions of the human colon. J Appl Bacteriol 72 (1), 57–64. [DOI] [PubMed] [Google Scholar]
- 45.Deplancke B et al. (2003) Gastrointestinal and microbial responses to sulfate-supplemented drinking water in mice. Exp Biol Med (Maywood) 228 (4), 424–33. [DOI] [PubMed] [Google Scholar]
- 46.Brown GC and Cooper CE (1994) Nanomolar concentrations of nitric oxide reversibly inhibit synaptosomal respiration by competing with oxygen at cytochrome oxidase. FEBS Lett 356 (2–3), 295–8. [DOI] [PubMed] [Google Scholar]
- 47.Hagen T et al. (2003) Redistribution of intracellular oxygen in hypoxia by nitric oxide: effect on HIF1alpha. Science 302 (5652), 1975–8. [DOI] [PubMed] [Google Scholar]
- 48.Trimmer BA et al. (2001) Nitric oxide and the control of firefly flashing. Science 292 (5526), 2486–8. [DOI] [PubMed] [Google Scholar]
- 49.Semenza GL (2003) Targeting HIF-1 for cancer therapy. Nat Rev Cancer 3 (10), 721–32. [DOI] [PubMed] [Google Scholar]
- 50.Dickhout JG et al. (2012) Integrated stress response modulates cellular redox state via induction of cystathionine gamma-lyase: cross-talk between integrated stress response and thiol metabolism. J Biol Chem 287 (10), 7603–14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Longchamp A et al. (2018) Amino Acid Restriction Triggers Angiogenesis via GCN2/ATF4 Regulation of VEGF and H2S Production. Cell 173 (1), 117–129 e14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Kabil O et al. (2016) Heme-dependent metabolite switching regulates H2S synthesis in response to ER stress. J Biol Chem 291, 16418–16423. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Banerjee R (2017) Catalytic promiscuity and heme-dependent redox regulation of H2S synthesis. Curr Opin Chem Biol 37, 115–121. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Taoka S and Banerjee R (2001) Characterization of NO binding to human cystathionine beta-synthase: possible implications of the effects of CO and NO binding to the human enzyme. J Inorg Biochem 87 (4), 245–51. [DOI] [PubMed] [Google Scholar]
- 55.Weeks CL et al. (2009) Heme regulation of human cystathionine beta-synthase activity: insights from fluorescence and Raman spectroscopy. J Am Chem Soc 131 (35), 12809–16. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Maklashina E et al. (1998) Anaerobic expression of Escherichia coli succinate dehydrogenase: functional replacement of fumarate reductase in the respiratory chain during anaerobic growth. J Bacteriol 180 (22), 5989–96. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Van Hellemond JJ et al. (1995) Rhodoquinone and complex II of the electron transport chain in anaerobically functioning eukaryotes. J Biol Chem 270 (52), 31065–70. [DOI] [PubMed] [Google Scholar]
- 58.Elrod JW et al. (2007) Hydrogen sulfide attenuates myocardial ischemia-reperfusion injury by preservation of mitochondrial function. Proc Natl Acad Sci U S A 104 (39), 15560–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Goodman RP et al. (2020) Hepatic NADH reductive stress underlies common variation in metabolic traits. Nature 583 (7814), 122–126. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Rutter J et al. (2010) Succinate dehydrogenase - Assembly, regulation and role in human disease. Mitochondrion 10 (4), 393–401. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Astrom K et al. (2003) Altitude is a phenotypic modifier in hereditary paraganglioma type 1: evidence for an oxygen-sensing defect. Hum Genet 113 (3), 228–37. [DOI] [PubMed] [Google Scholar]
- 62.Krakow G and Vennesland B (1961) The equilibrium constant of the dihydroorotic dehydrogenase reaction. J Biol Chem 236, 142–4. [PubMed] [Google Scholar]
- 63.Becker DF and Thomas EA (2001) Redox properties of the PutA protein from Escherichia coli and the influence of the flavin redox state on PutA-DNA interactions. Biochemistry 40 (15), 4714–21. [DOI] [PubMed] [Google Scholar]
- 64.Zhang J et al. (2006) Structure of electron transfer flavoprotein-ubiquinone oxidoreductase and electron transfer to the mitochondrial ubiquinone pool. Proc Natl Acad Sci U S A 103 (44), 16212–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Koppenol WH and Bounds PL (2017) Signaling by sulfur-containing molecules. Quantitative aspects. Arch Biochem Biophys 617, 3–8. [DOI] [PubMed] [Google Scholar]
- 66.Mracek T et al. (2013) The function and the role of the mitochondrial glycerol-3-phosphate dehydrogenase in mammalian tissues. Biochim Biophys Acta 1827 (3), 401–10. [DOI] [PubMed] [Google Scholar]
- 67.Zu Y et al. (2003) Reversible, electrochemical interconversion of NADH and NAD+ by the catalytic (Ilambda) subcomplex of mitochondrial NADH:ubiquinone oxidoreductase (complex I). J Am Chem Soc 125 (20), 6020–1. [DOI] [PubMed] [Google Scholar]
- 68.Urban PF and Klingenberg M (1969) On the redox potentials of ubiquinone and cytochrome b in the respiratory chain. Eur J Biochem 9 (4), 519–25. [DOI] [PubMed] [Google Scholar]