Some of the most threatening human diseases are due to a blockage of the mitochondrial electron transport chain (ETC). In a variety of plants, fungi, and prokaryotes, there is a naturally evolved mechanism for such threats to viability, namely a bypassing of the blocked portion of the ETC by alternative enzymes of the respiratory chain. One such enzyme is the alternative oxidase (AOX). When AOX is expressed, it enables its host to survive life-threatening conditions or, as in parasites, to evade host defenses. In vertebrates, this mechanism has been lost during evolution. However, we and others have shown that transfer of AOX into the genome of the fruit fly and mouse results in a catalytically engaged AOX. This implies that not only is the AOX a promising target for combating human or agricultural pathogens but also a novel approach to elucidate disease mechanisms or, in several cases, potentially a therapeutic cure for human diseases. In this review, we highlight the varying functions of AOX in their natural hosts and upon xenotopic expression, and discuss the resulting need to develop species-specific AOX inhibitors.

Enzymes act as catalysts for chemical reactions and thereby enable living organisms to adopt to their environment and convert energy from nutrients ready for all vital processes such as growth, maintenance, or reproduction [1–3]. A class of enzymes that has gained particular importance in recent years are the alternative oxidases (AOXs). AOXs belong to a subfamily of di-iron carboxylate proteins which include methane mono-oxygenases, ribonucleotide reductases, fatty acyl desaturases, and rubrerythrin [4]. Although AOX is a long-established redox partner of the mitochondrial electron transport chain (ETC) (Figure 1), its exact function in the process of energy conversion has for a long time been debated. Initially, AOXs were thought to be restricted to plant mitochondria, with their main function being to maintain TCA cycle activity under conditions when energy is provided by photosynthesis [5]. Seminal work later discovered that in plants the protein is induced by different forms of stress, such as inhibition of electron flux through the ETC, excessive ROS production or, importantly, during pathogen attack [6,7]. With the advent of technologies allowing DNA analyses and the availability of genome sequences, as well as the availability of AOX-specific monoclonal antibodies, AOX has subsequently been shown to be present in most kingdoms of life [5,8–10] implying a prominent role of this protein in cell viability. Furthermore, evidence emerged suggesting that AOX evolved very early in evolution, probably from a common ancestor of prokaryotes [8]. Amino-acid sequence analyses revealed that the AOX protein is highly conserved, irrespective of the host organism in which it is expressed. The broad presence within various phyla and subphyla has been confirmed by a variety of techniques including bioinformatic identification of its sequence, gene expression data upon amplification by reverse-transcriptase polymerase chain reaction (RT/PCR) and/or functional assays. Indeed respirometry confirmed the presence of cyanide-resistant respiration [9], validated using relatively non-specific AOX inhibitors, i.e. salicylhydroxamic acid (SHAM) and propyl gallates (PG) [11,12]. It is clear that AOX is not only critical for respiratory activity in plants but also necessary for the survival of numerous phytopathogenic and parasitic organisms. Of particular significance was the finding that several protist organisms which are exceptionally pathogenic to humans contain an AOX such as trypanosomes [13], Cryptosporidium parvum [14], Blastocystis hominis [15] and the microsporidia [16]. The full extent of AOX distribution in prokaryotes is still relatively uncertain, but it is becoming increasingly clear that AOX is also widespread among proteobacteria, including Novosphingobium aromaticivorans [17], members of the Vibrio genus [8,18,19] and Thiobacillus denitrificans [8]. In addition to its wide distribution in plants and prokaryotes, AOX is present in phytopathogenic fungi such as Blumeria graminis [20], Magnaporthe grisea [20,21], Septoria tritici [22], Ustilago maydis [20,23], Sclerotinia sclerotiorum [24,25] and Botrytis cinerea [20,26] to name but a few. Other notable species expressing AOX include the invertebrate marine worms Arenicola marina [27] and Sipunculus nudus [28], the bivalve Genkendia demissa [9], the Pacific oyster Crassostrea gigas [9], and the sea squirt Ciona intestinalis [29,30]. Since vertebrates lost AOX from their genome during evolution, Ciona intestinalis is now the closest known relative to humans still expressing a functional AOX enzyme. Remarkably, AOXs in fungi and Ciona appear to facilitate similar stress responses as has been described for plants, namely the conversion of cellular energy, redox homeostasis, and the control of mitochondrial ROS content [27,30,31] (Figure 1). This role in stress response makes AOX a valuable therapeutic target in both agriculture and environmental protection and possibly also in in human infectious diseases. For this very reason, AOXs in general and the one isolated from Ciona in particular have been utilized as a means to correct ETC impairments in humans, and we suggest that their use in vivo may even represent a potential novel therapeutic treatment for mitochondrial diseases [29,32–36].

The canonical electron transport chain (ETC) consists of four oxidoreductase complexes namely cI (NADH:ubiquinone oxidoreductase), cII (succinate:ubiquinone oxidoreductase), cIII (ubiquinol:cytochrome c oxidoreductase or the cytochrome bc1 complex) and cIV (cytochrome c oxidase).

Figure 1.
The canonical electron transport chain (ETC) consists of four oxidoreductase complexes namely cI (NADH:ubiquinone oxidoreductase), cII (succinate:ubiquinone oxidoreductase), cIII (ubiquinol:cytochrome c oxidoreductase or the cytochrome bc1 complex) and cIV (cytochrome c oxidase).

The transfer of electrons from respiratory substrates at cI and cII to oxygen at cIV is coupled to proton translocation across the inner mitochondrial membrane (IM) at cI, cIII, and cIV generating a H+ electrochemical gradient which is the main driving force (protonmotive force, pmf) for ATP production at cV (mitochondrial F1F0-ATPase). This coupled process is hence termed oxidative phosphorylation (OXPHOS). Oxidation of respiratory substrates, generated by the TCA cycle (tricarboxylic acid cycle or Krebs cycle), maintains the TCA and keeps upstream metabolic circuits operational. The main functions of the ETC include maintenance of cellular redox balance, heat production, carbon turnover for biosynthetic purposes and ATP generation by OXPHOS. If cIII and/or cIV are impaired, electron flux through the ETC is stalled, causing an arrest of the TCA cycle and upstream metabolic cycles, as well as proton translocation. This leads to redox imbalance, ROS production and eventually decreases OXPHOS. AOX (alternative oxidase) restores electron flux through the ETC and thus most mitochondrial and metabolic functions even though cIII and cV are bypassed.

Figure 1.
The canonical electron transport chain (ETC) consists of four oxidoreductase complexes namely cI (NADH:ubiquinone oxidoreductase), cII (succinate:ubiquinone oxidoreductase), cIII (ubiquinol:cytochrome c oxidoreductase or the cytochrome bc1 complex) and cIV (cytochrome c oxidase).

The transfer of electrons from respiratory substrates at cI and cII to oxygen at cIV is coupled to proton translocation across the inner mitochondrial membrane (IM) at cI, cIII, and cIV generating a H+ electrochemical gradient which is the main driving force (protonmotive force, pmf) for ATP production at cV (mitochondrial F1F0-ATPase). This coupled process is hence termed oxidative phosphorylation (OXPHOS). Oxidation of respiratory substrates, generated by the TCA cycle (tricarboxylic acid cycle or Krebs cycle), maintains the TCA and keeps upstream metabolic circuits operational. The main functions of the ETC include maintenance of cellular redox balance, heat production, carbon turnover for biosynthetic purposes and ATP generation by OXPHOS. If cIII and/or cIV are impaired, electron flux through the ETC is stalled, causing an arrest of the TCA cycle and upstream metabolic cycles, as well as proton translocation. This leads to redox imbalance, ROS production and eventually decreases OXPHOS. AOX (alternative oxidase) restores electron flux through the ETC and thus most mitochondrial and metabolic functions even though cIII and cV are bypassed.

Close modal

Here, we examine the various functions of AOX in selected host species and show how its expression may help to exacerbate the pathogenicity of the respective species. Furthermore, we discuss recent studies that use xenotopic expression of AOX to challenge the etiology of mitochondrial diseases in an attempt to provide a cure.

Numerous human pathogens express AOX, the activity of which facilitates a number of vital functions in the pathogen adaptation processes, which in turn can enhance their pathogenicity. One example is in Trypanosoma brucei (a causative agent of trypanosomiasis), where AOX plays a crucial role in adapting the metabolism to facilitate its life-cycle either in the tsetse fly (genus Glossina) or within the human bloodstream. Originally, it was thought that the bloodstream form of Trypanosoma brucei relied on AOX for ATP production by OXPHOS [37,38]. This seems counterintuitive considering that at this stage the pathogen is almost exclusively dependent on glycolysis for ATP production [39,40]. This notion is supported by the finding of a single mitochondrion with few flattened cristae in the bloodstream form [41] while a fully branched mitochondrial network is present in the procyclic insect form [39,40]. The switch in metabolism from oxidative phosphorylation to glycolysis and vice versa occurs rapidly and is likely regulated by substrate availability, eventually also leading to an altered mitochondrial proteome [42]. Whilst the procyclic form harbors all the ETC and Krebs cycle enzymes required for ATP production, the bloodstream form lacks the heme proteins required for respiratory complexes III and IV. In the bloodstream form of trypanosomes, two membrane-bound enzymes are expressed instead, i.e. glycerol-3-phosphate dehydrogenase and AOX, which together form a futile cycle rather than a fully functional ETC. Neither of these enzymes are protonmotive and hence cannot contribute to the protonmotive force (pmf), a prerequisite for mitochondrial ATP formation [39], suggesting that a contribution of AOX-mediated respiration to ATP production by OXPHOS in the bloodstream is an unlikely event. The actual role of AOX in the bloodstream form is rather to maintain redox balance, i.e. decreasing accumulating NADH levels thereby allowing glycolysis to continue to operate [40]. Interestingly, ATP inhibits trypanosomal AOX [40]. To avoid ATP-mediated inhibition of AOX and thus a stalling of the parasitic metabolism, the trypanosomal F1F0-ATPase hydrolyses ATP to pump protons across the inner mitochondrial membrane [40]. The paradoxic conclusion, therefore, is that the bloodstream form of Trypanosoma brucei indeed relies on AOX activity for ATP production, albeit via glycolysis rather than OXPHOS [37,38].

A further example of a pathogen in which a functional AOX has been discovered is the apicomplexan Cryptosporidium parvum [43]. C. parvum causes severe diarrheal disease (cryptosporidiosis), contributing significantly to mortality in children and immunocompromised individuals [44]. Unlike trypanosomes, C. parvum has a single-host life cycle in the intestine during which obligate asexual and sexual processes occur [45]. To date, no appropriate treatment exists. However, inhibition of AOX by SHAM and 8-hydroxyquinoline has been shown to inhibit the growth of this parasite in vitro [43] suggesting a vital role of AOX for its pathogenicity. The exact role of AOX in the pathogenicity is still unclear but as demonstrated for trypanosomes, its metabolism is largely glycolytic [46]. Interestingly, cryptosporidia have both aerobic and anaerobic metabolic pathways, which undoubtedly increases metabolic flexibility necessary for it to exist in a changing environment [44,47]. Notably, C. parvum possesses a degenerate mitochondrion and an incomplete F1-ATPase (only the α- and β-subunits are expressed [48]). It is thus tempting to speculate that these F1-ATPase subunits are expressed to control ATP content and consequently AOX activity thereby increasing both viability and pathogenicity of C. parvum.

A seemingly different situation is observed in Blastocystis hominis, a parasite generally considered to be an obligate anaerobe [15,49]. Intriguingly, it was discovered that B. hominis possess only mitochondria-like organelles with an incomplete ETC and a partial Krebs cycle alongside other metabolic pathways but does contain a functional AOX [49]. Since AOX uses oxygen as a final electron acceptor, its presence implies that oxygen fluctuations occur which represent a threat to the parasite. Indeed a recent study suggests that the blastocyst AOX can buffer oxygen fluctuations in the gut [15]. According to this study, AOX accepts electrons from ETC complex II (succinate:ubiquinone oxidoreductase/fumarate reductase) [15] and since neither complex II nor AOX are protonmotive, they form a futile cycle. Again, AOX cannot primarily contribute to ATP production by OXPHOS. A possible advantageous effect of expressing AOX could instead be that it prevents excessive mitochondrial ROS production, preserves the redox balance and/or influences the direction of important metabolic transport systems, such as the malate-aspartate shuttle (Figure 1). Eventually one or a combination of the proposed mechanisms may increase the pathogen's metabolic flexibility and thus the observed resilience of blastocysts.

AOX is also expressed in Acanthamoeba [50] which can cause a rare granulomatous amebic encephalitis involving the central nervous system that characteristically occurs in immunocompromised patients and nearly always causes death, or a painful keratitis resulting in blindness [51,52]. Acanthamoeba co-expresses two fully functional and competing quinol-oxidizing pathways, the phosphorylating canonical ETC and the cyanide-resistant AOX pathway [53,54]. Both pathways not only share the substrate (ubiquinol), but also interact kinetically, thereby directly or indirectly affecting the activity of the other. For instance, it was demonstrated that a direct protein–protein interaction exists between AOX and subunits of complex III [55]. Moreover, it has been shown that binding of GMP has a positive allosteric effect whilst the binding of ATP has a negative allosteric effect, and that both operate competitively, in that they overcome each other's effect [56,57]. Although the role of AOX in pathogenicity has not been fully elucidated, data suggests that AOX is induced in response to stress, e.g. during the encystment process [50], to buffer mitochondrial ROS generation [58] or to generate the heat needed to allow growth at low temperatures [59].

Other emerging human pathogens that broadly express AOX include the microsporidia, a group of obligate intracellular parasitic eukaryotes [16]. The identification of AOX in these parasites is particularly surprising because microsporidia were for a long time considered to be amitochondriate until remnants of mitochondria, now called mitosomes, were discovered [60]. A hallmark of microsporidia is that they have adapted perfectly to life as intracellular parasite reflected by a drastically reduced biological complexity such as loss of cell organelles and a massively shrunk genome [61–65]. Mitosomes lack a genome and genes encoding subunits of the ETC, the mitochondrial ATPase, and Krebs cycle enzymes appear not to exist [16,61]. Therefore, most microsporidia rely on glycolysis for ATP production, or are exclusively dependent on ATP import from the host as they even lack a fully functional glycolytic pathway [66]. Beside proteins required for the assembly of iron-sulfur clusters, mitosomes harbor a glycerol-3- phosphate dehydrogenase and, notably, an AOX [16,61]. The biological role of AOX in microsporidia likely depends on the stage in their respective life cycle [67], but it is reasonable to conclude that it plays a role similar to that in trypanosomes, namely to facilitate the recycling of reducing equivalents to keep glycolysis and/or other metabolic circuits operational. Whether an ATPase activity also exists to regulate AOX activity has yet to be determined.

AOX is also expressed in opportunistic human fungal pathogens, i.e. pathogens that do not represent a health threat unless an individual is immunocompromised or susceptible through other conditions, such as Candida albicans (candidiasis) [68,69], Cryptococcus neoformans (cryptococcosis) [70] and Aspergillus fumigatus (aspergillosis) [71]. The biological role of AOX in Candida albicans is of particular interest since its expression appears to be a reaction to different forms of increased stress resistance. In other words, the protein is expressed to protect against mitochondrial ROS generation during copper starvation [72], to decrease the susceptibility of the cell to antifungal drugs [73], or to prevent the loss of viability and cell wall rearrangements that eventually lead to increased recognition and uptake by macrophages [74]. A worldwide concern has recently emerged due to the discovery of Candida auris [75–77], a multidrug-resistant form of Candida, first isolated from a patient's auricle in Japan in 2009 [78]. This multidrug-resistant and primarily nosocomial pathogen has been associated with fungemia, ventriculitis, osteomyelitis, malignant otitis and otomastoiditis, and complicated forms of intra-abdominal infections, pericarditis, pleural effusions, and vulvovaginitis in a broad range of patients, including immunocompromised individuals [75–77]. Since its first description, C. auris infections have been reported in India, South Africa, Kuwait, Venezuela, Brazil, the United States, Colombia, Pakistan, Spain, Germany, Israel, Norway, Oman, and the United Kingdom [75–77]. The latter is the first description of a hospital-acquired transmission, which has led to a greater outbreak in a European country, adding to the notion that this multidrug resistant pathogen is capable of transmission in the health care setting potentially causing serious infections of a global concern. Particularly alarming is the finding that C. auris appears not only resistant to fluconazole [79], a drug commonly used to treat candidemia, but also to the other three main classes of antifungal drugs, azoles, echinocandins and polyenes [76].

Whilst metabolic flexibility is considered a key characteristic of Candida spp. enabling it to swiftly adapt to varying enviromental conditions and host niches [80], little is known about specific virulence factors of C. auris. Since AOX enables a metabolic stress resistance of C. auris, we have previously suggested that AOX inhibitors may become a safe and effective treatment approach and thereby could become a ‘game changer' in the clinic [69]. In support of this notion, functional studies with chemical or genetic silencing of AOX activity have provided further insights into the role of this enzyme in the virulence and pathogenicity of numerous human pathogens while bypassing the mode of action of conventional drugs. For instance, Paracoccidioides brasiliensis is one of the causative agents of paracoccidioidomycosis, and treatment of P. brasiliensis [81] with the non-specific AOX inhibitor SHAM prevented the transition from the mycelial to the yeast form, a critical step for host colonization. Likewise deletion of the C. neoformans [70] aox1 gene, which is induced at 37 °C presumably after sensing the human body temperature, led to a decrease in C. neoformans virulence in a murine model. Finally, silencing of the A. fumigatus AOX has led to a decreased viability in the presence of ROS and a higher rate of killing by macrophages [82].

As an increasing number of mechanisms of action of antibiotics and antifungal drugs against human pathogens are bypassed, the presence of multidrug-resistant organisms thus poses a serious threat to human welfare. AOX is emerging as a suitable pharmacological target. To date many compounds have been reported as inhibitors of the AOX, with the earliest and most used historically being the alkyl-gallate and hydroxamic acid derivatives from the 1980s (Table 1). Unfortunately, the low potency and low selectivity of most compounds have prevented their clinical use [83], which is now changing in large part due to advances in the biochemical and structural characterization of purified AOXs. The development of potent AOX inhibitors which are specific and exhibit broad-spectrum species sensitivity, act in the nM range, and have no or minimal side effects appears to be a viable approach. A number of such promising new molecules have been identified and tested both in our and other laboratories and are currently awaiting further clinical assessment. Latest advances and differences in the mode of action of individual inhibitors have recently been reviewed in detail elsewhere and will not be part of this review [84].

Table 1
AOX inhibitors or inhibitor classes
CompoundPubChem CIDStructureReferences
Benzoic acid derivatives 135  [123,234–236
Ascofuranone derivatives 6434242  [69,122,125,127,237
Decyl-Aurachin C 6439171  [238
Chloroquine 2719  [239
Coumarin derivatives 5650046  [240
Disulfiram 3117  [241
Ferulenol 54679300  [126
Gallic acid derivatives 61253  [11,235,238,242
Hydroxamic acid derivatives 66644  [235,242,243
Linoleic acid 5280450  [244–246
N-phenylbenzamide derivatives 3532749  [98,247
5-n-undecyl-6-hydroxy-4,7-dioxobenzothiazole 3016416  [244,248
CompoundPubChem CIDStructureReferences
Benzoic acid derivatives 135  [123,234–236
Ascofuranone derivatives 6434242  [69,122,125,127,237
Decyl-Aurachin C 6439171  [238
Chloroquine 2719  [239
Coumarin derivatives 5650046  [240
Disulfiram 3117  [241
Ferulenol 54679300  [126
Gallic acid derivatives 61253  [11,235,238,242
Hydroxamic acid derivatives 66644  [235,242,243
Linoleic acid 5280450  [244–246
N-phenylbenzamide derivatives 3532749  [98,247
5-n-undecyl-6-hydroxy-4,7-dioxobenzothiazole 3016416  [244,248

Structures column depicts core inhibitor structures, with R groups indicating positions of functional group manipulation used within the reference list.

The rising presence of fungal infections is also recognized as a global threat to food safety in agriculture, which endangers both ecosystem and human health [85]. Fungal infections which have a negative impact on food quality are well-described for cereal crops. For instance, Fusarium graminearum in wheat [86], Magnaporthe oryzae in rice [87], Ustilago maydis in maize [88], and Sclerotinia sclerotiorum in potatoes [89] and soybean [90]. Such infections are also devastating non-cereal crops, such as Moniliophthora perniciosa in cocoa [91,92], Mycosphaerella fijiensis in banana [93], Venturia spp. in apple [94], and various others [95]. Disturbingly, a simultaneous occurrence of severe epidemics in all five cereals would most certainly have a catastrophic impact on food security, as there would not be enough food available for the world's growing population. Analagous to the aforementioned human pathogens, a great number of phytopathogenic fungi also use AOX to escape stress response mechanisms thereby enhancing pathogenicity [21,25,96,97].

Unfortunately, only very few studies have addressed the biochemical differences between the various AOX enzymes and the different roles that fungal AOXs may play in plant pathogenesis. Hitherto, it has been demonstrated that in the cocoa pathogen M. perniciosa, AOX is necessary for spore germination and host infection [98] while M. grisea virulence in barley leaves appears to use different mechanisms [21]. A more indisputable role, and perhaps more relevant from an agronomic perspective, is attributed to the importance of AOX as a source of resistance to fungicides used as crop protection agents. The best-studied case is the inhibition of complex III (cytochrome bc1 complex) of the ETC by strobilurins and other so-called quinone-outside inhibitor (QoI) fungicides [99]. Due to the functional bypass of the cytochrome bc1 complex, AOX can restore mitochondrial functions (Figure 1) and sustain fungal growth. Unsurprisingly it has been demonstrated that the inhibition of both pathways, the canonical ETC and the alternative pathway, is a promising approach to control several phytopathogens of agronomical importance. For instance, both Septoria tritici, a fungus that causes major leaf spot diseases in wheat [96], and Gaeumannomyces graminis var. tritici [100], the wheat ‘take-all' fungus, express AOX specifically when treated with respiratory inhibitors, thereby facilitating a fungicide strobilurin-resistant respiratory pathway. Similar results have been observed with Magnaporthe grisea [101], Sclerotinia sclerotiorum [25], M. perniciosa [97,98], Venturia inaequalis [102,103] and Microdochium majus [104]. QoI fungicides have been extensively used in the agriculture since the 1990's especially following the discovery of strobilurin (a natural compound isolated from the fungus Strobilurus tenacellus) and the development of synthetic analogs with enhanced field performance [99]. Azoxystrobin became the top-selling fungicide as little as four years after its launching and within 10 years strobilurins represented 20% of worldwide sales [105]. An alarming recent observation, however, is the increasing identification of field isolates resistant to strobilurins which at least in part can be attributed to AOX expression. Moreover, it has been shown in M. grisea that AOX allows fungal survival long enough for the emergence of point-mutations in the cytochrome bc1 complex, such as the change of glycine 143 to alanine (G143A), thereby rendering azoxystrobin completely ineffective [106,107]. AOX has also been associated with resistance to fungicides using other modes of action such as procymidone, fluconazole and the imidazoles miconazole, econazole and ketoconazole [73,108]. Mechanistically this may be due to the contribution of AOX to metabolic and redox homeostasis under stress conditions, and/or ROS scavenging. As with QoI fungicides, it is likely that new resistant mutants will evolve after prolonged exposure to those fungicides. Since fungicides play a key role in the control of diseases in crops and are thus a vital part of ensuring global food supply and safety, the increasing resistance to fungicides not only threatens the efficacy of individual fungicides but is also a major cause of reductions in crop yield and subsequent environmental degradation.

A further aspect which deserves attention is the possible role of AOX in the production of harmful mycotoxins by food colonizing fungi which can occur in the field or during post-harvest storage. In addition to AOX allowing fungal growth as discussed above, which by itself can increase mycotoxin incidence, this enzyme may contribute to the metabolic pathways associated with mycotoxin production. There is a known positive correlation between AOX expression in Acremonium chrysogenum [109] and Aspergillus nidulans [110,111] and the production of cephalosporin C and sterigmatocystin, respectively. This implies that AOX may have a broad role in fungal secondary metabolites production, some of which can have important effects on the human welfare. For instance, sterigmatocystin is a precursor to a skin-permeable carcinogen aflatoxin B1 produced mainly by Aspergillus flavus but also by other Aspergillus species [112]. Since mycotoxins can cause serious health issues including cancer, immune system suppression, liver failure and congenital diseases, regulatory bodies around the world have created strict guidelines for mycotoxin contamination levels. In the European Union the maximum accepted level of aflatoxin B1 in several nuts and cereals is precautionarily restricted to 2 to 12 parts per billion (ppb), but it could be as low as 0.1 ppb in foods destined to infants (European Commission Regulation No 1881/2006 Dec 19, 2006; https://t1p.de/8t67). Health and economic losses, however, may not only occur through direct contact with human beings but also by contamination and killing of poultry and livestock. In addition, many mycotoxins are highly stable meaning that they may remain present even after the toxin-producing fungus has been eliminated during the food processing [113–115].

Because AOX is present in both the fungal pathogen and the infected plant, putative AOX inhibitors that can be approved as fungicides must have exceptional specificity to inhibit the pathogen without harming the host, and this specificity must be accompanied by high efficacy and (food) safety. A key challenge in understanding the structure-function relationship of AOX has been the identification of its substrate- and inhibitor-binding site and its mechanisms of catalytic action. A detailed knowledge of the nature of this binding site is thus important, since it will reveal if there is a common architecture for the binding of its substrate or an inhibitor, thereby providing unprecendented insights into differential mechanisms of binding. Only such knowledge will allow the rational design of phytopathogenic and anti-parasitic drugs that target AOX in a species-specific manner.

AOXs are self-folding di-iron carboxylate proteins harboring four highly conserved ExxH motifs, a characteristic of the common iron binding sequences found in non-heme iron binding proteins [116,117]. Initial AOX models proposed that a four-helix bundle acts as a scaffold to hold the iron core, the di-iron center being ligated by three glutamates, two histidines and one aspartate comparable to the binding sequences of methane monooxygenase or ribonucleotide reductase (Figure 2). These initial models also included two additional transmembrane helices with hydrophobic sections in the mitochondrial intermembrane space that were initially thought to function as substrate binding domains. Further modeling removed the transmembrane region in favor of a simpler association with the membrane based on hydrophobicity plots, and the proposed iron binding aspartate residue was replaced by a glutamate to evenly space the di-iron core binding across all four helices of the bundle [118,119]. Yet, it took another 13 years for the first AOX crystal structure to become available, largely due to the problematic nature of purification of membrane bound proteins in a stable form. Initial crystallization attempts of the trypanosomal AOX (often referred to as TAO) using digitonin were unsuccessful [120], but later improvements using n-octyl-β-d-glucopyranoside [121] increased both activity and stability of the protein, allowing for generation of crystal structures to a resolution of 2.3 Å [122]. Notably, the current structure lacks a section of the N-terminal region due to its ‘unstructured' nature which is thought to be the mitochondrial targeting sequence as its removal does not affect functionality of the recombinant protein [123]. Although attempts to apply these techniques to AOXs from other species [124,125] have been successful in generating stable purified protein, they nevertheless have failed to generate refracting crystals of sufficient quality for crystallography. Despite this shortcoming, most structural predictions previously made were proven to be correct, with the overall structure of TAO being that of a di-iron protein consisting of six long α-helices and four shorter α-helices, the metal core being ligated by four glutamates residues, E123, E162, E213 and E266, and two histidine residues, H165 and H269. A striking feature of the protein is its apparent crystallization in a dimeric form, with each of the monomers aligning in an antiparallel fashion. Given the conserved nature of the dimer interface across all AOX species and the symmetrical nature of the inter-monomer interaction of the unstructured N-terminal arm this arrangement is unlikely to be an artifact of the crystallization procedure, with the protein behaving as a structural dimer.

Crystal structure (PDB:5ZDR) of the alternative oxidase (AOX) dimer.

Figure 2.
Crystal structure (PDB:5ZDR) of the alternative oxidase (AOX) dimer.

(A and B) Surface model of the dimer showing hydrophobic residues in orange, hydrophilic in grey. (B) shows the entrance to the hydrophobic cavity in the membrane associating side of the protein, with the inhibitor depicted as green sticks and a leucine ‘gate', which controls inhibitor orientation, colored blue. (C) Helical structure of the dimer in the same orientation as A, with monomers colored magenta and yellow, respectively. Inhibitor can be seen as green sticks to show position of the hydrophobic cavity between helices 1 and 4 leading directly to the di-iron core, which shown as orange spheres. (D) The binding pocket within the protein, showing all amino acids within 5 Å of the inhibitor. Key hydrogen bonding between the phenol, Arg-118 and Thr-219 shown as yellow dotted line (inlet).

Figure 2.
Crystal structure (PDB:5ZDR) of the alternative oxidase (AOX) dimer.

(A and B) Surface model of the dimer showing hydrophobic residues in orange, hydrophilic in grey. (B) shows the entrance to the hydrophobic cavity in the membrane associating side of the protein, with the inhibitor depicted as green sticks and a leucine ‘gate', which controls inhibitor orientation, colored blue. (C) Helical structure of the dimer in the same orientation as A, with monomers colored magenta and yellow, respectively. Inhibitor can be seen as green sticks to show position of the hydrophobic cavity between helices 1 and 4 leading directly to the di-iron core, which shown as orange spheres. (D) The binding pocket within the protein, showing all amino acids within 5 Å of the inhibitor. Key hydrogen bonding between the phenol, Arg-118 and Thr-219 shown as yellow dotted line (inlet).

Close modal

One of the most important structural features for the protein with respect to both its species-specific functionality and inhibition is the hydrophobic substrate cavity. This cavity is located between the α1 and α4 helices and leads directly to the di-iron core from the proposed membrane binding region [122] providing quinol, the substrate for AOX, direct access from the lipid bilayer to the reactive metal core. Whilst there is no crystal data with the substrate bound, numerous structures are available with inhibitors, such as colletochlorin B, AF2779OH or ferulenol, bound within the cavity [122,126], all of which have been shown to behave as transition state analogs indicating that the binding-site is highly likely to be the quinol binding site. Moreover, mutational analyses of the conserved residues lining the cavity has highlighted the significant changes in quinol kinetics that accompany slight changes to the amino acid composition, demonstrating the importance of this region to substrate/inhibitor affinity. Consequently, this region is of particular importance for the development of inhibitors for the control of pathogenic diseases mentioned above.

The trypanocidal activity of ascofuranone has been known since the mid 1990's, however, its low yielding chemical synthesis and low efficacy in blood clearance without the need for significant concentrations of glycerol led to a loss of scientific interest [127–129]. More recently there has been a significant resurgence in interest with this compound and its analogs, with numerous structure-function studies [130,131] and QSAR modeling work from the TAO crystal structure [132]. These studies highlighted the importance of the aldehyde functionality and the need for a hydrogen bond acceptor in the para position from the hydrophobic chain. In addition to ascofuranone-like compounds, the gallates are also considered to be effective inhibitors in high nanomolar region (Table 1). Recent developments in the gallate analogs (4-hydrobenzoates) show significant improvements in efficacy with the addition of TPP+, a charged lipophilic cation, aiding in targeting the compound to the mitochondria. An extensive review of all alternative oxidase inhibitors known to date, with a particular focus on their efficacy against TAO, however, can be found elsewhere [84].

Although, there is crystal data for the trypanosomal variant, significant homology modeling has been performed on AOXs from fungal [125] and human pathogens [15,133]. Of note, in most of these variants studied to date, the predicted cavity size appears decreased in comparison with TAO. Such a decrease in cavity size would significantly decrease the ability of compounds to inhibit the enzyme. This is especially apparent with ascochlorin, which, while active at nanomolar concentrations in TAO has its efficacy reduced 1000-fold when tested against the AOX from M. perniciosa [125]. Thus, effective inhibition of TAO has been widely studied, but that such inhibitors are also effective against fungal targets is still under investigation.

In addition to the role of AOX in enabling human and fungal pathogens to seek opportunistic ways of causing diseases, there are also perspectives for the use of AOX to challenge disease paradigms of human pathologies. One such emerging opportunity is within the area of AOX gene therapy, in attempts to overcome mitochondrial dysfunctions [134]. However, besides the potential therapeutic benefits of xenotopic AOX expression [35,36], one must be aware of possible side effects and pitfalls [29]. An obvious threat to viability could be that mitochondrial ATP production decreases due to the catalytic engagement of AOX and the consequent bypass of two of the three energy-conserving sites in the ETC (Figure 1). This could counteract putative beneficial effects such as the restoration of electron flux through the ETC, the redox balance, or heat production. Indeed, we found that xenotopic expression of AOX was deleterious in two independent mouse disease models, namely a model of mitochondrial myopathy based on skeletal muscle-specific ablation of the complex IV subunit COX15 [135] and in the postischemic heart [136]. In both, impaired signaling and maladaptive organ remodeling appeared to be the consequence of AOX expression. Yet, AOX is currently being considered for gene therapy in mitochondrial diseases ranging from Alzheimer's disease to emphysema and endotoxemia [137–139] after previous studies indicated that the symptoms of a disease may be alleviated or the condition can be completely cured. Notably, AOX gene therapy does not necessarily restore ATP levels [135]. Instead, it may reactivate respiratory complex I, thereby indirectly re-establishing the mitochondrial membrane potential which itself is indispensable for ATP production [140]. To date, AOX has been shown to efficiently bypass chemical inhibition of the ETC in flies [141] and mice [142,143], to rescue from lethality in cytochrome c oxidase (COX) deficient flies [141,144], and to partially rescue locomotor degeneration in flies exhibiting Parkinson's-like symptoms [141,145]. More recently it was shown that AOX can mitigate beta-amyloid production [137] (a causative agent of Alzheimer's disease) and toxicity in the fly and human cultured cells, and AOX also decreased lethality in an endotoxemia model of the mouse [138] and alleviated detrimental lung remodeling upon toxic smoke exposure [139].

Parkinson's disease (PD) is a neurodegenerative disease with an adult-onset that is characterized by the degeneration of dopaminergic neurons in the pars compacta of the substantia nigra. The dopaminergic neurons communicate with neurons of the basal ganglia by dopamine as a neurotransmitter and collectively are responsible for the fine tuning of movements. Consequently, clinical hallmarks of PD are slowness of movement (bradykinesia), rigidity, tremor, and the inability to maintain an upright position (postural instability). A mitochondrial involvement in the pathology has been postulated, since pharmacological inhibition of OXPHOS complexes such as the respiratory complex I inhibitor and neurotoxin l-methyl-4-phenyl-l,2,3,6-tetrahydropyridine (MPTP) [146–148] and the potent respiratory complex IV inhibitors carbon monoxide (CO) and cyanide produce symptoms that mimic PD to a great extent [149,150]. Furthermore, deletions in mitochondrial DNA have been found in the striatum of PD patients, as well as defects in OXPHOS due to mutations in nuclear encoded subunits of respiratory complexes I and IV in isolated mitochondria from muscle biopsies of PD patients [151] supporting the assumption of a mitochondrial involvement. Together, these findings have nurtured the idea of PD being a mitochondrial disorder.

To ultimately identify the position of mitochondrial OXPHOS dysfunction in the hierarchy of PD, the combination of animal models and genetic tools became inevitable. One such approach used the fly model, Drosophila melanogaster, in which the catalytic subunit of the mitochondrial DNA polymerase was down-regulated by RNAi technology [145]. Expectedly, this down-regulation disturbed ETC function. The use of cell-specific drivers allowed an ablation in specific neurons, e.g. dopaminergic neurons. This resulted in adult onset of progressive dopaminergic neurodegeneration and premature ageing with age-related motor deficits. Interestingly, the xenotopic expression of AOX, and partial restoration of PINK1/parkin signaling or Drp1, rescued the observed phenotype [145]. AOX restored ATP levels which argues for an increased electron flux through complex I, and more importantly, preventing dopaminergic neurodegeneration. This is a strong direct indication for a causal involvement of OXPHOS dysfunction in PD and subsequently a novel treatment option directly on-site of dysfunction for a disease that so far escaped a proper treatment approach.

Further evidence for a positive effect of AOX on the course of PD arose in another Drosophila disease model, the dj-1b mutant [141]. DJ1 is a highly conserved protein that was previously identified as a disease gene causing autosomal recessive early-onset parkinsonism [152]. The exact mechanism of DJ1 action in the pathology is not fully understood, but DJ1 was thought to affect ROS signaling thereby inducing apoptosis. Interestingly, the exceptional susceptibility of dopaminergic neurons to mitochondrial ROS contrasts to a relatively low mitochondria mass [153], although this may actually reinforce an OXPHOS dependence. In support of this, loss of respiratory complexes I and IV was detected in cultured neurons and forebrain samples from DJ1 knockout mice, suggesting energetically inefficient mitochondria and increased ROS production as a likely cause [154]. Another recent observation provided an additional perspective. It was found in brain mitochondria from DJ1 knock-out mice that respiration-dependent hydrogen peroxide consumption was increased [155], rather than decreased, which was attributed to increased activity of mitochondrial thioredoxin and glutaredoxin, and higher total GSH levels. A possible explanation for the seemingly counterintuitive finding may come from a study using an alternative genetic model, Dictyostelium discoideum. The results showed that DJ1 controls extramitochondrial functions such as phagocytosis and, to a lesser extent, pinocytosis under physiologic conditions [156]. Loss of function may thus only indirectly affect mitochondrial respiration which, nevertheless, may phenocopy conditions of OXPHOS deficiency. Of course, it is equally possible that different forms of PD exist based on entirely independent etiologies that, nevertheless, result in very similar disease states again exemplifying the complexity of PD. Irrespective of the precise underlying mechanism, AOX rescued dj-1b fly mutant [141], which is consistent with the idea of a disrupted ETC and subsequent up-regulation of mitochondrial ROS levels.

Alzheimer's disease (AD) is the most common form of dementia based on a chronic neurodegenerative process that worsens over time [157,158]. A clinical hallmark of AD is the difficulty to remember recent events whilst older information remains temporarily accessible. Over time AD patients virtually lose their personality. Since vital body functions also decline the mean survival time from the onset of dementia ranges only from 3.3 to 11.7 years [159]. The etiology of AD is only incompletely understood. It is known that the risk of developing AD rises with age and thus the lifestyle may be an etiologic factor. Although, due to the long incubation time, it is difficult to identify cause-and-effect relations it is assumed that cardiovascular fitness is a protective factor against a cognitive decline [160,161]. Besides life-style factors, there is a documented genetic risk for late-onset and early-onset AD in carriers of apolipoprotein E4 (APOE4) [162]. While APOE is synthesized upon neuronal damage or stress and thus represents a regular repair mechanism, its conformation causes the generation of adverse neurotoxic fragments via proteolysis [163]. APOE4-derived fragments have a negative effect on the cytoskeleton and, of note, also cause mitochondrial dysfunctions. Furthermore it has been suggested that APOE is involved in clearance of amyloid beta (Aβ) peptide from the brain and that different APOE alleles may contribute differentially to the clearance [164]. Interestingly, while it is widely assumed that toxicity arises mainly from the extracellular insoluble protease-resistant fibrils, this concept has been disputed [165]. It appears that oligomeric Aβ peptides can disrupt the integrity of cell membranes [166] and, indeed, injected amyloid peptides impair synaptic transmission and induce cell death and cognitive decline if able to aggregate [167]. This would make the process of aggregation and subsequent membrane interaction the toxic event and not the passive presence of fibrils [168]. Toxicity may thus be primarily due to detergent-like properties on membranes and disrupted ionic homeostasis [169,170]. This mechanism may equally underlie the observed sensitivity of mitochondria to amyloid peptides [171]. Indeed, Aβ peptides impair cellular autophagy including mitophagy [172]. Interestingly, Aβ peptides are imported into mitochondria using the classical import machinery [173,174], which then has a number of effects [175]. For instance, an impairment of respiratory complex IV (cytochrome c oxidase) [171,176] has been demonstrated to be due to direct interaction in association with Aβ accumulation [177,178] which supports postmortem observations in brains of AD patients [179,180]. The proposed Aβ-induced disruption of electron transport through the ETC may also explain an observed decrease in the TCA cycle enzyme α-ketoglutarate [181] and an increase in ETC-related ROS production [182,183].

Despite the progress in understanding underlying disease mechanisms, no treatment has thus far has been developed which could stop or reverse the loss of cognitive functions. Based on current knowledge, however, mitochondria appear as a valuable target to shift fate [184,185]. This was recently tested in different models of AD pathology [137]. Specifically, AOX acted as antioxidant and thus decreased Aβ40 production which was induced in human HEK293-derived cells by the cytochrome bc1 complex inhibitor antimycin A. Furthermore, AOX partially alleviated a neurodegenerative phenotype, i.e. a decreased lifespan in fly models expressing human Aβ peptides specifically in neurons. This improvement in survival was associated with a decrease in marker expression indicative of oxidative stress pointing towards disruption of electron flow through the ETC and thus induced ROS production as causative factor for the development of AD.

Sepsis is defined as ‘life-threatening organ dysfunction caused by a dysregulated host response to infection' [186]. Sepsis-related organ failure has a high prevalence of morbidity and mortality, particularly associated with age, frailty, and severe comorbidities [187,188]. In the United States, for example, sepsis affects ∼1.7 million adults each year, and is thought to account for more than 250 000 deaths [187,189]. It has been estimated that sepsis occurs in up to 50% of hospital admissions eventually leading to death [187,189,190]. Unfortunately, sepsis survivors have also an increased risk of developing cognitive impairment and cardiovascular disease resulting in a shorter life expectancy [191]. Despite decades of research there is still no causal treatment option for sepsis [192]. Part of the reason for the failure of drug development is that molecular mechanisms underlying the development of sepsis are still ill-defined likely due to the complexity and systemic nature of the condition. One concept proposes that a so-called cytokine storm systemically dysregulates several signaling cascades thereby affecting all organs simultaneously [193].

The complex network of dysregulated signaling cascades observed during sepsis affects all levels of cellular metabolism and hence organs which rely heavily on OXPHOS, such as the heart, are particularly sensitive [194]. Recently, evidence emerged that mitochondria act not as helpless bystanders of a greater disturbance but are rather active players in the process. Due to conflicting results, the exact mode of impairment and the hierarchy of mitochondria in the pathology are still under debate [138,195–197], so-called mitochondrial interventions (often referred to as metabolic resuscitation) [198] have been identified as a potential therapeutic approach. Seemingly paradoxical, mitochondrial biogenesis is not necessarily able to recover mitochondrial respiration in sepsis [196]. In the liver, this may be explained by the stress-induced expression of nitric oxide (NO), carbon monoxide (CO) and hydrogen sulfide (H2S) [199], all three being potent inhibitors of complex IV [200–202]. Although therapeutically valuable at low or moderate levels [203,204], nitric oxide, carbon monoxide and hydrogen sulfide levels are in the context of pathologies significantly increased [202,205,206]. Given that hepatic parenchymal injury is critical for the development of the sepsis-related multi-organ failure, and respiratory disruption is the leading mitochondrial impairment in sepsis, it seems possible that increasing mitochondrial mass alone is not sufficient to overcome such potent inhibitors. Instead, mitochondrial biogenesis may have the opposite effect since more ROS may originate from a more dysfunctional ETC, possibly even accelerating multi-organ failure and septic shock. Conversely, protective effects have been described for antioxidants targeted specifically to the mitochondrial compartment, measured as a decrease in cytokine levels and oxidative stress while markers of mitochondrial and organ functions improve [207].

Cardiac contractile dysfunction is another important component towards fatality during sepsis. While systolic dysfunction paradoxically predicts a better outcome, diastolic dysfunction is a common and major predictor of mortality in sepsis [208]. Again, impaired OXPHOS was identified as the driving trigger of sepsis-associated myocardial depression and specifically myocardial complex IV has been shown to be impaired [209]. Supplementation with exogenous cytochrome c partially restored complex IV activity and improved cardiac function [210,211]. Interestingly, myocardial cell death is rare in sepsis and contractility is fully restored in survivors, arguing for a functional rather than structural impairment [212,213]. A phenomenon which is compatible with exposure to high levels of respiratory inhibitors as being the underlying mechanism.

Overcoming mitochondrial respiratory impairment has been identified as the key outcome of clinical and biochemical organ failure. Published data indicates that AOX expression provides protection against septic shock, based on mitochondrial protection or acceleration of the recovery process through stimulation of mitochondrial activity, e.g. in a mouse model of LPS-induced endotoxemia [138]. Mechanistically, it was demonstrated in isolated bone-marrow-derived macrophages that AOX prevents succinate-driven reverse electron transport (RET) and dramatically decreases mitochondrial ROS. Such results suggest that AOX may have a clinical application in the control of sepsis in addition to other mitochondrial diseases highlighted above.

Chronic obstructive pulmonary disease (COPD) describes a number of conditions of chronic inflammatory lung disease which obstruct the airflow [214]. Most prominent conditions contributing to COPD are emphysema and chronic bronchitis. Emphysema describes a pathology in which the terminal ends of the airways (alveoli) are gradually destructed, which ultimately impairs vital gas exchange. The greatest risk factors for the development of emphysema are cigarette smoke but also other toxic gases as well as particulate matter and household air pollution [215–218]. The lack of the endogenous ability to regenerate and the almost complete absence of therapeutic measures to halt or reverse tissue damage make COPD an incurable disease whose progression can currently at best be slowed [219].

One reason for the observed cigarette smoke toxicity may be its content of detrimental oxidants [220] which cause local inflammation and consequently trigger long-term lung remodeling [221,222]. The identification of the main toxin may prove difficult since cigarette smoke is a mix of more than 5000 toxic and carcinogenic compounds [223]. Importantly, from a mitochondrial perspective, is that cigarette smoke contains carbon monoxide and hydrogen cyanide both of which are potent complex IV inhibitors. These two compounds alone potentially disrupt the electron flow through the ETC impairing ATP production, redox homeostasis, and ROS production. Indeed, previous studies have provided substantial evidence for a smoke-induced respiratory impairment [224–227]. To directly study the involvement of mitochondrial ETC dysfunction and its hierarchy in the development of COPD, we tested the role of AOX in a murine model of smoke-induced emphysema [139]. In this study, AOX attenuated the degree of lung tissue destruction and loss of lung function. To identify underlying mechanisms, embryonic fibroblasts isolated from mouse were treated with cigarette smoke condensate (CSC) [139], a well-established surrogate for cigarette smoke [228–230]. Expression of AOX increased cellular viability and decreased mitochondrial ROS level upon CSC exposure. An unexpected finding of our study was that the so-called early-phase inflammatory response, i.e. the acutely induced infiltration of macrophages and neutrophils, and the expression of cytokines upon smoke exposure was unaffected. This is surprising since a pro-inflammatory cytokine signature was considered a trigger for adverse lung remodeling [231–233]. Apparently, this assumption needs to be revisited.

Bypassing a blocked ETC by AOX increases the viability of organisms. This mechanism helps plants, fungi and other organisms adapt to their ever-changing environment, but is also used by pathogens to evade the host's immune response (Figure 3). To protect the host while attacking a pathogen, several AOX inhibitors have been developed and are currently being tested. The fact that the AOX is naturally absent from humans can facilitate the design of targeted drugs with fewer side effects. The use of such compounds may become a ‘game changer' for many applications, whether as agrochemicals or as antibiotics to combat multidrug-resistant organisms. Interestingly, many human diseases are based on a blockade of the ETC. We and others have shown that xenotopic expression of AOX, a cross-species gene transfer, can challenge disease paradigms in appropriate disease models and, as such, AOX expression has the potential to cure lethal so-called mitochondrial diseases. However, the ultimate goal must be to optimize AOX for better applicability with tailored biochemical properties.

AOX is a naturally evolved stress response gene that restores electron flux through the ETC when the canonical pathway is blocked.

Figure 3.
AOX is a naturally evolved stress response gene that restores electron flux through the ETC when the canonical pathway is blocked.

By facilitating a bypass of the canonical ETC, it enables oxygen consumption and heat production, prevents mitochondrial ROS production, indirectly supports ATP generation by OXPHOS, and maintains cellular redox balance and TCA cycle turnover. Plants, fungi and other natural hosts use AOX to increase their metabolic flexibility and stress resistance. Pathogens utilize AOX to increase pathogenicity. Since a number of human diseases are based on ETC impairment, it has been proposed that xenotopic expression of AOX can be used as a therapeutic in human diseases to eventually alleviate disease-related side effects.

Figure 3.
AOX is a naturally evolved stress response gene that restores electron flux through the ETC when the canonical pathway is blocked.

By facilitating a bypass of the canonical ETC, it enables oxygen consumption and heat production, prevents mitochondrial ROS production, indirectly supports ATP generation by OXPHOS, and maintains cellular redox balance and TCA cycle turnover. Plants, fungi and other natural hosts use AOX to increase their metabolic flexibility and stress resistance. Pathogens utilize AOX to increase pathogenicity. Since a number of human diseases are based on ETC impairment, it has been proposed that xenotopic expression of AOX can be used as a therapeutic in human diseases to eventually alleviate disease-related side effects.

Close modal

MS states that he is a shareholder in a spin-off company to develop AOX-based therapeutics. ALM is a shareholder and scientific director of a company that develops AOX inhibitors and aims to improve the protein properties of AOX.

Anthony Moore: Conceptualization, Writing - original draft, Writing — review and editing. Marten Szibor: Investigation, Writing - original draft, Writing — review and editing. Christina Schenkl: Investigation, Writing — original draft, Writing — review and editing. Luke Young: Investigation, Writing — original draft, Writing — review and editing. Mario R.O. Barsottini: Investigation, Writing — original draft, Writing — review and editing.

ALM gratefully acknowledges support from the Biotechnology and Biological Sciences Research Council (BBSRC) (grants BB/L022915/1 and BB/NO10051/1). C.S. thanks the Deutsche Forschungsgemeinschaft (DFG) for financial support (grant SCHE2126/1-1).

AD

Alzheimer's disease

AOX

alternative oxidase

APOE4

apolipoprotein E4

COPD

chronic obstructive pulmonary disease

CSC

cigarette smoke condensate

ETC

electron transport chain

PD

Parkinson's disease

SHAM

salicylhydroxamic acid

1
Kraut
,
J.
(
1988
)
How Do enzymes work?
Science
242
,
533
540
2
Arcus
,
V.L.
and
Mulholland
,
A.J.
(
2020
)
Temperature, dynamics, and enzyme-catalyzed reaction rates
.
Annu. Rev. Biophys.
49
,
1
18
3
Ribeiro
,
A.J.M.
,
Tyzack
,
J.D.
,
Borkakoti
,
N.
,
Holliday
,
G.L.
and
Thornton
,
J.M.
(
2020
)
A global analysis of function and conservation of catalytic residues in enzymes
.
J. Biol. Chem.
295
,
314
324
4
Moore
,
A.L.
,
Shiba
,
T.
,
Young
,
L.
,
Harada
,
S.
,
Kita
,
K.
and
Ito
,
K.
(
2013
)
Unraveling the heater: new insights into the structure of the alternative oxidase
.
Annu. Rev. Plant Biol.
64
,
637
663
5
Rogov
,
A.G.
,
Sukhanova
,
E.I.
,
Uralskaya
,
L.A.
,
Aliverdieva
,
D.A.
and
Zvyagilskaya
,
R.A.
(
2014
)
Alternative oxidase: distribution, induction, properties, structure, regulation, and functions
.
Biochemistry
79
,
1615
1634
6
Simons
,
B.H.
,
Millenaar
,
F.F.
,
Mulder
,
L.
,
Loon
,
L.C.V.
and
Lambers
,
H.
(
1999
)
Enhanced expression and activation of the alternative oxidase during infection of arabidopsis with Pseudomonas syringae pv tomato
.
Plant Physiol.
120
,
529
538
7
Maxwell
,
D.P.
,
Nickels
,
R.
and
McIntosh
,
L.
(
2002
)
Evidence of mitochondrial involvement in the transduction of signals required for the induction of genes associated with pathogen attack and senescence
.
Plant J.
29
,
269
279
8
McDonald
,
A.E.
and
Vanlerberghe
,
G.C.
(
2006
)
Origins, evolutionary history, and taxonomic distribution of alternative oxidase and plastoquinol terminal oxidase
.
Comp. Biochem. Physiol. Part D Genomics Proteomics
1
,
357
364
9
McDonald
,
A.E.
,
Vanlerberghe
,
G.C.
and
Staples
,
J.F.
(
2009
)
Alternative oxidase in animals: unique characteristics and taxonomic distribution
.
J. Exp. Biol.
212
,
2627
2634
10
McDonald
,
A.E.
(
2009
)
Alternative oxidase: what information can protein sequence comparisons give us?
Physiol. Plantarum
137
,
328
341
11
Siedow
,
J.N.
and
Girvin
,
M.E.
(
1980
)
Alternative respiratory pathway; its role in seed respiration and its inhibition by propyl gallate
.
Plant Physiol.
65
,
669
674
12
Young
,
P.G.
(
1983
)
The SHAM-sensitive alternative oxidase in Tetrahymena pyriformis: activity as a function of growth state and chloramphenicol treatment
.
Microbiology
129
,
1357
1363
13
Chaudhuri
,
M.
and
Hill
,
G.C.
(
1996
)
Cloning, sequencing, and functional activity of the Trypanosoma brucei brucei alternative oxidase
.
Mol. Biochem. Parasitol.
83
,
125
129
14
Suzuki
,
T.
,
Hashimoto
,
T.
,
Yabu
,
Y.
,
Kido
,
Y.
,
Sakamoto
,
K.
,
Nihei
,
C.
, et al (
2004
)
Direct evidence for cyanide-insensitive quinol oxidase (alternative oxidase) in apicomplexan parasite Cryptosporidium parvum: phylogenetic and therapeutic implications
.
Biochem. Biophys. Res. Commun.
313
,
1044
1052
15
Tsaousis
,
A.D.
,
Hamblin
,
K.A.
,
Elliott
,
C.R.
,
Young
,
L.
,
Rosell-Hidalgo
,
A.
,
Gourlay
,
C.W.
et al (
2018
)
The human gut colonizer blastocystis respires using complex II and alternative oxidase to buffer transient oxygen fluctuations in the gut
.
Front. Cell Infect. Microbiol.
8
,
371
16
Williams
,
B.A.P.
,
Elliot
,
C.
,
Burri
,
L.
,
Kido
,
Y.
,
Kita
,
K.
,
Moore
,
A.L.
et al (
2010
)
A broad distribution of the alternative oxidase in microsporidian parasites
.
PLoS Pathog.
6
,
e1000761
17
Stenmark
,
P.
and
Nordlund
,
P.
(
2003
)
A prokaryotic alternative oxidase present in the bacterium Novosphingobium aromaticivorans
.
FEBS Lett.
552
,
189
192
18
Dunn
,
A.K.
,
Karr
,
E.A.
,
Wang
,
Y.
,
Batton
,
A.R.
,
Ruby
,
E.G.
and
Stabb
,
E.V.
(
2010
)
The alternative oxidase (AOX) gene in vibrio fischeri is controlled by NsrR and upregulated in response to nitric oxide
.
Mol. Microbiol.
77
,
44
55
19
Dunn
,
A.K.
(
2018
)
Alternative oxidase activity reduces stress in vibrio fischeri cells exposed to nitric oxide
.
J. Bacteriol.
200
,
e00797-17
20
Kaneko
,
I.
and
Ishii
,
H.
(
2009
)
Effect of azoxystrobin on activities of antioxidant enzymes and alternative oxidase in wheat head blight pathogens Fusarium graminearum and Microdochium nivale
.
J. Gen. Plant Pathol.
75
,
388
21
Avila-Adame
,
C.
and
Köller
,
W.
(
2002
)
Disruption of the alternative oxidase gene in magnaporthe grisea and Its impact on host infection
.
Mol. Plant Microbe Interact.
15
,
493
500
22
Affourtit
,
C.
,
Heaney
,
S.P.
and
Moore
,
A.L.
(
2000
)
Mitochondrial electron transfer in the wheat pathogenic fungus Septoria tritici: on the role of alternative respiratory enzymes in fungicide resistance
.
Biochim. Biophys. Acta Bioenergetics
1459
,
291
298
23
Sierra-Campos
,
E.
,
Velázquez
,
I.
,
Matuz-Mares
,
D.
,
Villavicencio-Queijeiro
,
A.
and
Pardo
,
J.P.
(
2009
)
Functional properties of the Ustilago maydis alternative oxidase under oxidative stress conditions
.
Mitochondrion
9
,
96
102
24
Xu
,
T.
,
Yao
,
F.
,
Liang
,
W.-S.
,
Li
,
Y.-H.
,
Li
,
D.-R.
,
Wang
,
H.
et al (
2012
)
Involvement of alternative oxidase in the regulation of growth, development, and resistance to oxidative stress of Sclerotinia sclerotiorum
.
J. Microbiol.
50
,
594
602
25
Xu
,
T.
,
Wang
,
Y.-T.
,
Liang
,
W.-S.
,
Yao
,
F.
,
Li
,
Y.-H.
,
Li
,
D.-R.
et al (
2013
)
Involvement of alternative oxidase in the regulation of sensitivity of Sclerotinia sclerotiorum to the fungicides azoxystrobin and procymidone
.
J. Microbiol.
51
,
352
358
26
Tamura
,
H.
,
Mizutani
,
A.
,
Yukioka
,
H.
,
Miki
,
N.
,
Ohba
,
K.
and
Masuko
,
M.
(
1999
)
Effect of the methoxyiminoacetamide fungicide, SSF129, on respiratory activity in Botrytis cinerea
.
Pest. Sci
55
,
681
686
27
Hildebrandt
,
T.M.
and
Grieshaber
,
M.K.
(
2008
)
Redox regulation of mitochondrial sulfide oxidation in the lugworm, arenicola marina
.
J. Exp. Biol.
211
,
2617
2623
28
Buchner
,
T.
,
Abele
,
D.
and
Pörtner
,
H.-O.
(
2001
)
Oxyconformity in the intertidal worm Sipunculus nudus: the mitochondrial background and energetic consequences
.
Comp. Biochem. Physiol. B Biochem. Mol. Biol.
129
,
109
120
29
Saari
,
S.
,
Garcia
,
G.S.
,
Bremer
,
K.
,
Chioda
,
M.M.
,
Andjelković
,
A.
,
Debes
,
P.V.
, et al (
2019
)
Alternative respiratory chain enzymes: therapeutic potential and possible pitfalls
.
Biochim. Biophys. Acta Mol. Basis Dis.
1865
,
854
866
30
Bremer
,
K.
,
Yasuo
,
H.
,
Debes
,
P.V.
and
Jacobs
,
H.T.
(
2021
)
The alternative oxidase (AOX) increases sulphide tolerance in the highly invasive marine invertebrate Ciona intestinalis
.
J. Exp. Biol.
224
,
jeb242985
31
Joseph-Horne
,
T.
,
Hollomon
,
D.W.
and
Wood
,
P.M.
(
2001
)
Fungal respiration: a fusion of standard and alternative components
.
Biochim. Biophys. Acta Bioenergetics
1504
,
179
195
32
Rustin
,
P.
,
Jacobs
,
H.T.
,
Dietrich
,
A.
,
Lightowlers
,
R.N.
,
Tarassov
,
I.
and
Corral-Debrinski
,
M.
(
2007
)
Targeting allotopic material to the mitochondrial compartment: new tools for better understanding mitochondrial physiology and prospect for therapy
.
Méd. Sci.
23
,
519
525
33
Dassa
,
E.P.
,
Dufour
,
E.
,
Goncalves
,
S.
,
Jacobs
,
H.T.
and
Rustin
,
P.
(
2009
)
The alternative oxidase, a tool for compensating cytochrome c oxidase deficiency in human cells
.
Physiol. Plantarum
137
,
427
434
34
Rustin
,
P.
and
Jacobs
,
H.T.
(
2009
)
Respiratory chain alternative enzymes as tools to better understand and counteract respiratory chain deficiencies in human cells and animals
.
Physiol. Plantarum
137
,
362
370
35
Schiff
,
M.
,
Bénit
,
P.
,
Jacobs
,
H.T.
,
Vockley
,
J.
and
Rustin
,
P.
(
2012
)
Therapies in inborn errors of oxidative metabolism
.
Trends Endocrinol. Metab.
23
,
488
495
36
El-Khoury
,
R.
,
Kemppainen
,
K.K.
,
Dufour
,
E.
,
Szibor
,
M.
,
Jacobs
,
H.T.
and
Rustin
,
P.
(
2014
)
Engineering the alternative oxidase gene to better understand and counteract mitochondrial defects: state of the art and perspectives
.
Br. J. Pharmacol.
171
,
2243
2249
37
Clarkson
,
A.B.
,
Bienen
,
E.J.
,
Pollakis
,
G.
and
Grady
,
R.W.
(
1989
)
Respiration of bloodstream forms of the parasite Trypanosoma brucei brucei is dependent on a plant-like alternative oxidase
.
J. Biol. Chem.
264
,
17770
6
38
Bienen
,
E.J.
,
Maturi
,
R.K.
,
Pollakis
,
G.
and
Clarkson
,
A.B.
(
1993
)
Non-cytochrome mediated mitochondrial ATP production in bloodstream form Trypanosoma brucei brucei
.
Eur. J. Biochem.
216
,
75
80
39
van Hellemond
,
J.J.
,
Bakker
,
B.M.
and
Tielens
,
A.G.M.
(
2005
)
Energy metabolism and Its compartmentation in Trypanosoma brucei
.
Adv. Microb. Physiol.
50
,
199
226
40
Luévano-Martínez
,
L.A.
,
Girard
,
R.M.B.M.
,
Alencar
,
M.B.
and
Silber
,
A.M.
(
2020
)
ATP regulates the activity of an alternative oxidase in Trypanosoma brucei
.
FEBS Lett.
594
,
2150
2158
41
Jakob
,
M.
,
Hoffmann
,
A.
,
Amodeo
,
S.
,
Peitsch
,
C.
,
Zuber
,
B.
and
Ochsenreiter
,
T.
(
2016
)
Mitochondrial growth during the cell cycle of Trypanosoma brucei bloodstream forms
.
Sci. Rep.
6
,
36565
42
Durieux
,
P.O.
,
Schütz
,
P.
,
Brun
,
R.
and
Köhler
,
P.
(
1991
)
Alterations in Krebs cycle enzyme activities and carbohydrate catabolism in two strains of Trypanosoma brucei during in vitro differentiation of their bloodstream to procyclic stages
.
Mol. Biochem. Parasitol.
45
,
19
27
43
Roberts
,
C.W.
,
Roberts
,
F.
,
Henriquez
,
F.L.
,
Akiyoshi
,
D.
,
Samuel
,
B.U.
,
Richards
,
T.A.
et al (
2004
)
Evidence for mitochondrial-derived alternative oxidase in the apicomplexan parasite Cyptosporidium parvum: a potential anti-microbial agent target
.
Int. J. Parasitol.
34
,
297
308
44
Bouzid
,
M.
,
Hunter
,
P.R.
,
Chalmers
,
R.M.
and
Tyler
,
K.M.
(
2013
)
Cryptosporidium pathogenicity and virulence
.
Clin. Microbiol. Rev.
26
,
115
134
45
Tandel
,
J.
,
English
,
E.D.
,
Sateriale
,
A.
,
Gullicksrud
,
J.A.
,
Beiting
,
D.P.
,
Sullivan
,
M.C.
et al (
2019
)
Life cycle progression and sexual development of the apicomplexan parasite Cryptosporidium parvum
.
Nat. Microbiol.
4
,
2226
2236
46
Zhang
,
H.
,
Guo
,
F.
,
Zhou
,
H.
and
Zhu
,
G.
(
2012
)
Transcriptome analysis reveals unique metabolic features in the cryptosporidium parvum oocysts associated with environmental survival and stresses
.
BMC Genomics
13
,
647
647
47
Barta
,
J.R.
and
Thompson
,
R.C.A.
(
2006
)
What is cryptosporidium? Reappraising its biology and phylogenetic affinities
.
Trends Parasitol.
22
,
463
468
48
Abrahamsen
,
M.S.
,
Templeton
,
T.J.
,
Enomoto
,
S.
,
Abrahante
,
J.E.
,
Zhu
,
G.
,
Lancto
,
C.A.
et al (
2004
)
Complete genome sequence of the apicomplexan, cryptosporidium parvum
.
Science
304
,
441
445
49
Denoeud
,
F.
,
Roussel
,
M.
,
Noel
,
B.
,
Wawrzyniak
,
I.
,
Silva
,
C.D.
,
Diogon
,
M.
et al (
2011
)
Genome sequence of the stramenopile Blastocystis, a human anaerobic parasite
.
Genome Biol.
12
,
R29
50
Henriquez
,
F.L.
,
McBride
,
J.
,
Campbell
,
S.J.
,
Ramos
,
T.
,
Ingram
,
P.R.
,
Roberts
,
F.
et al (
2009
)
Acanthamoeba alternative oxidase genes: identification, characterisation and potential as antimicrobial targets
.
Int. J. Parasitol.
39
,
1417
1424
51
Khan
,
N.A.
(
2006
)
Acanthamoeba: biology and increasing importance in human health
.
FEMS Microbiol. Rev.
30
,
564
595
52
Kot
,
K.
,
Łanocha-Arendarczyk
,
N.A.
and
Kosik-Bogacka
,
D.I.
(
2018
)
Amoebas from the genus acanthamoeba and their pathogenic properties
.
Ann. Parasitol.
64
,
299
308
53
Jarmuszkiewicz
,
W.
,
Sluse-Goffart
,
C.M.
,
Hryniewiecka
,
L.
,
Michejda
,
J.
and
Sluse
,
F.E.
(
1998
)
Electron partitioning between the two branching quinol-oxidizing pathways in Acanthamoeba castellanii mitochondria during steady-state state 3 respiration*
.
J. Biol. Chem.
273
,
10174
10180
54
Jarmuszkiewicz
,
W.
,
Sluse
,
F.E.
,
Hryniewiecka
,
L.
and
Sluse-Goffart
,
C.M.
(
2002
)
Interactions between the cytochrome pathway and the alternative oxidase in isolated Acanthamoeba castellanii mitochondria
.
J. Bioenerg. Biomembr.
34
,
31
40
55
Navet
,
R.
,
Jarmuszkiewicz
,
W.
,
Douette
,
P.
,
Sluse-Goffart
,
C.M.
and
Sluse
,
F.E.
(
2004
)
Mitochondrial respiratory chain complex patterns from acanthamoeba castellanii and lycopersicon esculentum: comparative analysis by BN-PAGE and evidence of protein–protein interaction between alternative oxidase and complex III
.
J. Bioenerg. Biomembr.
36
,
471
479
56
Woyda-Ploszczyca
,
A.M.
,
Sluse
,
F.E.
and
Jarmuszkiewicz
,
W.
(
2009
)
Regulation of acanthamoeba castellanii alternative oxidase activity by mutual exclusion of purine nucleotides; ATP's inhibitory effect
.
Biochim. Biophys. Acta Bioenergetics
1787
,
264
271
57
Antos-Krzeminska
,
N.
and
Jarmuszkiewicz
,
W.
(
2014
)
Functional expression of the Acanthamoeba castellanii alternative oxidase in Escherichia coli; regulation of the activity and evidence for Acaox gene function
.
Biochem. Cell Biol.
92
,
235
241
58
Czarna
,
M.
and
Jarmuszkiewicz
,
W.
(
2005
)
Activation of alternative oxidase and uncoupling protein lowers hydrogen peroxide formation in amoeba Acanthamoeba castellanii mitochondria
.
FEBS Lett.
579
,
3136
3140
59
Jarmuszkiewicz
,
W.
,
Fraczyk
,
O.
and
Hryniewiecka
,
L.
(
2001
)
Effect of growth at low temperature on the alternative pathway respiration in Acanthamoeba castellanii mitochondria
.
Acta Biochim. Pol.
48
,
729
737
60
Williams
,
B.A.P.
,
Hirt
,
R.P.
,
Lucocq
,
J.M.
and
Embley
,
T.M.
(
2002
)
A mitochondrial remnant in the microsporidian trachipleistophora hominis
.
Nature
418
,
865
869
61
Katinka
,
M.D.
,
Duprat
,
S.
,
Cornillot
,
E.
,
Méténier
,
G.
,
Thomarat
,
F.
,
Prensier
,
G.
et al (
2001
)
Genome sequence and gene compaction of the eukaryote parasite Encephalitozoon cuniculi
.
Nature
414
,
450
453
62
Keeling
,
P.J.
and
Fast
,
N.M.
(
2002
)
MICROSPORIDIA: biology and evolution of highly reduced intracellular parasites
.
Microbiology
56
,
93
116
63
Beznoussenko
,
G.V.
,
Dolgikh
,
V.V.
,
Seliverstova
,
E.V.
,
Semenov
,
P.B.
,
Tokarev
,
Y.S.
,
Trucco
,
A.
et al (
2007
)
Analogs of the Golgi complex in microsporidia: structure and avesicular mechanisms of function
.
J. Cell Sci.
120
,
1288
1298
64
Keeling
,
P.J.
and
Corradi
,
N.
(
2011
)
Shrink it or lose it: balancing loss of function with shrinking genomes in the microsporidia
.
Virulence
2
,
67
70
65
Peyretaillade
,
E.
,
Alaoui
,
H.E.
,
Diogon
,
M.
,
Polonais
,
V.
,
Parisot
,
N.
,
Biron
,
D.G.
et al (
2011
)
Extreme reduction and compaction of microsporidian genomes
.
Res. Microbiol.
162
,
598
606
66
Keeling
,
P.J.
,
Corradi
,
N.
,
Morrison
,
H.G.
,
Haag
,
K.L.
,
Ebert
,
D.
,
Weiss
,
L.M.
et al (
2010
)
The reduced genome of the parasitic microsporidian enterocytozoon bieneusi lacks genes for core carbon metabolism
.
Genome Biol. Evol.
2
,
304
309
67
Dolgikh
,
V.V.
,
Senderskiy
,
I.V.
,
Pavlova
,
O.A.
,
Naumov
,
A.M.
and
Beznoussenko
,
G.V.
(
2011
)
Immunolocalization of an alternative respiratory chain in antonospora (Paranosema) locustae spores: mitosomes retain their role in microsporidial energy metabolism
.
Eukaryot. Cell
10
,
588
593
68
Helmerhorst
,
E.J.
,
Murphy
,
M.P.
,
Troxler
,
R.F.
and
Oppenheim
,
F.G.
(
2002
)
Characterization of the mitochondrial respiratory pathways in Candida albicans
.
Biochim. Biophys. Acta Bioenergetics
1556
,
73
80
69
Copsey
,
A.C.
,
Barsottini
,
M.R.O.
,
May
,
B.
,
Xu
,
F.
,
Albury
,
M.S.
,
Young
,
L.
et al (
2021
)
Kinetic characterisation and inhibitor sensitivity of Candida albicans and Candida auris recombinant AOX expressed in a self-assembled proteoliposome system
.
Sci. Rep.
11
,
14748
70
Akhter
,
S.
,
McDade
,
H.C.
,
Gorlach
,
J.M.
,
Heinrich
,
G.
,
Cox
,
G.M.
and
Perfect
,
J.R.
(
2003
)
Role of alternative oxidase gene in pathogenesis of Cryptococcus neoformans
.
Infect. Immun.
71
,
5794
5802
71
Magnani
,
T.
,
Soriani
,
F.M.
,
Martins
,
V.P.
,
Nascimento
,
A.M.
,
Tudella
,
V.G.
,
Curti
,
C.
et al (
2007
)
Cloning and functional expression of the mitochondrial alternative oxidase of Aspergillus fumigatus and its induction by oxidative stress
.
FEMS Microbiol. Lett.
271
,
230
238
72
Broxton
,
C.N.
and
Culotta
,
V.C.
(
2016
)
An adaptation to low copper in Candida albicans involving SOD enzymes and the alternative oxidase
.
PLoS ONE
11
,
e0168400
73
Yan
,
L.
,
Li
,
M.
,
Cao
,
Y.
,
Gao
,
P.
,
Cao
,
Y.
,
Wang
,
Y.
et al (
2009
)
The alternative oxidase of Candida albicans causes reduced fluconazole susceptibility
.
J. Antimicrob. Chemother.
64
,
764
773
74
Duvenage
,
L.
,
Walker
,
L.A.
,
Bojarczuk
,
A.
,
Johnston
,
S.A.
,
MacCallum
,
D.M.
,
Munro
,
C.A.
et al (
2019
)
Inhibition of classical and alternative modes of respiration in Candida albicans leads to cell wall remodeling and increased macrophage recognition
.
mBio
10
,
e02535-18
75
Sears
,
D.
and
Schwartz
,
B.S.
(
2017
)
Candida auris: an emerging multidrug-resistant pathogen
.
Int. J. Infect. Dis.
63
,
95
98
76
Forsberg
,
K.
,
Woodworth
,
K.
,
Walters
,
M.
,
Berkow
,
E.L.
,
Jackson
,
B.
,
Chiller
,
T.
et al (
2019
)
Candida auris: the recent emergence of a multidrug-resistant fungal pathogen
.
Med. Mycol.
57
,
1
12
77
Nett
,
J.E.
(
2019
)
Candida auris: an emerging pathogen ‘incognito’?
PLoS Pathog.
15
,
e1007638
78
Satoh
,
K.
,
Makimura
,
K.
,
Hasumi
,
Y.
,
Nishiyama
,
Y.
,
Uchida
,
K.
and
Yamaguchi
,
H.
(
2009
)
Candida auris sp. nov., a novel ascomycetous yeast isolated from the external ear canal of an inpatient in a Japanese hospital
.
Microbiol. Immunol.
53
,
41
44
79
Chaabane
,
F.
,
Graf
,
A.
,
Jequier
,
L.
and
Coste
,
A.T.
(
2019
)
Review on antifungal resistance mechanisms in the emerging pathogen candida auris
.
Front. Microbiol.
10
,
2788
80
Mba
,
I.E.
and
Nweze
,
E.I.
(
2020
)
Mechanism of Candida pathogenesis: revisiting the vital drivers
.
Eur. J. Clin. Microbiol.
39
,
1797
1819
81
Martins
,
V.P.
,
Dinamarco
,
T.M.
,
Soriani
,
F.M.
,
Tudella
,
V.G.
,
Oliveira
,
S.C.
,
Goldman
,
G.H.
et al (
2011
)
Involvement of an alternative oxidase in oxidative stress and mycelium-to-yeast differentiation in Paracoccidioides brasiliensis
.
Eukaryot. Cell
10
,
237
248
82
Magnani
,
T.
,
Soriani
,
F.M.
,
Martins
,
V.P.
,
Policarpo
,
A.C.
,
Sorgi
,
C.A.
,
Faccioli
,
L.H.
et al (
2009
)
Silencing of mitochondrial alternative oxidase gene of Aspergillus fumigatus enhances reactive oxygen species production and killing of the fungus by macrophages
.
J. Bioenerg. Biomembr.
40
,
631
83
Clarkson
,
A.B.
and
Brohn
,
F.H.
(
1976
)
Trypanosomiasis: an approach to chemotherapy by the inhibition of carbohydrate catabolism
.
Science
194
,
204
206
84
Ebiloma
,
G.U.
,
Balogun
,
E.O.
,
Cueto-Díaz
,
E.J.
,
Koning
,
H.P.
and
Dardonville
,
C.
(
2019
)
Alternative oxidase inhibitors: mitochondrion-targeting as a strategy for new drugs against pathogenic parasites and fungi
.
Med. Res. Rev.
39
,
1553
1602
85
Davies
,
C.R.
,
Wohlgemuth
,
F.
,
Young
,
T.
,
Violet
,
J.
,
Dickinson
,
M.
,
Sanders
,
J.-W.
et al (
2021
)
Evolving challenges and strategies for fungal control in the food supply chain
.
Fungal Biol. Rev.
36
,
15
26
86
Drakopoulos
,
D.
,
Kägi
,
A.
,
Gimeno
,
A.
,
Six
,
J.
,
Jenny
,
E.
,
Forrer
,
H.-R.
et al (
2020
)
Prevention of Fusarium head blight infection and mycotoxins in wheat with cut-and-carry biofumigation and botanicals
.
Field Crop Res.
246
,
107681
87
Cruz-Mireles
,
N.
,
Eisermann
,
I.
,
Garduño-Rosales
,
M.
,
Molinari
,
C.
,
Ryder
,
L.S.
,
Tang
,
B.
et al (
2021
)
Magnaporthe oryzae, methods and protocols
.
Methods Mol. Biol.
2356
,
19
40
88
Vollmeister
,
E.
,
Schipper
,
K.
,
Baumann
,
S.
,
Haag
,
C.
,
Pohlmann
,
T.
,
Stock
,
J.
et al (
2012
)
Fungal development of the plant pathogen Ustilago maydis
.
FEMS Microbiol. Rev.
36
,
59
77
89
Moreira
,
X.
,
Granjel
,
R.R.
,
Fuente
,
M.
,
Fernández-Conradi
,
P.
,
Pasch
,
V.
,
Soengas
,
P.
et al (
2021
)
Apparent inhibition of induced plant volatiles by a fungal pathogen prevents airborne communication between potato plants
.
Plant Cell Environ.
44
,
1192
1201
90
Ranjan
,
A.
,
Jayaraman
,
D.
,
Grau
,
C.
,
Hill
,
J.H.
,
Whitham
,
S.A.
,
Ané
,
J.
et al (
2018
)
The pathogenic development of Sclerotinia sclerotiorum in soybean requires specific host NADPH oxidases
.
Mol. Plant Pathol.
19
,
700
714
91
Meinhardt
,
L.W.
,
Rincones
,
J.
,
Bailey
,
B.A.
,
Aime
,
M.C.
,
Griffith
,
G.W.
,
Zhang
,
D.
et al (
2008
)
Moniliophthora perniciosa, the causal agent of witches’ broom disease of cacao: what's new from this old foe?
Mol. Plant Pathol.
9
,
577
588
92
Bailey
,
B.A.
,
Evans
,
H.C.
,
Phillips-Mora
,
W.
,
Ali
,
S.S.
and
Meinhardt
,
L.W.
(
2018
)
Moniliophthora roreri, causal agent of cacao frosty pod rot
.
Mol. Plant Pathol.
19
,
1580
1594
93
Churchill
,
A.C.L.
(
2011
)
Mycosphaerella fijiensis, the black leaf streak pathogen of banana: progress towards understanding pathogen biology and detection, disease development, and the challenges of control
.
Mol. Plant Pathol.
12
,
307
328
94
González-Domínguez
,
E.
,
Armengol
,
J.
and
Rossi
,
V.
(
2017
)
Biology and epidemiology of venturia species affecting fruit crops: a review
.
Front. Plant Sci.
8
,
1496
95
Tournas
,
V.H.
(
2008
)
Spoilage of vegetable crops by bacteria and fungi and related health hazards
.
Crit. Rev. Microbiol.
31
,
33
44
96
Miguez
,
M.
,
Reeve
,
C.
,
Wood
,
P.M.
and
Hollomon
,
D.W.
(
2004
)
Alternative oxidase reduces the sensitivity of Mycosphaerella graminicola to QOI fungicides
.
Pest. Manag. Sci.
60
,
3
7
97
Thomazella
,
D.P.T.
,
Teixeira
,
P.J.P.L.
,
Oliveira
,
H.C.
,
Saviani
,
E.E.
,
Rincones
,
J.
,
Toni
,
I.M.
et al (
2012
)
The hemibiotrophic cacao pathogen Moniliophthora perniciosa depends on a mitochondrial alternative oxidase for biotrophic development
.
New Phytol.
194
,
1025
1034
98
Barsottini
,
M.R.
,
Pires
,
B.A.
,
Vieira
,
M.L.
,
Pereira
,
J.G.
,
Costa
,
P.C.
,
Sanitá
,
J.
et al (
2019
)
Synthesis and testing of novel alternative oxidase (AOX) inhibitors with antifungal activity against Moniliophthora perniciosa (Stahel), the causal agent of witches’ broom disease of cocoa, and other phytopathogens
.
Pest Manag. Sci.
75
,
1295
1303
99
Musso
,
L.
,
Fabbrini
,
A.
and
Dallavalle
,
S.
(
2020
)
Natural compound-derived cytochrome bc1 complex inhibitors as antifungal agents
.
Molecules
25
,
4582
100
Joseph-Horne
,
T.
,
Wood
,
P.M.
,
Wood
,
C.K.
,
Moore
,
A.L.
,
Headrick
,
J.
and
Hollomon
,
D.
(
1998
)
Characterization of a split respiratory pathway in the wheat ‘take-all’ fungus, gaeumannomyces graminis var.tritici *
.
J. Biol. Chem.
273
,
11127
11133
101
Avila-Adame
,
C.
and
Köller
,
W.
(
2003
)
Impact of alternative respiration and target-site mutations on responses of germinating conidia of Magnaporthe grisea to Qo-inhibiting fungicides
.
Pest Manag. Sci.
59
,
303
309
102
Steinfeld
,
U.
,
Sierotzki
,
H.
,
Parisi
,
S.
,
Poirey
,
S.
and
Gisi
,
U.
(
2001
)
Sensitivity of mitochondrial respiration to different inhibitors in Venturia inaequalis
.
Pest Manag. Sci.
57
,
787
796
103
Färber
,
R.B.K.
,
Chin
,
K.M.
and
Leadbitter
,
N.
(
2002
)
Sensitivity of Venturia inaequalis to trifloxystrobin
.
Pest Manag. Sci.
58
,
261
267
104
Walker
,
A.
,
Auclair
,
C.
,
Gredt
,
M.
and
Leroux
,
P.
(
2009
)
First occurrence of resistance to strobilurin fungicides in Microdochium nivale and Microdochium majus from French naturally infected wheat grains
.
Pest. Manag. Sci.
65
,
906
915
105
Fernández-Ortuño
,
D.
,
Torés
,
J.A.
,
Vicente
,
A.D.
and
Pérez-García
,
A.
(
2010
)
The QoI fungicides, the rise and fall of a successful class of agricultural fungicides
. In
Fungicides
(Carisse, O., ed.), pp.
203
220
. ISBN: 978-953-307-266-1, InTech, Available from: http://www.intechopen.com/books/fungicides/the-qoifungicides-the-rise-and-fall-of-a-successful-class-of-agricultural-fungicides https://cdn.intechopen.com/pdfs/12388/InTech-The_qoi_fungicides_the_rise_and_fall_of_a_successful_class_of_agricultural_fungicides.pdf
106
Avila-Adame
,
C.
and
Köller
,
W.
(
2003
)
Characterization of spontaneous mutants of Magnaporthe grisea expressing stable resistance to the Qo-inhibiting fungicide azoxystrobin
.
Curr. Genet.
42
,
332
338
107
Casida
,
J.E.
and
Durkin
,
K.A.
(
2017
)
Pesticide chemical research in toxicology: lessons from nature
.
Chem. Res. Toxicol.
30
,
94
104
108
Strippoli
,
V.
,
D'Auria
,
F.D.
,
Tecca
,
M.
,
Callari
,
A.
and
Simonetti
,
G.
(
2000
)
Propyl gallate increases in vitro antifungal imidazole activity against Candida albicans
.
Int. J. Antimicrob. Agents
16
,
73
76
109
Karaffa
,
L.
,
Sándor
,
E.
,
Kozma
,
J.
,
Kubicek
,
C.P.
and
Szentirmai
,
A.
(
1999
)
The role of the alternative respiratory pathway in the stimulation of cephalosporin C formation by soybean oil in Acremonium chrysogenum
.
Appl. Microbiol. Biotechnol.
51
,
633
638
110
Leiter
,
É.
,
Park
,
H.-S.
,
Kwon
,
N.-J.
,
Han
,
K.-H.
,
Emri
,
T.
,
Oláh
,
V.
et al (
2016
)
Characterization of the aodA, dnmA, mnSOD and pimA genes in Aspergillus nidulans
.
Sci. Rep.
6
,
20523
111
Molnár
,
Á.P
,
Németh
,
Z.
,
Fekete
,
E.
,
Flipphi
,
M.
,
Keller
,
N.P.
and
Karaffa
,
L.
(
2018
)
Analysis of the relationship between alternative respiration and sterigmatocystin formation in aspergillus nidulans
.
Toxins
10
,
168
112
Keller
,
N.P.
,
Kantz
,
N.J.
and
Adams
,
T.H.
(
1994
)
Aspergillus nidulans verA is required for production of the mycotoxin sterigmatocystin
.
Appl. Environ. Microb.
60
,
1444
1450
113
Kotinagu
,
K.
,
Mohanamba
,
T.
and
Kumari
,
L.R.
(
2015
)
Assessment of aflatoxin B1 in livestock feed and feed ingredients by high-performance thin layer chromatography
.
Vet. World
8
,
1396
1399
114
Rushing
,
B.R.
and
Selim
,
M.I.
(
2019
)
Aflatoxin B1: a review on metabolism, toxicity, occurrence in food, occupational exposure, and detoxification methods
.
Food Chem. Toxicol.
124
,
81
100
115
Carballo
,
D.
,
Tolosa
,
J.
,
Ferrer
,
E.
and
Berrada
,
H.
(
2019
)
Dietary exposure assessment to mycotoxins through total diet studies. A review
.
Food Chem. Toxicol.
128
,
8
20
116
Moore
,
A.L.
,
Umbach
,
A.L.
and
Siedow
,
J.N.
(
1995
)
Structure-function relationships of the alternative oxidase of plant mitochondria: a model of the active site
.
J. Bioenerg. Biomembr.
27
,
367
377
117
Siedow
,
J.N.
,
Umbach
,
A.L.
and
Moore
,
A.L.
(
1995
)
The active site of the cyanide-resistant oxidase from plant mitochondria contains a binuclear iron center
.
FEBS Lett.
362
,
10
14
118
Andersson
,
M.E.
and
Nordlund
,
P.
(
1999
)
A revised model of the active site of alternative oxidase
.
FEBS Lett.
449
,
17
22
119
Berthold
,
D.A.
,
Andersson
,
M.E.
and
Nordlund
,
P.
(
2000
)
New insight into the structure and function of the alternative oxidase
.
Biochim. Biophys. Acta Bioenergetics
1460
,
241
254
120
Nihei
,
C.
,
Fukai
,
Y.
,
Kawai
,
K.
,
Osanai
,
A.
,
Yabu
,
Y.
,
Suzuki
,
T.
et al (
2003
)
Purification of active recombinant trypanosome alternative oxidase
.
FEBS Lett.
538
,
35
40
121
Kido
,
Y.
,
Sakamoto
,
K.
,
Nakamura
,
K.
,
Harada
,
M.
,
Suzuki
,
T.
,
Yabu
,
Y.
et al (
2010
)
Purification and kinetic characterization of recombinant alternative oxidase from Trypanosoma brucei brucei
.
Biochim. Biophys. Acta Bioenergetics
1797
,
443
450
122
Shiba
,
T.
,
Kido
,
Y.
,
Sakamoto
,
K.
,
Inaoka
,
D.K.
,
Tsuge
,
C.
,
Tatsumi
,
R.
et al (
2013
)
Structure of the trypanosome cyanide-insensitive alternative oxidase
.
Proc. Natl Acad. Sci. U.S.A.
110
,
4580
4585
123
Ebiloma
,
G.U.
,
Ayuga
,
T.D.
,
Balogun
,
E.O.
,
Gil
,
L.A.
,
Donachie
,
A.
,
Kaiser
,
M.
et al (
2018
)
Inhibition of trypanosome alternative oxidase without its N-terminal mitochondrial targeting signal (ΔMTS-TAO) by cationic and non-cationic 4-hydroxybenzoate and 4-alkoxybenzaldehyde derivatives active against T. brucei and T. congolense
.
Eur. J. Med. Chem.
150
,
385
402
124
Elliott
,
C.
,
Young
,
L.
,
May
,
B.
,
Shearman
,
J.
,
Albury
,
M.S.
,
Kido
,
Y.
et al (
2014
)
Purification and characterisation of recombinant DNA encoding the alternative oxidase from Sauromatum guttatum
.
Mitochondrion
19
,
261
268
125
Barsottini
,
M.R.O.
,
Copsey
,
A.
,
Young
,
L.
,
Baroni
,
R.M.
,
Cordeiro
,
A.T.
,
Pereira
,
G.A.G.
et al (
2020
)
Biochemical characterization and inhibition of the alternative oxidase enzyme from the fungal phytopathogen Moniliophthora perniciosa
.
Commun. Biol.
3
,
263
126
Shiba
,
T.
,
Inaoka
,
D.K.
,
Takahashi
,
G.
,
Tsuge
,
C.
,
Kido
,
Y.
,
Young
,
L.
et al (
2019
)
Insights into the ubiquinol/dioxygen binding and proton relay pathways of the alternative oxidase
.
Biochim. Biophys. Acta Bioenergetics
1860
,
375
382
127
Minagawa
,
N.
,
Yabu
,
Y.
,
Kita
,
K.
,
Nagai
,
K.
,
Ohta
,
N.
,
Meguro
,
K.
et al (
1996
)
An antibiotic, ascofuranone, specifically inhibits respiration and in vitro growth of long slender bloodstream forms of trypanosoma brucei brucei
.
Mol. Biochem. Parasitol.
81
,
127
136
128
Fukai
,
Y.
,
Amino
,
H.
,
Hirawake
,
H.
,
Yabu
,
Y.
,
Ohta
,
N.
,
Minagawa
,
N.
et al (
1999
)
Functional expression of the ascofuranone-sensitive Trypanosoma brucei brucei alternative oxidase in the cytoplasmic membrane of escherichia coli
.
Comp. Biochem. Physiol. C Pharmacol. Toxicol. Endocrinol.
124
,
141
148
129
Yabu
,
Y.
,
Suzuki
,
T.
,
Nihei
,
C.
,
Minagawa
,
N.
,
Hosokawa
,
T.
,
Nagai
,
K.
et al (
2006
)
Chemotherapeutic efficacy of ascofuranone in Trypanosoma vivax-infected mice without glycerol
.
Parasitol. Int.
55
,
39
43
130
Saimoto
,
H.
,
Kido
,
Y.
,
Haga
,
Y.
,
Sakamoto
,
K.
and
Kita
,
K.
(
2012
)
Pharmacophore identification of ascofuranone, potent inhibitor of cyanide-insensitive alternative oxidase of trypanosoma brucei
.
J. Biochem.
153
,
267
273
131
West
,
R.A.
,
O'Doherty
,
O.G.
,
Askwith
,
T.
,
Atack
,
J.
,
Beswick
,
P.
,
Laverick
,
J.
et al (
2017
)
African trypanosomiasis: synthesis & SAR enabling novel drug discovery of ubiquinol mimics for trypanosome alternative oxidase
.
Eur. J. Med. Chem.
141
,
676
689
132
Rosell-Hidalgo
,
A.
,
Young
,
L.
,
Moore
,
A.L.
and
Ghafourian
,
T.
(
2021
)
QSAR and molecular docking for the search of AOX inhibitors: a rational drug discovery approach
.
J. Comput. Aid. Mol. Des.
35
,
245
260
133
May
,
B.
,
Young
,
L.
and
Moore
,
A.L.
(
2017
)
Structural insights into the alternative oxidases: are all oxidases made equal?
Biochem. Soc. Trans.
45
,
731
740
134
Perales-Clemente
,
E.
,
Bayona-Bafaluy
,
M.P.
,
Pérez-Martos
,
A.
,
Barrientos
,
A.
,
Fernández-Silva
,
P.
and
Enriquez
,
J.A.
(
2008
)
Restoration of electron transport without proton pumping in mammalian mitochondria
.
Proc. Natl Acad. Sci. U.S.A.
105
,
18735
18739
135
Dogan
,
S.A.
,
Cerutti
,
R.
,
Benincá
,
C.
,
Brea-Calvo
,
G.
,
Jacobs
,
H.T.
,
Zeviani
,
M.
et al (
2018
)
Perturbed redox signaling exacerbates a mitochondrial myopathy
.
Cell Metab.
28
,
764
775.e5
136
Szibor
,
M.
,
Schreckenberg
,
R.
,
Gizatullina
,
Z.
,
Dufour
,
E.
,
Wiesnet
,
M.
,
Dhandapani
,
P.K.
et al (
2020
)
Respiratory chain signalling is essential for adaptive remodelling following cardiac ischaemia
.
J. Cell. Mol. Med.
24
,
3534
3548
137
El-Khoury
,
R.
,
Kaulio
,
E.
,
Lassila
,
K.A.
,
Crowther
,
D.C.
,
Jacobs
,
H.T.
and
Rustin
,
P.
(
2016
)
Expression of the alternative oxidase mitigates beta-amyloid production and toxicity in model systems
.
Free Radic. Biol. Med.
96
,
57
66
138
Mills
,
E.L.
,
Kelly
,
B.
,
Logan
,
A.
,
Costa
,
A.S.H.
,
Varma
,
M.
,
Bryant
,
C.E.
et al (
2016
)
Succinate dehydrogenase supports metabolic repurposing of mitochondria to drive inflammatory macrophages
.
Cell
167
,
457
470.e13
139
Giordano
,
L.
,
Farnham
,
A.
,
Dhandapani
,
P.K.
,
Salminen
,
L.
,
Bhaskaran
,
J.
,
Voswinckel
,
R.
et al (
2019
)
Alternative oxidase attenuates cigarette smoke-induced lung dysfunction and tissue damage
.
Am. J. Respir. Cell Mol. Biol.
60
,
515
522
140
Szibor
,
M.
,
Gainutdinov
,
T.
,
Fernandez-Vizarra
,
E.
,
Dufour
,
E.
,
Gizatullina
,
Z.
,
Debska-Vielhaber
,
G.
et al (
2020
)
Bioenergetic consequences from xenotopic expression of a tunicate AOX in mouse mitochondria: switch from RET and ROS to FET
.
Biochim. Biophys. Acta Bioenergetics
1861
,
148137
141
Fernandez-Ayala
,
D.J.M.
,
Sanz
,
A.
,
Vartiainen
,
S.
,
Kemppainen
,
K.K.
,
Babusiak
,
M.
,
Mustalahti
,
E.
et al (
2009
)
Expression of the ciona intestinalis alternative oxidase (AOX) in drosophila complements defects in mitochondrial oxidative phosphorylation
.
Cell Metab.
9
,
449
460
142
El-Khoury
,
R.
,
Dufour
,
E.
,
Rak
,
M.
,
Ramanantsoa
,
N.
,
Grandchamp
,
N.
,
Csaba
,
Z.
et al (
2013
)
Alternative oxidase expression in the mouse enables bypassing cytochrome c oxidase blockade and limits mitochondrial ROS overproduction
.
PLoS Genet.
9
,
e1003182
143
Szibor
,
M.
,
Dhandapani
,
P.K.
,
Dufour
,
E.
,
Holmström
,
K.M.
,
Zhuang
,
Y.
,
Salwig
,
I.
et al (
2017
)
Broad AOX expression in a genetically tractable mouse model does not disturb normal physiology
.
Dis. Model. Mech.
10
,
163
171
144
Kemppainen
,
K.K.
,
Rinne
,
J.
,
Sriram
,
A.
,
Lakanmaa
,
M.
,
Zeb
,
A.
,
Tuomela
,
T.
et al (
2014
)
Expression of alternative oxidase in drosophila ameliorates diverse phenotypes due to cytochrome oxidase deficiency
.
Hum. Mol. Genet.
23
,
2078
2093
145
Humphrey
,
D.M.
,
Parsons
,
R.B.
,
Ludlow
,
Z.N.
,
Riemensperger
,
T.
,
Esposito
,
G.
,
Verstreken
,
P.
et al (
2012
)
Alternative oxidase rescues mitochondria-mediated dopaminergic cell loss in Drosophila
.
Hum. Mol. Genet.
21
,
2698
2712
146
Langston
,
J.
,
Ballard
,
P.
,
Tetrud
,
J.
and
Irwin
,
I.
(
1983
)
Chronic parkinsonism in humans due to a product of meperidine-analog synthesis
.
Science
219
,
979
980
147
Nicklas
,
W.J.
,
Vyas
,
I.
and
Heikkila
,
R.E.
(
1985
)
Inhibition of NADH-linked oxidation in brain mitochondria by 1-methyl-4-phenyl-pyridine, a metabolite of the neurotoxin, 1-methyl-4-phenyl-1,2,5,6-tetrahydropyridine
.
Life Sci.
36
,
2503
2508
148
Singer
,
T.P.
,
Castagnoli
,
N.
,
Ramsay
,
R.R.
and
Trevor
,
A.J.
(
1987
)
Biochemical events in the development of parkinsonism induced by 1-Methyl-4-Phenyl-1,2,3,6-Tetrahydropyridine
.
J. Neurochem.
49
,
1
8
149
Sun
,
T.-K.
,
Chen
,
Y.-Y.
,
Huang
,
S.-H.
,
Hsu
,
S.-W.
,
Lee
,
C.-C.
,
Chang
,
W.-N.
et al (
2018
)
Neurotoxicity of carbon monoxide targets caudate-mediated dopaminergic system
.
Neurotoxicology
65
,
272
279
150
Zhang
,
D.
,
Lee
,
B.
,
Nutter
,
A.
,
Song
,
P.
,
Dolatabadi
,
N.
,
Parker
,
J.
et al (
2015
)
Protection from cyanide-induced brain injury by the Nrf2 transcriptional activator carnosic acid
.
J. Neurochem.
133
,
898
908
151
Shoffner
,
J.M.
,
Watts
,
R.L.
,
Juncos
,
J.L.
,
Torroni
,
A.
and
Wallace
,
D.C.
(
1991
)
Mitochondrial oxidative phosphorylation defects in Parkinson's disease
.
Ann. Neurol.
30
,
332
339
152
Bonifati
,
V.
,
Rizzu
,
P.
,
van Baren
,
M.J.
,
Schaap
,
O.
,
Breedveld
,
G.J.
,
Krieger
,
E.
et al (
2003
)
Mutations in the DJ-1 gene associated with autosomal recessive early-onset parkinsonism
.
Science
299
,
256
259
153
Liang
,
C.-L.
,
Wang
,
T.T.
,
Luby-Phelps
,
K.
and
German
,
D.C.
(
2007
)
Mitochondria mass is low in mouse substantia nigra dopamine neurons: implications for Parkinson's disease
.
Exp. Neurol.
203
,
370
380
154
Lopez-Fabuel
,
I.
,
Martin-Martin
,
L.
,
Resch-Beusher
,
M.
,
Azkona
,
G.
,
Sanchez-Pernaute
,
R.
and
Bolaños
,
J.P.
(
2017
)
Mitochondrial respiratory chain disorganization in Parkinson's disease-relevant PINK1 and DJ1 mutants
.
Neurochem. Int.
109
,
101
105
155
Lopert
,
P.
and
Patel
,
M.
(
2014
)
Brain mitochondria from DJ-1 knockout mice show increased respiration-dependent hydrogen peroxide consumption
.
Redox Biol.
2
,
667
672
156
Chen
,
S.
,
Annesley
,
S.J.
,
Jasim
,
R.A.F.
,
Musco
,
V.J.
,
Sanislav
,
O.
and
Fisher
,
P.R.
(
2017
)
The Parkinson's disease-associated protein DJ-1 plays a positive nonmitochondrial role in endocytosis in dictyostelium cells
.
Dis. Model. Mech.
10
,
1261
1271
157
Scheltens
,
P.
,
Blennow
,
K.
,
Breteler
,
M.M.B.
,
de Strooper
,
B.
,
Frisoni
,
G.B.
,
Salloway
,
S.
et al (
2016
)
Alzheimer's disease
.
Lancet
388
,
505
517
158
Reiss
,
A.B.
,
Arain
,
H.A.
,
Stecker
,
M.M.
,
Siegart
,
N.M.
and
Kasselman
,
L.J.
(
2018
)
Amyloid toxicity in Alzheimer's disease
.
Rev. Neurosci.
29
,
613
627
159
Todd
,
S.
,
Barr
,
S.
,
Roberts
,
M.
and
Passmore
,
A.P.
(
2013
)
Survival in dementia and predictors of mortality: a review
.
Int. J. Geriatr. Psychiatry
28
,
1109
1124
160
Gottesman
,
R.F.
,
Schneider
,
A.L.C.
,
Albert
,
M.
,
Alonso
,
A.
,
Bandeen-Roche
,
K.
,
Coker
,
L.
et al (
2014
)
Midlife hypertension and 20-year cognitive change: the atherosclerosis risk in communities neurocognitive study
.
JAMA Neurol.
71
,
1218
1227
161
Nyberg
,
J.
,
Åberg
,
M.A.I.
,
Schiöler
,
L.
,
Nilsson
,
M.
,
Wallin
,
A.
,
Torén
,
K.
et al (
2014
)
Cardiovascular and cognitive fitness at age 18 and risk of early-onset dementia
.
Brain
137
,
1514
1523
162
Genin
,
E.
,
Hannequin
,
D.
,
Wallon
,
D.
,
Sleegers
,
K.
,
Hiltunen
,
M.
,
Combarros
,
O.
et al (
2011
)
APOE and Alzheimer disease: a major gene with semi-dominant inheritance
.
Mol. Psychiatry
16
,
903
907
163
Mahley
,
R.W.
and
Huang
,
Y.
(
2012
)
Apolipoprotein E sets the stage: response to injury triggers neuropathology
.
Neuron
76
,
871
885
164
Castellano
,
J.M.
,
Kim
,
J.
,
Stewart
,
F.R.
,
Jiang
,
H.
,
DeMattos
,
R.B.
,
Patterson
,
B.W.
et al (
2011
)
Human apoE isoforms differentially regulate brain amyloid-β peptide clearance
.
Sci. Transl. Med.
3
,
89ra57
89ra57
165
Benilova
,
I.
,
Karran
,
E.
and
Strooper
,
B.D.
(
2012
)
The toxic Aβ oligomer and Alzheimer's disease: an emperor in need of clothes
.
Nat. Neurosci.
15
,
349
357
166
Stefani
,
M.
(
2010
)
Biochemical and biophysical features of both oligomer/fibril and cell membrane in amyloid cytotoxicity
.
FEBS J.
277
,
4602
4613
167
Stéphan
,
A.
,
Laroche
,
S.
and
Davis
,
S.
(
2001
)
Generation of aggregated β-amyloid in the rat hippocampus impairs synaptic transmission and plasticity and causes memory deficits
.
J. Neurosci.
21
,
5703
5714
168
Butterfield
,
S.M.
and
Lashuel
,
H.A.
(
2010
)
Amyloidogenic protein–membrane interactions: mechanistic insight from model systems
.
Angew. Chem. Int. Ed. Engl.
49
,
5628
5654
169
Sheikh
,
K.
,
Giordani
,
C.
,
McManus
,
J.J.
,
Hovgaard
,
M.B.
and
Jarvis
,
S.P.
(
2012
)
Differing modes of interaction between monomeric Aβ1–40 peptides and model lipid membranes: an AFM study
.
Chem. Phys. Lipids
165
,
142
150
170
Gunn
,
A.P.
,
Wong
,
B.X.
,
Johanssen
,
T.
,
Griffith
,
J.C.
,
Masters
,
C.L.
,
Bush
,
A.I.
et al (
2016
)
Amyloid-β peptide Aβ3pE-42 induces lipid peroxidation, membrane permeabilization, and calcium influx in neurons*
.
J. Biol. Chem.
291
,
6134
6145
171
Manczak
,
M.
,
Anekonda
,
T.S.
,
Henson
,
E.
,
Park
,
B.S.
,
Quinn
,
J.
and
Reddy
,
P.H.
(
2006
)
Mitochondria are a direct site of Aβ accumulation in Alzheimer's disease neurons: implications for free radical generation and oxidative damage in disease progression
.
Hum. Mol. Genet.
15
,
1437
1449
172
Reddy
,
P.H.
and
Oliver
,
D.M.
(
2019
)
Amyloid beta and phosphorylated tau-induced defective autophagy and mitophagy in Alzheimer's disease
.
Cells
8
,
488
173
Devi
,
L.
,
Prabhu
,
B.M.
,
Galati
,
D.F.
,
Avadhani
,
N.G.
and
Anandatheerthavarada
,
H.K.
(
2006
)
Accumulation of amyloid precursor protein in the mitochondrial import channels of human Alzheimer's disease brain is associated with mitochondrial dysfunction
.
J. Neurosci.
26
,
9057
9068
174
Petersen
,
C.A.H.
,
Alikhani
,
N.
,
Behbahani
,
H.
,
Wiehager
,
B.
,
Pavlov
,
P.F.
,
Alafuzoff
,
I.
et al (
2008
)
The amyloid β-peptide is imported into mitochondria via the TOM import machinery and localized to mitochondrial cristae
.
Proc. Natl Acad. Sci. U.S.A.
105
,
13145
13150
175
Tillement
,
L.
,
Lecanu
,
L.
and
Papadopoulos
,
V.
(
2011
)
Alzheimer's disease: effects of β-amyloid on mitochondria
.
Mitochondrion
11
,
13
21
176
Parker
,
W.D.
,
Filley
,
C.M.
and
Parks
,
J.K.
(
1990
)
Cytochrome oxidase deficiency in Alzheimer's disease
.
Neurology
40
,
1302
1302
177
Casley
,
C.S.
,
Canevari
,
L.
,
Land
,
J.M.
,
Clark
,
J.B.
and
Sharpe
,
M.A.
(
2002
)
Β-Amyloid inhibits integrated mitochondrial respiration and key enzyme activities
.
J. Neurochem.
80
,
91
100
178
Hernandez-Zimbron
,
L.F.
,
Luna-Muñoz
,
J.
,
Mena
,
R.
,
Vazquez-Ramirez
,
R.
,
Kubli-Garfias
,
C.
,
Cribbs
,
D.H.
et al (
2012
)
Amyloid-β peptide binds to cytochrome C oxidase subunit 1
.
PLos ONE
7
,
e42344
179
Maurer
,
I.
,
Zierz
,
S.
and
Möller
,
H.
(
2000
)
A selective defect of cytochrome c oxidase is present in brain of Alzheimer disease patients
.
Neurobiol. Aging
21
,
455
462
180
Eckert
,
A.
,
Schmitt
,
K.
and
Götz
,
J.
(
2011
)
Mitochondrial dysfunction - the beginning of the end in Alzheimer's disease? separate and synergistic modes of tau and amyloid-β toxicity
.
Alzheimer's Res. Ther.
3
,
15
181
Mastrogiacomo
,
F.
,
Lindsay
,
J.G.
,
Bettendorff
,
L.
,
Rice
,
J.
and
Kish
,
S.J.
(
1996
)
Brain protein and α-ketoglutarate dehydrogenase complex activity in Alzheimer-s disease
.
Ann. Neurol.
39
,
592
598
182
Spuch
,
C.
,
Ortolano
,
S.
and
Navarro
,
C.
(
2012
)
New insights in the amyloid-beta interaction with mitochondria
.
J. Aging Res.
2012
,
324968
183
Bobba
,
A.
,
Amadoro
,
G.
,
Valenti
,
D.
,
Corsetti
,
V.
,
Lassandro
,
R.
and
Atlante
,
A.
(
2013
)
Mitochondrial respiratory chain complexes I and IV are impaired by β-amyloid via direct interaction and through complex I-dependent ROS production, respectively
.
Mitochondrion
13
,
298
311
184
Kumar
,
A.
and
Singh
,
A.
(
2015
)
A review on mitochondrial restorative mechanism of antioxidants in Alzheimer's disease and other neurological conditions
.
Front. Pharmacol.
6
,
206
185
Wang
,
J.
and
Chen
,
G.-J.
(
2016
)
Mitochondria as a therapeutic target in Alzheimer's disease
.
Genes Dis.
3
,
220
227
186
Singer
,
M.
,
Deutschman
,
C.S.
,
Seymour
,
C.W.
,
Shankar-Hari
,
M.
,
Annane
,
D.
,
Bauer
,
M.
et al (
2016
)
The third international consensus definitions for sepsis and septic shock (Sepsis-3)
.
JAMA
315
,
801
810
187
Rhee
,
C.
,
Jones
,
T.M.
,
Hamad
,
Y.
,
Pande
,
A.
,
Varon
,
J.
,
O'Brien
,
C.
et al (
2019
)
Prevalence, underlying causes, and preventability of sepsis-associated mortality in US acute care hospitals
.
Jama Netw. Open
2
,
e187571
188
Singer
,
M.
,
Inada-Kim
,
M.
and
Shankar-Hari
,
M.
(
2019
)
Sepsis hysteria: excess hype and unrealistic expectations
.
Lancet
394
,
1513
1514
189
Rhee
,
C.
,
Dantes
,
R.
,
Epstein
,
L.
,
Murphy
,
D.J.
,
Seymour
,
C.W.
,
Iwashyna
,
T.J.
et al (
2017
)
Incidence and trends of sepsis in US hospitals using clinical vs claims data, 2009–2014
.
JAMA
318
,
1241
190
Liu
,
V.
,
Escobar
,
G.J.
,
Greene
,
J.D.
,
Soule
,
J.
,
Whippy
,
A.
,
Angus
,
D.C.
et al (
2014
)
Hospital deaths in patients With sepsis from 2 independent cohorts
.
JAMA
312
,
90
92
191
Shankar-Hari
,
M.
and
Rubenfeld
,
G.D.
(
2016
)
Understanding long-term outcomes following sepsis: implications and challenges
.
Curr. Infect. Dis. Rep.
18
,
37
192
Cavaillon
,
J.
,
Singer
,
M.
and
Skirecki
,
T.
(
2020
)
Sepsis therapies: learning from 30 years of failure of translational research to propose new leads
.
EMBO Mol. Med.
12
,
e10128
193
Chousterman
,
B.G.
,
Swirski
,
F.K.
and
Weber
,
G.F.
(
2017
)
Cytokine storm and sepsis disease pathogenesis
.
Semin. Immunopathol.
39
,
517
528
194
Stanzani
,
G.
,
Duchen
,
M.R.
and
Singer
,
M.
(
2019
)
The role of mitochondria in sepsis-induced cardiomyopathy
.
Biochim. Biophys. Acta Mol. Basis Dis.
1865
,
759
773
195
Malaisse
,
W.J.
,
Nadi
,
A.B.
,
Ladriere
,
L.
and
Zhang
,
T.-M.
(
1997
)
Protective effects of succinic aciddimethyl ester infusion in experimental endotoxemia
.
Nutrition
13
,
330
341
196
Vanasco
,
V.
,
Saez
,
T.
,
Magnani
,
N.D.
,
Pereyra
,
L.
,
Marchini
,
T.
,
Corach
,
A.
et al (
2014
)
Cardiac mitochondrial biogenesis in endotoxemia is not accompanied by mitochondrial function recovery
.
Free Radic. Bio Med.
77
,
1
9
197
McCreath
,
G.
,
Scullion
,
M.M.F.
,
Lowes
,
D.A.
,
Webster
,
N.R.
and
Galley
,
H.F.
(
2016
)
Pharmacological activation of endogenous protective pathways against oxidative stress under conditions of sepsis
.
Br. J. Anaesth.
116
,
131
139
198
Leite
,
H.P.
and
de Lima
,
L.F.P.
(
2016
)
Metabolic resuscitation in sepsis: a necessary step beyond the hemodynamic?
J. Thorac. Dis.
8
,
E552
E557
199
Bauer
,
M.
,
Press
,
A.T.
and
Trauner
,
M.
(
2013
)
The liver in sepsis: patterns of response and injury
.
Curr. Opin. Crit. Care
19
,
123
127
200
Dorman
,
D.C.
,
Moulin
,
F.J.-M.
,
McManus
,
B.E.
,
Mahle
,
K.C.
,
James
,
R.A.
and
Struve
,
M.F.
(
2002
)
Cytochrome oxidase inhibition induced by acute hydrogen sulfide inhalation: correlation with tissue sulfide concentrations in the rat brain, liver, lung, and nasal epithelium
.
Toxicol. Sci.
65
,
18
25
201
Alonso
,
J.
,
Cardellach
,
F.
,
López
,
S.
,
Casademont
,
J.
and
Miró
,
Ò
. (
2003
)
Carbon monoxide specifically inhibits cytochrome C oxidase of human mitochondrial respiratory chain
.
Pharmacol. Toxicol.
93
,
142
146
202
Sarti
,
P.
,
Forte
,
E.
,
Mastronicola
,
D.
,
Giuffrè
,
A.
and
Arese
,
M.
(
2012
)
Cytochrome c oxidase and nitric oxide in action: molecular mechanisms and pathophysiological implications
.
Biochim. Biophys. Acta Bioenergetics
1817
,
610
619
203
Nisoli
,
E.
,
Clementi
,
E.
,
Paolucci
,
C.
,
Cozzi
,
V.
,
Tonello
,
C.
,
Sciorati
,
C.
et al (
2003
)
Mitochondrial biogenesis in mammals: the role of endogenous nitric oxide
.
Science
299
,
896
899
204
Bauer
,
I.
and
Pannen
,
B.H.
(
2009
)
Bench-to-bedside review: carbon monoxide: from mitochondrial poisoning to therapeutic use
.
Crit. Care
13
,
220
205
Bye
,
A.
,
Sørhaug
,
S.
,
Ceci
,
M.
,
Høydal
,
M.A.
,
Stølen
,
T.
,
Heinrich
,
G.
et al (
2008
)
Carbon monoxide levels experienced by heavy smokers impair aerobic capacity and cardiac contractility and induce pathological hypertrophy
.
Inhal. Toxicol.
20
,
635
646
206
Tiranti
,
V.
,
Viscomi
,
C.
,
Hildebrandt
,
T.
,
Meo
,
I.D.
,
Mineri
,
R.
,
Tiveron
,
C.
et al (
2009
)
Loss of ETHE1, a mitochondrial dioxygenase, causes fatal sulfide toxicity in ethylmalonic encephalopathy
.
Nat. Med.
15
,
200
205
207
Lowes
,
D.A.
,
Webster
,
N.R.
,
Murphy
,
M.P.
and
Galley
,
H.F.
(
2013
)
Antioxidants that protect mitochondria reduce interleukin-6 and oxidative stress, improve mitochondrial function, and reduce biochemical markers of organ dysfunction in a rat model of acute sepsis
.
Br. J. Anaesth.
110
,
472
480
208
Landesberg
,
G.
,
Gilon
,
D.
,
Meroz
,
Y.
,
Georgieva
,
M.
,
Levin
,
P.D.
,
Goodman
,
S.
et al (
2012
)
Diastolic dysfunction and mortality in severe sepsis and septic shock
.
Eur. Heart J.
33
,
895
903
209
Levy
,
R.J.
,
Vijayasarathy
,
C.
,
Raj
,
N.R.
,
Avadhani
,
N.G.
and
Deutschman
,
C.S.
(
2004
)
Competitive and noncompetitive inhibition of myocardial cytochrome C oxidase in sepsis
.
Shock
21
,
110
114
210
Piel
,
D.A.
,
Gruber
,
P.J.
,
Weinheimer
,
C.J.
,
Courtois
,
M.R.
,
Robertson
,
C.M.
,
Coopersmith
,
C.M.
et al (
2007
)
Mitochondrial resuscitation with exogenous cytochrome c in the septic heart*
.
Crit. Care Med.
35
,
2120
2127
211
Piel
,
D.A.
,
Deutschman
,
C.S.
and
Levy
,
R.J.
(
2008
)
Exogenous cytochrome c restores myocardial cytochrome oxidase activity into the late phase of sepsis
.
Shock
29
,
612
616
212
Watts
,
J.A.
,
Kline
,
J.A.
,
Thornton
,
L.R.
,
Grattan
,
R.M.
and
Brar
,
S.S.
(
2004
)
Metabolic dysfunction and depletion of mitochondria in hearts of septic rats
.
J. Mol. Cell. Cardiol.
36
,
141
150
213
Rudiger
,
A.
and
Singer
,
M.
(
2007
)
Mechanisms of sepsis-induced cardiac dysfunction
.
Crit. Care Med.
35
,
1599
1608
214
Metzger
,
N.L.
and
Lundquist
,
L.M.
(
2012
)
A review of the advances in chronic obstructive pulmonary disease treatment
.
J. Pharm. Pract.
25
,
576
582
215
Smith
,
T.C.
,
Heller
,
J.M.
,
Hooper
,
T.I.
,
Gackstetter
,
G.D.
and
Gray
,
G.C.
(
2002
)
Are gulf war veterans experiencing illness due to exposure to smoke from Kuwaiti oil well fires? Examination of department of defense hospitalization data
.
Am. J. Epidemiol.
155
,
908
917
216
Silverman
,
D.T.
,
Samanic
,
C.M.
,
Lubin
,
J.H.
,
Blair
,
A.E.
,
Stewart
,
P.A.
,
Vermeulen
,
R.
et al (
2012
)
The diesel exhaust in miners study: a nested case–control study of lung cancer and diesel exhaust
.
J. Natl Cancer Inst.
104
,
855
868
217
Granados-Castro
,
L.F.
,
Rodríguez-Rangel
,
D.S.
,
Montaño
,
M.
,
Ramos
,
C.
and
Pedraza-Chaverri
,
J.
(
2015
)
Wood smoke exposure induces a decrease in respiration parameters and in the activity of respiratory complexes I and IV in lung mitochondria from Guinea pigs
.
Environ. Toxicol.
30
,
461
471
218
Santoro
,
A.
,
Tomino
,
C.
,
Prinzi
,
G.
,
Lamonaca
,
P.
,
Cardaci
,
V.
,
Fini
,
M.
et al (
2019
)
Tobacco smoking: risk to develop addiction, chronic obstructive pulmonary disease, and lung cancer
.
Recent Pat Anticancer Drug Discov.
14
,
39
52
219
Vega-Sánchez
,
Á.E
,
Téllez-Navarrete
,
N.A.
and
Pérez-Padilla
,
R.
(
2019
)
Chronic obstructive pulmonary disease: perspectives for primary health care
.
Rev. Invest. Clin.
71
,
55
63
220
Pryor
,
W.A.
and
Stone
,
K.
(
1993
)
Oxidants in cigarette smoke radicals, hydrogen peroxide, peroxynitrate, and peroxynitrite
.
Ann. N. Y. Acad. Sci.
686
,
12
27
221
Chung
,
K.F.
and
Adcock
,
I.M.
(
2008
)
Multifaceted mechanisms in COPD: inflammation, immunity, and tissue repair and destruction
.
Eur. Respir. J.
31
,
1334
1356
222
Jiang
,
Y.
,
Wang
,
X.
and
Hu
,
D.
(
2017
)
Mitochondrial alterations during oxidative stress in chronic obstructive pulmonary disease
.
Int. J. Chron. Obstruct. Pulmon. Dis.
12
,
1153
1162
223
Talhout
,
R.
,
Schulz
,
T.
,
Florek
,
E.
,
van Benthem
,
J.
,
Wester
,
P.
and
Opperhuizen
,
A.
(
2011
)
Hazardous compounds in tobacco smoke
.
Int. J. Environ. Res. Public Health
8
,
613
628
224
van der Toorn
,
M.
,
Slebos
,
D.-J.
,
de Bruin
,
H.G.
,
Leuvenink
,
H.G.
,
Bakker
,
S.J.L.
,
Gans
,
R.O.B.
et al (
2007
)
Cigarette smoke-induced blockade of the mitochondrial respiratory chain switches lung epithelial cell apoptosis into necrosis
.
Am. J. Physiol. Lung Cell. Mol. Physiol.
292
,
L1211
L1218
225
Hoffmann
,
R.F.
,
Zarrintan
,
S.
,
Brandenburg
,
S.M.
,
Kol
,
A.
,
de Bruin
,
H.G.
,
Jafari
,
S.
et al (
2013
)
Prolonged cigarette smoke exposure alters mitochondrial structure and function in airway epithelial cells
.
Respir. Res.
14
,
97
226
Ballweg
,
K.
,
Mutze
,
K.
,
Königshoff
,
M.
,
Eickelberg
,
O.
and
Meiners
,
S.
(
2014
)
Cigarette smoke extract affects mitochondrial function in alveolar epithelial cells
.
Am. J. Physiol.
307
,
L895
L907
227
Cloonan
,
S.M.
,
Glass
,
K.
,
Laucho-Contreras
,
M.E.
,
Bhashyam
,
A.R.
,
Cervo
,
M.
,
Pabón
,
M.A.
et al (
2016
)
Mitochondrial iron chelation ameliorates cigarette smoke–induced bronchitis and emphysema in mice
.
Nat. Med.
22
,
163
174
228
Messner
,
B.
,
Frotschnig
,
S.
,
Steinacher-Nigisch
,
A.
,
Winter
,
B.
,
Eichmair
,
E.
,
Gebetsberger
,
J.
et al (
2012
)
Apoptosis and necrosis: two different outcomes of cigarette smoke condensate-induced endothelial cell death
.
Cell Death Dis.
3
,
e424
229
Giordano
,
L.
,
Deceglie
,
S.
,
d'Adamo
,
P.
,
Valentino
,
M.L.
,
Morgia
,
C.L.
,
Fracasso
,
F.
et al (
2015
)
Cigarette toxicity triggers leber's hereditary optic neuropathy by affecting mtDNA copy number, oxidative phosphorylation and ROS detoxification pathways
.
Cell Death Dis.
6
,
e2021
230
Rao
,
P.
,
Ande
,
A.
,
Sinha
,
N.
,
Kumar
,
A.
and
Kumar
,
S.
(
2016
)
Effects of cigarette smoke condensate on oxidative stress, apoptotic cell death, and HIV replication in human monocytic cells
.
PLos ONE
11
,
e0155791
231
Zhang
,
X.W.
,
Liu
,
Q.
,
Wang
,
Y.
and
Thorlacius
,
H.
(
2001
)
CXC chemokines, MIP-2 and KC, induce P-selectin-dependent neutrophil rolling and extravascular migration in vivo
.
Br. J. Pharmacol.
133
,
413
421
232
Thatcher
,
T.H.
,
McHugh
,
N.A.
,
Egan
,
R.W.
,
Chapman
,
R.W.
,
Hey
,
J.A.
,
Turner
,
C.K.
et al (
2005
)
Role of CXCR2 in cigarette smoke-induced lung inflammation
.
Am. J. Physiol.
289
,
L322
L328
233
Papaporfyriou
,
A.
,
Loukides
,
S.
,
Kostikas
,
K.
,
Simoes
,
D.C.M.
,
Papatheodorou
,
G.
,
Konstantellou
,
E.
et al (
2014
)
Increased levels of osteopontin in sputum supernatant in patients with COPD
.
Chest
146
,
951
958
234
Grady
,
R.W.
,
Bienen
,
E.J.
and
Clarkson
,
A.B.
(
1986
)
Esters of 3,4-dihydroxybenzoic acid, highly effective inhibitors of the sn-glycerol-3-phosphate oxidase of Trypanosoma brucei brucei
.
Mol. Biochem. Parasitol.
21
,
55
63
235
Clarkson
,
A.B.
,
Bienen
,
E.J.
,
Pollakis
,
G.
and
Grady
,
R.W.
(
1989
)
Trypanocidal CoQ analogues: their effect on other mitochondrial systems
.
Comp. Biochem. Physiol. B
94
,
245
251
236
Grady
,
R.W.
,
Bienen
,
E.J.
,
Dieck
,
H.A.
,
Saric
,
M.
and
Clarkson
,
A.B.
(
1993
)
N-n-alkyl-3,4-dihydroxybenzamides as inhibitors of the trypanosome alternative oxidase: activity in vitro and in vivo
.
Antimicrob. Agents Chemother.
37
,
1082
1085
237
Young
,
L.
,
May
,
B.
,
Pendlebury-Watt
,
A.
,
Shearman
,
J.
,
Elliott
,
C.
,
Albury
,
M.S.
et al (
2014
)
Probing the ubiquinol-binding site of recombinant Sauromatum guttatum alternative oxidase expressed in E. coli membranes through site-directed mutagenesis
.
Biochim. Biophys. Acta Bioenergetics
1837
,
1219
1225
238
Hoefnagel
,
M.H.N.
,
Wiskich
,
J.T.
,
Madgwick
,
S.A.
,
Patterson
,
Z.
,
Oettmeier
,
W.
and
Rich
,
P.R.
(
1995
)
New inhibitors of the ubiquinol oxidase of higher plant mitochondria
.
Eur. J. Biochem.
233
,
531
537
239
James
,
T.W.
and
Spencer
,
M.S.
(
1982
)
Inhibition of cyanide-resistant respiration in pea cotyledon mitochondria by chloroquine
.
Plant Physiol.
69
,
1113
1115
240
Balogun
,
E.O.
,
Inaoka
,
D.K.
,
Shiba
,
T.
,
Tsuge
,
C.
,
May
,
B.
,
Sato
,
T.
et al (
2019
)
Discovery of trypanocidal coumarins with dual inhibition of both the glycerol kinase and alternative oxidase of Trypanosoma brucei brucei
.
FASEB J.
33
,
13002
13013
241
Grover
,
S.D.
and
Laties
,
G.G.
(
1981
)
Disulfiram inhibition of the alternative respiratory pathway in plant mitochondria
.
Plant Physiol.
68
,
393
400
242
Janes
,
H.W.
and
Wiest
,
S.C.
(
1982
)
Inhibition of O2 consumption resistant to cyanide and its development by n-propyl gallate and salicylhydroxamic acid
.
Plant Physiol.
70
,
853
857
243
Schonbaum
,
G.R.
,
Bonner
,
W.D.
,
Storey
,
B.T.
and
Bahr
,
J.T.
(
1971
)
Specific inhibition of the cyanide-insensitive respiratory pathway in plant mitochondria by hydroxamic acids
.
Plant Physiol.
47
,
124
128
244
Kay
,
C.J.
and
Palmer
,
J.M.
(
1985
)
Solubilization of the alternative oxidase of cuckoo-pint (Arum maculatum) mitochondria. stimulation by high concentrations of ions and effects of specific inhibitors
.
Biochem. J.
228
,
309
318
245
Sluse
,
F.E.
,
Almeida
,
A.M.
,
Jarmuszkiewicz
,
W.
and
Vercesi
,
A.E.
(
1998
)
Free fatty acids regulate the uncoupling protein and alternative oxidase activities in plant mitochondria
.
FEBS Lett.
433
,
237
240
246
Milani
,
G.
,
Jarmuszkiewicz
,
W.
,
Sluse-Goffart
,
C.M.
,
Schreiber
,
A.Z.
,
Vercesi
,
A.E.
and
Sluse
,
F.E.
(
2001
)
Respiratory chain network in mitochondria of Candida parapsilosis: ADP/O appraisal of the multiple electron pathways
.
FEBS Lett.
508
,
231
235
247
Costa
,
P.C.S.
,
Barsottini
,
M.R.O.
,
Vieira
,
M.L.L.
,
Pires
,
B.A.
,
Evangelista
,
J.S.
,
Zeri
,
A.C.M.
et al (
2020
)
N-Phenylbenzamide derivatives as alternative oxidase inhibitors: synthesis, molecular properties, 1H-STD NMR, and QSAR
.
J. Mol. Struct.
1208
,
127903
248
Cook
,
N.D.
and
Cammack
,
R.
(
1983
)
Effects of the quinone analogue 5-n-undecyl-6-hydroxy-4,7-dioxobenzothiazole on cyanide-sensitive and cyanide-insensitive plant mitochondria
.
Biochem. Soc. Trans.
11
,
785
785

Author notes

*

Present address: Cristália Produtos Químicos e Farmacêuticos Ltda, Biotechnology Division, Rod Itapira-Lindoia Km 14, Itapira – SP, 13974-900, Brazil

This is an open access article published by Portland Press Limited on behalf of the Biochemical Society and distributed under the Creative Commons Attribution License 4.0 (CC BY).