Next Article in Journal
HBM4EU Chromates Study: Urinary Metabolomics Study of Workers Exposed to Hexavalent Chromium
Previous Article in Journal
Metabolomic Associations of Asthma in the Hispanic Community Health Study/Study of Latinos
Previous Article in Special Issue
Tolerant Larvae and Sensitive Juveniles: Integrating Metabolomics and Whole-Organism Responses to Define Life-Stage Specific Sensitivity to Ocean Acidification in the American Lobster
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Exploring Thermal Sensitivities and Adaptations of Oxidative Phosphorylation Pathways

1
Faculty Saint-Jean, Department of Medicine, Faculty of Medicine and Dentistry, University of Alberta, Edmonton, AB T6C 4G9, Canada
2
Department Biologie, Université du Québec à Rimouski, Rimouski, QC G5L 3A1, Canada
*
Authors to whom correspondence should be addressed.
Metabolites 2022, 12(4), 360; https://doi.org/10.3390/metabo12040360
Submission received: 22 March 2022 / Revised: 8 April 2022 / Accepted: 11 April 2022 / Published: 17 April 2022
(This article belongs to the Special Issue Ectotherms Metabolism: Plasticity and Adaptation)

Abstract

:
Temperature shifts are a major challenge to animals; they drive adaptations in organisms and species, and affect all physiological functions in ectothermic organisms. Understanding the origin and mechanisms of these adaptations is critical for determining whether ectothermic organisms will be able to survive when faced with global climate change. Mitochondrial oxidative phosphorylation is thought to be an important metabolic player in this regard, since the capacity of the mitochondria to produce energy greatly varies according to temperature. However, organism survival and fitness depend not only on how much energy is produced, but, more precisely, on how oxidative phosphorylation is affected and which step of the process dictates thermal sensitivity. These questions need to be addressed from a new perspective involving a complex view of mitochondrial oxidative phosphorylation and its related pathways. In this review, we examine the effect of temperature on the commonly measured pathways, but mainly focus on the potential impact of lesser-studied pathways and related steps, including the electron-transferring flavoprotein pathway, glycerophosphate dehydrogenase, dihydroorotate dehydrogenase, choline dehydrogenase, proline dehydrogenase, and sulfide:quinone oxidoreductase. Our objective is to reveal new avenues of research that can address the impact of temperature on oxidative phosphorylation in all its complexity to better portray the limitations and the potential adaptations of aerobic metabolism.

Graphical Abstract

1. Introduction

Anticipated global climate changes will be a major threat to ectothermic organisms and the overall biodiversity of the planet [1]. For different species, the effects of these environmental alterations will depend on the range of temperatures they can tolerate (thermal tolerance), the impact of the temperature changes on their physiological performance and life history traits, and their ability to adjust to new environmental conditions. Therefore, the thermal sensitivity of different physiological functions, and their capacity to compensate for increased temperatures, might dictate the long-term consequences of global climate changes on their survival.
The bioenergetics of organisms—particularly cellular bioenergetics upstream of a complex physiological function—is affected by temperature [2]. Changes in temperature, therefore, necessarily impact the whole organism through the thermal sensitivity of cellular bioenergetics machinery, particularly the mitochondria [3]. There is also some evidence that impairment of the mitochondrial structure or function might set the upper thermal limits of ectotherms [4,5]. Furthermore, temperature acclimation is associated with changes in mitochondrial membrane composition that could lead to changes in enzyme activities [2].
Documenting to what extent OXPHOS is affected by temperature in different species from different thermal habitats could, therefore, be a simple way to assess mitochondrial adaptability. Another approach would be to identify the steps and pathways that are more affected by temperature and could impede ATP production or ROS management. The identification of these steps, combined with a delineation of divergences in mitochondrial architecture and organization in different species adapted to different thermal niches, could identify potential selection sites. Further studies on the genetic and phenotypic variability of these traits, as well as their heritability, are a prerequisite for estimating their adaptability and, therefore, the possibilities and constraints for the evolution of cellular bioenergetics. For example, studies on Drosophila have demonstrated that different mitochondrial haplotypes are associated with different thermal sensitivities [6,7].
Mitochondrial DNA encodes for important peptides of the ETS complexes and ATP synthase; these peptides are involved in the metabolic adaptations of organisms [8]. In combination with these mtDNA-encoded peptides, mitochondria also require hundreds of nuclear genes (proteins) to orchestrate the assembly of fully functional mitochondria. The identification of all genes and proteins involved in the potential adaptations of new phenotypes through genomic and proteomic approaches will be more complex than just comparing metabolic phenotypes associated with potential adaptation and mtDNA (see [9]). Accordingly, the identification of sensitive steps through extensive characterization of OXPHOS capacity when mitochondria are presented with different metabolites, inhibitors, and uncouplers at different temperatures [10], the measurement of changes in various species following thermal acclimation, and the comparison of species from different thermal habitats, are powerful and simpler approaches for identifying sensitive or plastic traits and loci that are likely under selective pressure.
We first need, however, to identify the relevant pathways used by different species or tissues and determine how these pathways react to short- (acute effects) and long- (adaptation) term variations in temperature. Traditionally, OXPHOS measurement was limited to the quantification of mitochondrial oxygen consumption after providing substrates for the reduction of complex I (NADH dehydrogenase; NADH pathway) or complex II (succinate dehydrogenase; succinate pathway), separately at first, and more recently in combination. These two pathways deliver electrons to the ubiquinone pool (Q-junction) that sequentially supplies complex III, cytochrome c, complex IV, and molecular oxygen. While studies on the thermal sensitivity of mitochondria focused on the NADH and succinate pathways in the last few decades, recent studies have concentrated on the combination of these two pathways, or have included other pathways, such as electron-transferring flavoprotein (ETF), mitochondrial glycerol-3-phosphate dehydrogenase (mGpDH), proline dehydrogenase (ProDH), choline dehydrogenase (ChoDH), dihydroorotate dehydrogenase (DhoDH), and sulfide:quinone oxidoreductase (SQOR) pathways (Figure 1). Recent progress has shown that these alternate pathways, in combination with canonic pathways, can result in significantly higher capacities to reduce the Q-junction (reviewed by [11]). It is likely that the evolution of these pathways, as well as their orchestration and combination, could lead to new metabolic adaptations. The aim of our paper is to review these understudied pathways and their thermal sensitivities, and to suggest rigorous approaches to include them when exploring potential adaptations of OXPHOS to environmental change.

2. NADH and Succinate Pathways

The most extensive literature available concerning the effect of temperature on mitochondrial function focuses on the NADH and succinate pathways, along with their closely associated steps. Some reviews are already available on these topics [2,3], and we will not add further details here. Briefly, the pyruvate dehydrogenase complex, a key step in pyruvate oxidation that feeds the NADH pathway, has been found to be largely responsible for respiratory limitation at low temperatures in warm-adapted species, such as rats [10]. Low-temperature respiratory limitation by pyruvate dehydrogenase has been overcome in cold-adapted species, such as the Atlantic wolffish (Anarhichas lupus) [13]. The importance of pyruvate dehydrogenase in regulating temperature preference is also supported by a study on a cold-seeking mutant of Drosophila melanogaster, which showed that the mutant with a preference for a cold environment had higher activity in the pyruvate dehydrogenase complex (PDC) [14]. Once the ectotherm species have overcome the limitation by PDC in the cold environment, complex IV thermal sensitivity becomes highly correlated with the thermal sensitivity of mitochondrial respiration; this relationship is not observed in endotherm species [2]. Furthermore, results from titrations of complex IV inhibitors at various temperatures in the red muscle of brook trout (Salvelinus fontinalis) suggest that complex IV exerts a high control coefficient on the ETS in the normal temperature range [15], but that the control of mitochondrial catalytic capacities might switch at lower temperatures to processes upstream of the ETS, such as the PDC [2].
Succinate and succinate dehydrogenase also contribute to electron transfer through FAD at the Q-junction, leading to an increased respiration rate. This additive respiration varies depending on species, strains, organs, and experimental conditions. In endotherms, the respiration rate following succinate addition to the mitochondria already fed at the complex I level is higher by 1.6- to 2.0-fold in rat heart [16,17], 1.2- to 1.8-fold in rat skeletal muscle [18,19,20], 1.4-fold in mouse skeletal muscle [21], and 1.3- to 2.1-fold in human skeletal muscle (reviewed by [22]). In fish heart mitochondria, the increase in OXPHOS resulting from the addition of succinate is rather variable, depending on species, time, and temperature. In salmon (Salmo salar) acclimated to 12 °C, the increase goes from 1.5-fold when measured at 20 °C, to approximately 2-fold when measured between 24 °C and 28 °C. When acclimated to 20 °C, the increase is lower, from 1.12-fold when measured at low temperatures, to 1.4-fold when measured at higher temperatures [23]. In European perch (Perca fluviatilis), the increment is lower compared to salmon and mammal tissues, and it is similar at two temperatures of measurement (i.e., from 1.10- to 1.12-fold at 16 °C and from 1.06- to 1.11-fold at 23 °C) [24]. In rainbow trout (Oncorhynchus mykiss) heart mitochondria, the increment was close to 1.2-fold between 5 °C and 30 °C [25] and between 1.1- and 1.2-fold in the wrasse (Notolabrus celidotus) when measured between 15 °C and 32.5 °C [26]. The magnitude of increment in respiratory capacity ensured by SDH activity is, therefore, species- and temperature-dependent, and is also modulated by acclimation temperature, at least in Atlantic salmon (Salmo salar). It has been reported that this reserve respiratory capacity, might be vital for cardiac myocytes [27], particularly for their ability to survive hypoxia [28], and potentially for their ability to survive temperature variation.
In the ETS, complexes I and III are known as the primary sites of ROS generation, with the suspected participation of SDH through the reverse flow of electrons to complex I generation sites [29,30,31,32]. It was later observed that SDH generates ROS through its FAD site [33,34,35]. Mitochondrial ROS management is of crucial importance for the governance of different pathways, not only in the aging process [36], but also for the tolerance of ectotherms to increased temperatures [4]. We can therefore speculate that the metabolic requirements of ROS could partly dictate ETS architecture, and particularly the relative quantity or activity of the different complexes [37]. In a comparative analysis of two species of bivalve (Arctica islandica and Mya arenaria), Munro et al. [38] observed a much lower rate of H2O2 efflux in the mitochondria of the long-lived species (A. islandica; maximum lifespan 507 years) compared to the short-lived species, which was also associated with considerably lower SDH activity relative to cytochrome c oxidase or citrate synthase. ROS management has been extensively associated with organism lifespans, and the activities of the major mitochondrial ROS producers (CI and CIII) have been associated with lifespan modulation (see [36,39] for reviews). These results, and the involvement of SDH in ROS regulation, call for a better portrayal of the relative contribution of these ETS complexes, including SDH, in future studies seeking to determine how mitochondria are precisely tailored and adapted to different environmental conditions.

3. Electron-Transferring Flavoprotein Pathway

The electron-transferring flavoprotein pathway (see details in Figure 1) is of major importance in many tissues and species. A few preliminary observations support the importance of fatty acid catabolism in adjustments of mitochondrial functions that occur during temperature variations. A recent study on zebrafish (Danio rerio) showed that (1) fasting enhances acute cold resistance, and (2) stimulation of lipid catabolism is associated with the improvement of cold resistance [40]. Interestingly, several observations in zebrafish have linked cold resistance to lipid metabolism, including (1) the inhibition of fatty acid ß-oxidation by the suppression of the mTOR pathway weakened fasting-induced cold resistance, (2) the stimulation of fatty acid ß-oxidation using fenofibrate increased survival rate following cold stress, and (3) the suppression of fatty acid ß-oxidation using mildronate caused a decrease in survival following cold stress [40]. These results clearly point out the importance of modulating fatty acid ß-oxidation for animal resistance to cold temperatures.
Exposure to cold temperatures has also been associated with changes in the activities of various enzymes involved in lipid metabolism in animals, especially carnitine palmitoyl transferase (CPT) and ß-hydroxyacyl-CoA dehydrogenase (HOAD), a mitochondrial matrix enzyme involved in ß-oxidation of fatty acids. For example, there is an increase in red muscle CPT1 in striped bass (Morone saxatilis) [41], and muscle HOAD in rainbow trout (Oncorhynchus mykiss) [42] and sea bream (Sparus aurata) [43], following cold acclimation. Similarly, the activities of muscle CPT1 and ß-hydroxyacyl-CoA dehydrogenase are increased in fish species from Antarctic regions compared to species from temperature zones [44], and in horse mussels (Modiolus modiolus) collected in winter compared to those collected in summer [45]. Higher muscle HOAD activity was also observed in cold-acclimated striped marsh frogs (Limnodynastes peronii) compared to warm-acclimated frogs [46]. Furthermore, another study revealed higher expression of genes linked to fatty acid metabolism, such as fatty acid binding protein 1 (FABP1) and/or acyl-coA binding protein (ACBP), in the liver and brain of carp (Cyprinus carpio L.) following cold adaptation [47]. These results are in line with a recent comparative study on the hearts of eight fish species collected over a wide environmental temperature gradient (optimal temperatures ranging from −1 °C to above 30 °C) that revealed much higher relative activities of CPT1 and HOAD (when normalized to citrate synthase activity) in cold-adapted fishes compared to warm-adapted ones [37]. Overall, these results suggest an increase in fatty acid oxidation when animals are exposed to cold temperatures.
Very few comparative studies on the impact of acute change of temperature on the metabolism of intact mitochondria have focused on fatty acid catabolism. Rare examples include a study on red muscle mitochondria from shorthorn sculpin (Myoxocephalus scorpius (Linnaeus, 1758)) [48]. This study demonstrated that the thermal sensitivity (expressed by the Q10) of palmitoylcarnitine oxidation, a long-chain fatty acid, is reduced by half after cold acclimation. However, mitochondrial respiration measurement with palmitoylcarnitine + malate as the only substrates cannot reveal at which step the modification occurs, i.e., adjustments in CACT, CPT2, long-chain fatty acid oxidation, or ETF can cause the observed compensation (Figure 1). To address this shortcoming, a recent study on the planarian Dugesia tigrina evaluated the impact of thermal acclimation on mitochondrial respiration using three sets of fatty acid ß-oxidation substrates: palmitoylcarnitine + malate, octanoylcarnitine + malate, and acetylcarnitine + malate (Mast and Lemieux, resubmitted). This allowed the assessment of the temperature effect on ß-oxidation for specific chain lengths (long- or medium-chain fatty acids, with palmitoylcarnitine or octanoylcarnitine, respectively) and on specific enzymes involved in the pathway, such as carnitine palmitoyltransferase 2 (CPT2; only required for palmitoylcarnitine and octanoylcarnitine oxidation), carnitine–acylcarnitine translocase (CACT; required for combinations of the three substrates), or carnitine acyltransferase (CAT; only required for acetylcarnitine oxidation). This study identified the medium-chain fatty acid oxidation capacity—but not the long-chain fatty acid oxidation capacity—as being adjusted following cold-acclimation in D. tigrina. To our knowledge, this is the only study that examines thermal sensitivity and the impact of thermal acclimation on multiple steps of fatty acid ß-oxidation in intact, functional mitochondria. More studies on substrate combinations in multiple tissues and different species are required to document the thermal sensitivity and plasticity of fatty acid oxidation and, therefore, the adaptability of the process.
Other mitochondrial enzymes that can sustain electron transport via ETF are enzymes involved in one-carbon metabolism, i.e., dimethylglycine dehydrogenase (DMGDH) and sarcosine dehydrogenase (SARDH) [49,50]. One-carbon metabolism includes the methionine and folate cycles, which provide one carbon unit (methyl group) during the synthesis of DNA, polyamines, amino acids, creatine, and phospholipids (for a review, see [51]). Both enzymes are also involved in choline metabolism. In mammalian liver cells, dimethylglycine and sarcosine are produced by choline oxidation [52,53,54]. Choline is involved in different physiological processes, including the transport of lipids, methylation reaction, neurotransmitter synthesis, and membrane phospholipid synthesis. Since choline is a micronutrient in different animal species, the function of DMGDH and SARDH in choline metabolism is oriented either to choline concentration adjustment, or to the production of sarcosine or glycine. Modulation of these enzymes is likely controlled by the requirements of anabolic or regulation pathways (regulation of choline, dimethylglycine, and sarcosine concentration and synthesis) rather than bioenergetic demands. Considering the role of choline in phospholipid synthesis, the modulation of the choline concentration by the regulation of DMGDH and SARDH activities could partly be involved in forming efficient mitochondrial membranes in ectotherms to face different thermal regimes. For example, mitochondrial respiratory capacity in the eurythermal teleost Fundulus heteroclitus is linked to the restructuring of membrane lipid composition via local adaptation as well as thermal acclimation [55].

4. Glycerophosphate Dehydrogenase

Mitochondrial glycerophosphate dehydrogenase (mGpDH, E.C. 1.1.99.5), found at the outer portion of the inner mitochondrial membrane [56], irreversibly oxidizes glycerol-3-phosphate to dihydroxyacetone phosphate (DHAP), feeding electrons from FADH2 into the Q-junction. The mGpDH functions together with the cytoplasmic form of the same enzyme (cGpDH, E.C. 1.1.1.8), which catalyzes the reverse conversion of DHAP into glycerol-3-phosphate (G3P), using NADH as the electron donor. In addition, mGpDH is also known as the rate-limiting enzyme for the glycerophosphate shuttle [57]. This enzyme shuttle, composed of mGpDH and cGpDH, is present in cells from a diverse range of animals, including insects [58], fish [59,60], yeast (see a review by [61]), and mammals [62,63,64,65,66].
A major function of mGpDH is likely the reoxidation of cytosolic NADH produced by glycolysis, and the transfer of reducing equivalents from the cytosol to the mitochondrial electron transport system [67]. However, its wide distribution in organisms, and its large variation at the tissue level, implies other functions [68]. The substrate for mGpDH—glycerol-3-phosphate—is, in fact, a pivotal metabolite that links carbohydrate to lipid metabolism. In addition to its link with OXPHOS, it can supply the carbon for gluconeogenesis and/or act as a backbone in triacylglycerol synthesis. It may also be provided as a substrate following the mobilization and lipolysis of acylglycerols, leading to free fatty acids and glycerol. This makes mGpDH a critical component in metabolism, being at the central junction between OXPHOS, glycolysis, and fatty acid metabolism [57,69,70,71,72]. Thus, it is not surprising that, changes in the expression of mGpDH have also been observed in several pathologies, such as cancer and metabolic diseases (reviewed by [68]). In recent studies, G3P has been proposed as a major substrate for mitochondrial respiration in some insects [73] and bivalves [74,75].
Very little work has been conducted on the impact of temperature on the GpDH shuttle; however, there is some indication that this enzyme system could be involved in the response to temperature changes. In a recent study on the yeast Saccharomyces, mGpDH activity was listed as one of 46 reactions important for growth at low temperatures [76]. A recent study showed that when various Drosophila species are held above their normal temperature range, NADH pathway (dependent on complex I) capacity is sharply reduced, but compensated for by a higher reliance on alternative pathways, such as mGpDH and ProDH, that act to maintain maximal OXPHOS capacity [77]. An increase in the expression of both cGpDH and mGpDH has been observed with increasing latitude (associated with decreasing temperatures) in Drosophila populations from southern Florida to Ontario [78]. This increase has been associated with the availability of nutrients along the latitudinal tropical–temperate gradient [78]. Thus, it seems that temperature along these gradients could lead to mitochondrial adjustments in pathway organization in ectotherms. Furthermore, mGpDH has been shown to mediate thermogenesis through modulation of activity in mouse brown adipose tissue following exposure to cold [79]. In these conditions, brown adipose tissue releases almost all the energy transduced by OXPHOS as heat via the uncoupling protein 1. The glycerophosphate shuttle supports brown adipose tissue thermogenesis by feeding additional reducing equivalents to the mitochondria [11]. In addition, knockout of mouse mGpDH leads to reduced energy turnover (food intake and oxygen consumption), increased thyroidal secretion, brown adipose tissue atrophy, and increased of uncoupling protein 3 when exposed to 32 °C, suggesting an attempt to compensate for a thermogenic deficiency [80].
Due to the fact that mGpDH function is affiliated with the cytosolic form cGpDH, we can expect that if one isozyme is affected by temperature change, the other would also be affected. There is some evidence that the cytosolic form of this enzyme is involved in the temperature acclimation of animal metabolism. First, the enzyme is activated following cold acclimation in endotherms. cGpDH in mice has been shown to greatly increase in the brown adipose tissue after exposure to cold conditions [81,82]. cGpDH gene expression also increases in the brain and liver of carp (Cyprinus carpio L.) following cold acclimation [47]. Second, cGpDH has been linked to resistance to cold damage. The loss of cGpDH decreases tolerance to freezing in wildtype Saccharomyces cerevisiae cells preincubated at low temperatures, but not in cells directly transferred from 30 °C to −20 °C [83]. This indicates that the mechanism behind the loss of tolerance requires time, and might be explained by processes dependent on cGpDH, but not necessarily the activity of the enzyme. This study identified the high-osmolarity glycerol pathway as an important process in (1) the transmission of the cold signal, (2) the regulation of expression in a subset of cold-induced genes (including cGpDH), and (3) the determination of freezing tolerance. Third, cGpDH has been identified as an enzyme that needs to be adjusted in hibernating mammals. Hibernation is not just a physiological process that lowers body temperature, it also modulates metabolism during a long period of starvation. During this starvation, new sources of glucose are required, and glycerol metabolism has been shown to provide 40–80% of this [84,85]. Glycerol is first phosphorylated by glycerol kinase to form glycerol-3-phospĥate, and is then converted by the reversible reaction of cGpDH into dihydroxyacetone phosphate (DHAP), an intermediate of gluconeogenesis and glycolysis [86]. When jerboas (Jaculus orientalis) enter hibernation, cGpDH can fuel catabolism by increasing activity up to two- to three-fold in various tissues, e.g., brown adipose tissue, skeletal muscle, kidneys, brain, and liver [87]. Furthermore, the liver mGpDH of hibernating ground squirrel (Urocitellus richardsonii) maintains better functionality at low temperature (5 °C) compared to the euthermic state form (higher affinities for glycerol-3-phosphate and NAD+), and displays lower perturbations when exposed to high temperature (50 °C) [88]. Another example is the black-tailed prairie dog (Cynomys ludivicianus); muscle cGpDH in this hibernator is structurally less rigid and maintains better functional integrity over a wide range of temperatures, has greater stability at high temperature, and is less sensitive to chemical denaturation compared to this enzyme in rabbits [86]. These specific properties may grant sustained enzyme functions over the wide range of body temperatures experienced by animals as they undergo the hibernating season.
From these studies, it appears that mitochondrial ability to oxidize G3P should be considered in future studies on either the thermal sensitivity of mitochondrial metabolism, or the adaptive or acclimatory responses to a changing environment. Another relevant reason to include G3P in the protocol of mitochondrial respiration is to ensure a proper estimation of maximal ETS capacity. An estimation of maximal ETS capacity is critical when evaluating the relative importance of different electron entrance pathways, for example, the NADH/CI or the FADH2/CII pathways. Respiration rates will often be normalized by the maximal oxygen consumption allowed by the ETS. Maximal ETS activity is obtained after saturating the Q pool with electrons (providing many of the mitochondrial substrates) and adding a membrane uncoupler that allows protons to leak into the matrix from outside the inner membrane, thus bypassing ATP synthase. Adequate saturation of the Q pool is, therefore, of critical importance for accurately estimating ETS capacity. It has been observed in invertebrates that the addition of G3P substrates can substantially increase the respiration rates of mitochondria that have already been provided with NADH and succinate pathway substrates [74,75,77]; consequently, we assume that feeding only these two pathways would significantly underestimate maximum ETS capacity. Accurately assessing the relative importance of different pathways feeding reducing equivalents to mitochondria at different temperatures requires proper normalization of the maximal ETS capacity, and, therefore, G3P may be needed as a substrate. For the same reasons, G3P could be useful for acclimation and adaptation (comparative) studies.

5. Dihydroorotate Dehydrogenase

Dihydroorotate (Dho) catabolism is an understudied pathway. To our knowledge, there is no study on the effect of temperature on Dho entry into the ETS, and there are reasons to believe that the research community would benefit from more information on this topic. This substrate is linked with many aspects of mitochondrial and cellular metabolism, and it is present in a wide variety of organisms. Here, we are considering dihydroorotate dehydrogenase (DhoDH) in animals. Bacteria and lower eukaryotes (e.g., S. cerevisiae) also have a form of DhoDH, but it is located in the cytoplasm, and uses either NAD+ or fumarate as an electron acceptor [89,90].
The Dho pathway in animals is driven by the enzyme DhoDH, a flavin mononucleotide (FMN) protein present in the outer phase of the inner mitochondrial membrane. It is the fourth step of the de novo pyrimidine synthesis pathway, it is vital for RNA and DNA synthesis, and it is the only step of this pathway within mitochondria [91]. DhoDH removes two electrons from Dho, converting it to orotate using its redox-active FMN prosthetic group; this requires a suitable electron acceptor to efficiently transfer the electron. This electron acceptor is the Q-junction of the mitochondrial ETS, which contributes to generating the electrochemical gradient [92]. It thus provides a link between the OXPHOS process and the control of biogenesis [93,94]. Pyrimidine nucleotides also serve essential functions in the activation mechanism of sugars for posttranslational glycosylation of proteins and lipids, and for phospholipid biosynthesis [95].
The Dho pathway has gained a great deal of interest recently because of its broad immunosuppressive effect in vivo, making it a promising therapeutic target for the treatment of cancer, viral infections, and auto-immune diseases [11,96,97]. Studies employing cell bypassing coenzyme Q oxidation have suggested that the only essential function of the ETS for the growth of some cancer cells is to preserve DhoDH activity [98,99], emphasizing the importance of this pathway when conditions change. Studies on the Dho pathway indicate that DhoDH inhibition can cause mitochondrial dysfunction other than what is directly linked to this pathway. First, DhoDH inhibition can induce the expression of GDF15 [91], one of the best circulating markers of mitochondrial dysfunction [100]. Second, even though the direct contribution of DhoDH to ATP generation is only marginal in various cancer cells at baseline (5–10% of routine respiration, coupled with ATP production [101]), it can vary depending on conditions. In fact, in the study of Zhang et al. [91], specific inhibition of DhoDH in cells led to a much greater reduction in mitochondrial respiration, suggesting that other processes, additional to the DhoDH pathway, were affected. Due to the fact that only routine and uncoupled respiration was used to estimate mitochondrial respiration in a standard cell culture medium (Dulbecco’s Modified Eagle Medium, DMEM), the data gave no indication as to which specific pathways were affected by the DhoDH inhibition. Third, the impact of DhoDH knockdown included a slowing of cell growth, the partial inhibition of complex III (affecting all other respiration pathways), a decrease in mitochondrial membrane potential, and an increase in mitochondrial ROS production [102]. Using immunoprecipitation and blue native-SDS/PAGE, the same study also showed that DhoDH interacts with complexes II and III [102], again suggesting potential effects beyond the provision of electrons to the Q pool via DhoDH. Fourth, DhoDH is downregulated in oral squamous carcinoma cells [103]; the loss of functional DhoDH not only impedes the de novo pyrimidine synthesis in these cells, it also disrupts mitochondrial respiration by destabilizing mitochondrial contact sites and the cristae organizing system, as well as mitochondrial homeostasis.
Even though there is no data on the effect of temperature on DhoDH, we can expect that its thermal sensitivity will affect more mitochondrial functions than only the reduction of the Q pool. Furthermore, even if it is loosely associated with the intermembrane face of the inner mitochondrial membrane, the DhoDH structure may be significantly modified due to membrane lipid alterations, impacting its activity and promoting its binding to ubiquinone [104]. It is then likely influenced by changes in temperature because temperature affects membrane structure and composition. Moreover, the association between DhoDH and the mitochondria’s inner membrane implies that its activity should be measured using a respirometry assay with the intact mitochondria, as with complex III and IV, in order to isolate the specific contribution of DhoDH to overall oxygen consumption in the OXPHOS system [101].

6. Choline Dehydrogenase

Glycine betaine (GB) can be synthetized by a two-step pathway: choline → betaine aldehyde → GB [105]. In animals, some fungi, and bacteria, the first step is catalyzed by mitochondrial membrane-bound ChoDH (E.C. 1.1.99.1) [106,107,108] or soluble choline oxidase [109,110,111,112], whereas it is catalyzed by choline monooxygenase in plants [113,114]. It has been shown in rat liver mitochondria that coenzyme Q might be the primary electron acceptor for ChoDH [105]. The second step—the transformation of betaine aldehyde into GB—seems to be catalyzed by NAD-dependent betaine aldehyde dehydrogenase in all organisms (reviewed by [115]), and this can be performed in the mitochondrial matrix [116]. Synthesis and/or transport and accumulation of GB is possible in a wide range of organisms.
GB is known to be involved in many cellular and biochemical processes, including the metabolism of carbohydrates, lipids, homocysteine/methionine, and ethanol as well as in macromolecule stabilization, antioxidant activity, and protein synthesis. It acts by changing the activity of various enzymes, either by directly regulating gene expression, by changing the phosphorylation status through specific kinases, or by controlling gene expression through changes in the degree of methylation of target gene promoters (reviewed by [115]). Considering its involvement in many metabolic pathways, GB is believed to be involved in many diseases, including diabetes, Alzheimer’s, Parkinson’s, Huntington’s disease, cardiovascular diseases, nonalcoholic steatohepatitis, and kidney diseases (reviewed by [115]). Interestingly, even though mutations in genes associated with choline transporters and choline metabolism enzymes can cause the childhood onset of severe neurologic and metabolic diseases, no diseases have been associated with ChoDH mutations in humans [117]. In fact, mice that have undergone ChoDH knockout are viable and overtly healthy, although both males and females are infertile (https://www.mousephenotype.org/ from [11], accessed on 1 April 2022). In contrast, choline is an essential nutrient in some conditions, and its deficiency is associated with liver diseases (reviewed by [11]).
Many bacteria, plants, and animals will accumulate glycine betaine (GB) under abiotic stress conditions [118,119]. In plants, betaine is synthesized and accumulated as an osmoprotectant against salt and temperature stresses (reviewed by [120]). Transgenic plants engineered with bacterial enzymes that synthesize glycine betaine showed enhanced tolerance towards various environmental stresses, including freezing, high temperatures, and hypersalinity [106]. Furthermore, tobacco cells showed an increase in thermotolerance following the overproduction of proline while exposed to high temperatures [121]. Synechococcus sp. and Arabidopsis thaliana transformed with genes from the soil bacteria Arthrobacter globiformis that encode choline oxidase (codA) have demonstrated improved tolerance to both high and freezing temperatures [122,123,124]. Enhanced tolerance to low temperatures has been also observed during germination of transgenic A. thaliana seeds transformed with codA [125,126]. The cryoprotective effect of GB might come from its ability to stabilize tertiary protein structures and prevent, or reverse, the disruption of tertiary protein structures caused by non-compatible (perturbing) solutes [127]. Increased GB levels in plants also cause a significant increase in freezing tolerance [128,129,130,131]. Plants facing thermal stress accumulate osmoprotective solutes with a high N content, such as quaternary ammonium compounds (QACs) (reviewed by [132]). The most common QACs are glycine, betaine, and choline, and they notably accumulate under both low [133] and high [134,135] temperatures. Choline is a fundamental precursor of phosphatidylcholine, which is a dominant constituent of eukaryote membrane phospholipids, and a large proportion of free choline is released into the cytoplasm during environmental stress [136,137], potentially contributing to the protective role of choline during thermal stress.
Choline dehydrogenase therefore appears to be mainly involved in providing metabolites that protect plants from stressful conditions linked to temperature changes. However, it is still unknown whether similar mechanisms occur in animals. It is known that ChoDH is associated with the modulation of mitophagy in mammalian cells [138]. This study suggested that ChoDH is required for the recruitment of SQSTM1 and LC3, which are essential for PARK2-mediated mitophagy. Knowing that mitophagy promotes the maintenance of homeostasis and is sensitive to oxidative stress and [139], and that temperature significantly affects oxidative stress in invertebrates [140,141] and fish [4], it would be worthwhile to explore the evolution of ChoDH in the context of thermal adaptation and acclimation.

7. Proline Dehydrogenase

Proline can be a fuel source in bacteria, protists, plants, and animals (reviewed by [142]). In eukaryotic cells, proline dehydrogenase (ProDH) is located in the inner mitochondrial membrane, and catalyzes the oxidation of L-proline into pyrroline-5-carboxylate. The enzyme ProDH shuttles electrons directly into the ETS Q-junction, or to oxygen [142]. In mammals, ProDH has been shown to support cell respiration under acute nutrient stress [143,144]. Pyrroline-5-carboxylate produced by ProDH can be hydrolyzed nonenzymatically to glutamic semialdehyde, which is further oxidized to glutamate during an NAD+-dependent reaction catalyzed by pyrroline-5-carboxylate dehydrogenase (P5CDH). Another route for pyrroline-5-carboxylate is conversion into ornithine, which is connected to the urea cycle [145]. Glutamate produced by P5CDH can be deaminated by alanine aminotransferase, or by glutamate dehydrogenase. This leads to α-ketoglutarate, which acts as a sparker metabolite that replenishes the citric acid cycle, feeding reducing equivalents from complexes I and II to the ETS, and maintains a higher potential for pyruvate oxidation (reviewed by [11,68]). The metabolism of proline is also closely linked with the pentose phosphate pathway in animals [145,146].
The importance of proline catabolism in cellular homeostasis is underscored by its link with many other metabolic pathways (ETS, CAC, pentose phosphate pathway, urea cycle). For example, an association between ProDH and genotoxic, inflammatory, and metabolic stress has been suggested [143]. When a cell faces variations in its internal and environmental context, programmed cell death can be induced through ProDH [147,148,149]. The induction of ProDH expression by p53 leads to increased proline oxidation, reactive oxygen species formation, and the induction of apoptosis; however, proline dehydrogenase may also be required to ensure proline protection against peroxide-induced cell death [150]. It has been suggested that proline and ProDH help to maintain NADPH, which is required to prevent and/or attenuate oxidative stress. The same authors [150] observed that ProDH activates Akt during H2O2 exposure, which inhibits FoxO3a and blocks cell death. Proline metabolism has also been associated with lifespan in nematodes [151] and yeast [152].
Proline can be a direct electron donor to the Q-junction (independent of complex I and II activity) or an indirect electron donor in the citric acid cycle. It was shown that most proline-induced mitochondrial respiration is independent of complex I in insects [153], mouse liver mitochondria, a human cell model, and cancer cells [143]. However, succinate inhibits ProDH in a noncompetitive way and prevents ROS generation by ProDH, suggesting that there is cross-talk between proline and succinate in terms of their roles in partially controlling mitochondrial respiration and their different signaling pathways associated with stress responses [143].
Sporadic data on the proline oxidation pathway seem to show that the activities and functions of ProDH are highly diverse and depend on the environmental, cellular, and phylogenetic contexts, even within a group of closely related species (reviewed by [68]). Proline can be used as a main fuel, as a co-substrate, or simply as a sparker metabolite [68,153]. An exceptional reliance on proline as an energy substrate has been observed in various species of invertebrates, such as some flying and blood-feeding insects [153,154,155,156,157,158,159]. It was shown that proline oxidation coupled with respiration is more prominent in female mosquitoes (Aedes aegypti) compared to males, and that blood-feeding is also restricted to females [73]. In at least one bivalve species (Mercenaria mercenaria; [160]), proline is the preferred substrate of the mitochondria. In some other species of insects (e.g., beetles) and in cephalopods, proline is used not as a main substrate, but as a sparker substrate, working in combination with carbohydrates and providing additional intermediates to the TCA cycle to support high activity rates (reviewed by [161]).
Considering the limited information on the utilization of proline as a mitochondrial substrate by eukaryotes, the impact of temperature on proline metabolism, as well as its importance in temperature adaptation or acclimation, are still open fields to explore. Some studies have shown that proline accumulates prior to overwintering in gall fly larva (Eurosta solidaginis) [162] and in southwestern (Diatraea grandiosella) and European (Ostrinia nubilalis) corn borers [163]. Proline accumulation depends on alterations of ProDH or proline oxidase functions. Changes in proline metabolism have been related to survival when facing thermal stress in the parasite Trypanosoma cruzi, the etiological agent of Chagas disease [164]. The involvement of proline in resistance to high temperature stress has been linked to the extreme thermal resistance—up to 70 °C—of the enzyme Δ1-pyrroline-5-carboxylate reductase, which catalyzes proline biosynthesis [165]. A conclusive metabolomics analysis of CI-deficient muscle has shown that proline oxidation, along with other substrates, such as succinate, glycerophophate, and fatty acids, can compensate for complex I deficiency by feeding electrons directly into the ETS Q-junction [166]. Modulation of rel ProDH content and activity could, therefore, be an adaptive response to compensate for the alterations in key electron transport processes (for example CI) following changes in environmental conditions, including temperature.
Proline metabolism has also been linked with tolerance to temperature changes in plants and fungi. Many plants under thermal (cold and heat) stress accumulate osmoprotective solutes with a high N content, such as choline and proline (e.g., [132,133,134,135,167,168]). Proline content has even been identified as one of the major desirable criteria for screening heat tolerance in wheat [169]. Proline acts to protect cell structure [170] and encourages growth following the stress period [170,171,172]. In creeping bentgrass (Agrostis stolonifera L.), a close association between cold stress tolerance and the proline-associated pentose phosphate pathway has been established [173]. Enhanced proline concentration has also been observed when facing high temperature stress in the fungus Aspergillus flavus [174]. More research is needed to understand the role of this pathway in tolerance to temperature variation in animals.

8. Sulfide:Quinone Oxidoreductase

Hydrogen sulfide (H2S) is an environmental toxin; at high concentrations, it can inhibit mitochondrial respiration through tight binding to cytochrome c oxidase [175,176,177] and mitochondrial depolarization [178]. To prevent this poisoning, sulfide:quinone oxidoreductase (SQOR) couples H2S oxidation to ubiquinone reduction, providing electrons to complex III (see [179] for a review). When present at lower concentrations, H2S can play a signaling role in the cardiovascular, central nervous, and gastrointestinal systems of mammals [180]. Mitochondria have been identified as a major target. For example, sulfidration of ATP synthase at low concentrations increases this enzyme’s activity in human-derived cell lines [181]. These concentration-dependent antagonistic impacts of sulfide on mitochondrial functions might require the tight regulation of its concentrations. The adjustment and control of SQOR activity could, therefore, respond not only to insure detoxification, but also to modulate sulfide concentration in the range of its signaling functions. It remains to be explored to what extent SQOR is involved in the adaptation to sulfide-rich and fluctuating environments, such as the deep sea and ocean sediments, or if this enzyme has any adaptive functions during temperature fluctuations. A recent study of the transcriptomic response of deep-sea mussels (Gigantas platifrons) to exposure to high sulfide concentrations revealed upregulation of oxidative phosphorylation and sulfide oxidation, suggesting plasticity in sulfide detoxification capacity [182]. In other studies on zebrafish (Danio rerio), it has been observed that SQOR expression is modulated by temperature [183,184,185,186]. These observations indicate a role of SQOR in adaptive responses to thermal stress, and call for further investigations.

9. Conclusions

To understand the effect of temperature on metabolism, it is first important to acquire a good picture of the preferred mitochondrial substrates and pathways specific to the species, tissues, and thermal conditions. Considering the number of substrates and pathways involved, determining the impact of temperature on mitochondrial metabolism thus becomes very complex. Determining how persistently preferred pathways are used requires respiration measurements for each substrate under various temperature conditions. The maximum catalytic capacity of mitochondria can also be informative. For this, we need to verify that the Q-junction is saturated at the mitochondria’s highest capacity by ensuring an adequate provision of electrons to the ETS. The sensitivity of mitochondrial respiratory control can also be explored. The impact of adding an uncoupler could reveal to what extent OXPHOS is controlled by the phosphorylation system (i.e., ATP synthase, adenine nucleotide translocase, phosphate carrier), which varies greatly among species [10,187,188]. A more detailed approach to examine respiratory control would be to estimate the control strength of different steps using inhibitors of specific pathways (see [74]). Most studies on sensitivity to environmental conditions have focused on respiration rates usually measured by high-resolution respirometry, but respiration rate and ATP synthesis are not the only outputs of mitochondria; the management of ROS production and buffering also become major concerns [189,190]. Finally, we should be aware of the consequences that changes in environmental conditions have on the efficiency of mitochondrial respiration, i.e., the amount of ATP synthetized per unit of oxygen consumed [191,192]. All these considerations for studying mitochondrial sensitivity are also valuable for acclimation and adaptation studies.
To further forecast the ability of ectotherms to adapt to a changing environment, we should be aware of highly sensitive traits of physiological and biochemical phenotypes that might impose limitations following change. In view of the central role of mitochondria in bioenergetics and the regulation of different pathways, it is likely that mitochondrial adaptation is a key feature determining the ability to colonize new thermal habitats. The fundamental question is which parts and pathways associated with the mitochondria can impose constraints in a changing environment. Knowledge of the weak areas will allow the exploration of their variability, heritability, and, therefore, evolvability. The first step required to identify these weak traits is to better portray the thermal sensitivity of the key pathways and steps identified in this review. This represents a major enterprise that calls for the involvement of many more comparative physiologists and biochemists than the few already active this field.
a = 1

Author Contributions

Conceptualization, H.L. and P.U.B.; writing—original draft preparation, H.L. and P.U.B.; writing—review and editing, H.L. and P.U.B.; visualization, H.L. and P.U.B.; funding acquisition, H.L. and P.U.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Discovery Grants from the Natural Sciences and Engineering Research Council of Canada, grant numbers RGPIN-2021-02924 (to H. Lemieux) and RGPIN-2019-05992 (to P.U. Blier).

Acknowledgments

The authors thank Laure Devine for editorial review of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Antão, L.H.; Bates, A.E.; Blowes, S.A.; Waldock, C.; Supp, S.R.; Magurran, A.E.; Dornelas, M.; Schipper, A.M. Temperature-related biodiversity change across temperate marine and terrestrial systems. Nat. Ecol. Evol. 2020, 4, 927–933. [Google Scholar] [CrossRef] [PubMed]
  2. Blier, P.U.; Lemieux, H.; Pichaud, N. Holding our breath in our modern world: Will mitochondria keep the pace with global changes? Can. J. Zool. 2014, 92, 591–601. [Google Scholar] [CrossRef]
  3. Chung, D.J.; Schulte, P.M. Mitochondria and the thermal limits of ectotherms. J. Exp. Biol. 2020, 223, 1–14. [Google Scholar] [CrossRef] [PubMed]
  4. Christen, F.; Desrosiers, V.; Dupont-Cyr, B.A.; Vandenberg, G.W.; Le Francois, N.R.; Tardif, J.C.; Dufresne, F.; Lamarre, S.G.; Blier, P.U. Thermal tolerance and thermal sensitivity of heart mitochondria: Mitochondrial integrity and ROS production. Free Radic. Biol. Med. 2018, 116, 11–18. [Google Scholar] [CrossRef] [PubMed]
  5. Iftikar, F.I.; Hickey, A.J. Do mitochondria limit hot fish hearts? Understanding the role of mitochondrial function with heat stress in Notolabrus celidotus. PLoS ONE 2013, 8, e64120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Pichaud, N.; Ballard, J.W.; Tanguay, R.M.; Blier, P.U. Thermal sensitivity of mitochondrial functions in permeabilized muscle fibers from two populations of Drosophila simulans with divergent mitotypes. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2011, 301, R48–R59. [Google Scholar] [CrossRef] [Green Version]
  7. Pichaud, N.; Ballard, J.W.O.; Tanguay, R.M.; Blier, P.U. Naturally occurring mitochondrial DNA haplotypes exhibit metabolic differences: Insight into functional properties of mitochondria. Evolution 2012, 66, 3189–3197. [Google Scholar] [CrossRef]
  8. Blier, P.U.; Dufresne, F.; Burton, R.S. Natural selection and the evolution of mtDNA-encoded peptides: Evidence for intergenomic co-adaptation. Trends Genet. 2001, 17, 400–406. [Google Scholar] [CrossRef]
  9. Baris, T.Z.; Wagner, D.N.; Dayan, D.I.; Du, X.; Blier, P.U.; Pichaud, N.; Oleksiak, M.F.; Crawford, D.L. Evolved genetic and phenotypic differences due to mitochondrial-nuclear interactions. PLoS Genet 2017, 13, e1006517. [Google Scholar] [CrossRef] [Green Version]
  10. Lemieux, H.; Blier, P.U.; Gnaiger, E. Remodeling pathway control of oxidative phosphorylation by temperature in the heart. Sci. Rep. 2017, 7, 2840. [Google Scholar] [CrossRef] [Green Version]
  11. Banerjee, R.; Purhonen, J.; Kallijärvi, J. The mitochondrial coenzyme Q junction and complex III: Biochemistry and pathophysiology. FEBS J. 2021, 1–23. [Google Scholar] [CrossRef] [PubMed]
  12. Longo, N.; Frigeni, M.; Pasquali, M. Carnitine transport and fatty acid oxidation. Biochim. Biophys. Acta 2016, 1863, 2422–2435. [Google Scholar] [CrossRef] [PubMed]
  13. Lemieux, H.; Tardif, J.C.; Dutil, J.D.; Blier, P.U. Thermal sensitivity of cardiac mitochondrial metabolism in an ectothermic species from a cold environment, Atlantic wolffish (Anarhichas lupus). J. Exp. Mar. Biol. Ecol. 2010, 384, 113–118. [Google Scholar] [CrossRef]
  14. Takeuchi, K.; Nakano, Y.; Kato, U.; Kaneda, M.; Aizu, M.; Awano, W.; Yonemura, S.; Kiyonaka, S.; Mori, Y.; Yamamoto, D.; et al. Changes in temperature preferences and energy homeostasis in dystroglycan mutants. Science 2009, 323, 1740–1743. [Google Scholar] [CrossRef] [PubMed]
  15. Blier, P.U.; Lemieux, H. The impact of the thermal sensitivity of cytochrome c oxidase on the respiration rate of Arctic charr red muscle mitochondria. J. Comp. Physiol. B 2001, 171, 247–253. [Google Scholar] [CrossRef] [PubMed]
  16. Lemieux, H.; Vazquez, E.J.; Fujioka, H.; Hoppel, C.L. Decrease in mitochondrial function in rat cardiac permeabilized fibers correlates with the aging phenotype. J. Gerontol. A Biol. Sci. Med. Sci. 2010, 65, 1157–1164. [Google Scholar] [CrossRef] [Green Version]
  17. Costa, L.E.; Boveris, A.; Koch, O.R.; Taquini, A.C. Liver and heart mitochondria in rats submitted to chronic hypobaric hypoxia. Am. J. Physiol.-Cell Physiol. 1988, 255, C123–C129. [Google Scholar] [CrossRef]
  18. Garait, B.; Couturier, K.; Servais, S.; Letexier, D.; Perrin, D.; Batandier, C.; Rouanet, J.L.; Sibille, B.; Rey, B.; Leverve, X.; et al. Fat intake reverses the beneficial effects of low caloric intake on skeletal muscle mitochondrial H2O2 production. Free Radic. Biol. Med. 2005, 39, 1249–1261. [Google Scholar] [CrossRef]
  19. Llesuy, S.; Evelson, P.; González-Flecha, B.; Peralta, J.; Carreras, M.C.; Poderoso, J.J.; Boveris, A. Oxidative stress in muscle and liver of rats with septic syndrome. Free Rad. Biol. Med. 1994, 16, 445–451. [Google Scholar] [CrossRef]
  20. Warren, B.E.; Lou, P.H.; Lucchinetti, E.; Zhang, L.; Clanachan, A.S.; Affolter, A.; Hersberger, M.; Zaugg, M.; Lemieux, H. Early mitochondrial dysfunction in glycolytic muscle, but not oxidative muscle, of the fructose-fed insulin-resistant rat. Am. J. Physiol. Endocrinol. Metab. 2014, 306, E658–E667. [Google Scholar] [CrossRef] [Green Version]
  21. Aragonés, J.; Schneider, M.; Van Geyte, K.; Fraisl, P.; Dresselaers, T.; Mazzone, M.; Dirkx, R.; Zacchigna, S.; Lemieux, H.; Nam Ho Jeoung, N.H.; et al. Deficiency or inhibition of oxygen sensor Phd1 induces hypoxia tolerance by reprogramming basal metabolism. Nat. Genet. 2008, 40, 170–180. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Gnaiger, E. Capacity of oxidative phosphorylation in human skeletal muscle. New perspectives of mitochondrial physiology. Int. J. Biochem. Cell Biol. 2009, 41, 1837–1845. [Google Scholar] [CrossRef] [PubMed]
  23. Gerber, L.; Clow, K.A.; Mark, F.C.; Gamperl, A.K. Improved mitochondrial function in salmon (Salmo salar) following high temperature acclimation suggests that there are cracks in the proverbial “ceiling”. Sci. Rep. 2020, 10, 21636. [Google Scholar] [CrossRef] [PubMed]
  24. Pichaud, N.; Ekström, A.; Breton, S.; Sundström, F.; Rowinski, P.; Blier, P.U.; Sandblom, E. Adjustments of cardiac mitochondrial phenotype in a warmer thermal habitat is associated with oxidative stress in European perch, Perca fluviatilis. Sci. Rep. 2020, 10, 17697. [Google Scholar] [CrossRef]
  25. Michaelsen, J.; Fago, A.; Bundgaard, A. High temperature impairs mitochondrial function in rainbow trout cardiac mitochondria. J. Exp. Biol. 2021, 224, jeb242382. [Google Scholar] [CrossRef]
  26. Iftikar, F.I.; Morash, A.J.; Cook, D.G.; Herbert, N.A.; Hickey, A.J. Temperature acclimation of mitochondria function from the hearts of a temperate wrasse (Notolabrus celidotus). Comp. Biochem. Physiol. A Mol. Integr. Physiol. 2015, 184, 46–55. [Google Scholar] [CrossRef]
  27. Dhingra, R.; Kirshenbaum, L.A. Succinate dehydrogenase/complex II activity obligatorily links mitochondrial reserve respiratory capacity to cell survival in cardiac myocytes. Cell Death Dis. 2015, 6, e1956. [Google Scholar] [CrossRef] [Green Version]
  28. Pfleger, J.; He, M.; Abdellatif, M. Mitochondrial complex II is a source of the reserve respiratory capacity that is regulated by metabolic sensors and promotes cell survival. Cell Death Dis. 2015, 6, e1835. [Google Scholar] [CrossRef] [Green Version]
  29. Treberg, J.R.; Quinlan, C.L.; Brand, M.D. Hydrogen peroxide efflux from muscle mitochondria underestimates matrix superoxide production-a correction using glutathione depletion. FEBS J. 2010, 277, 2766–2778. [Google Scholar] [CrossRef] [Green Version]
  30. St-Pierre, J.; Buckingham, J.A.; Roebuck, S.J.; Brand, M.D. Topology of superoxide production from different sites in the mitochondrial electron transport chain. J. Biol. Chem. 2002, 277, 44784–44790. [Google Scholar] [CrossRef] [Green Version]
  31. Hansford, R.G.; Hogue, B.A.; Mildaziene, V. Dependence of H2O2 formation by rat heart mitochondria on substrate availability and donor age. J. Bioenerg. Biomembr. 1997, 29, 89–95. [Google Scholar] [CrossRef] [PubMed]
  32. Lambert, A.J.; Brand, M.D. Superoxide production by NADH:ubiquinone oxidoreductase (complex I) depends on the pH gradient across the mitochondrial inner membrane. Biochem. J. 2004, 382, 511–517. [Google Scholar] [CrossRef] [PubMed]
  33. Quinlan, C.L.; Orr, A.L.; Perevoshchikova, I.V.; Treberg, J.R.; Ackrell, B.A.; Brand, M.D. Mitochondrial complex II can generate reactive oxygen species at high rates in both the forward and reverse reactions. J. Biol. Chem. 2012, 287, 27255–27264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Siebels, I.; Dröse, S. Q-site inhibitor induced ROS production of mitochondrial complex II is attenuated by TCA cycle dicarboxylates. Biochim. Biophys. Acta 2013, 1827, 1156–1164. [Google Scholar] [CrossRef] [Green Version]
  35. Kluckova, K.; Sticha, M.; Cerny, J.; Mracek, T.; Dong, L.; Drahota, Z.; Gottlieb, E.; Neuzil, J.; Rohlena, J. Ubiquinone-binding site mutagenesis reveals the role of mitochondrial complex II in cell death initiation. Cell Death Dis. 2015, 6, e1749. [Google Scholar] [CrossRef] [Green Version]
  36. Blier, P.U.; Munro, D.; Degletagne, C.; Rodriguez, E.T. What modulates animal longevity? Fast and slow aging in bivalves as a model for the study of lifespan. Semin. Cell Dev. Biol. 2017, 70, 130–140. [Google Scholar] [CrossRef]
  37. Hunter-Manseau, F.; Desrosiers, V.; Le François, N.R.; Dufresne, F.; Detrich, H.W.; Nozais, C.; Blier, P.U. From Africa to Antarctica: Exploring the Metabolism of Fish Heart Mitochondria Across a Wide Thermal Range. Front. Physiol. 2019, 10, 1–11. [Google Scholar] [CrossRef]
  38. Munro, D.; Pichaud, N.; Paquin, F.; Kemeid, V.; Blier, P.U. Low hydrogen peroxide production in mitochondria of the long-lived Arctica islandica: Underlying mechanisms for slow aging. Aging Cell 2013, 12, 584–592. [Google Scholar] [CrossRef]
  39. Barja, G. Towards a unified mechanistic theory of aging. Exp. Gerontol. 2019, 124, 110627. [Google Scholar] [CrossRef]
  40. Lu, D.L.; Ma, Q.; Wang, J.; Li, L.Y.; Han, S.L.; Limbu, S.M.; Li, D.L.; Chen, L.Q.; Zhang, M.L.; Du, Z.Y. Fasting enhances cold resistance in fish through stimulating lipid catabolism and autophagy. J. Physiol. 2019, 597, 1585–1603. [Google Scholar] [CrossRef]
  41. Rodnick, K.J.; Sidell, B.D. Cold acclimation increases carnitine palmitoyltransferase I activity in oxidative muscle of striped bass. Am. J. Physiol. 1994, 266, R405–R412. [Google Scholar] [CrossRef] [PubMed]
  42. Guderley, H.; Gawlicka, A. Qualitative modification of muscle metabolic organization with thermal acclimation of rainbow trout, Oncorhynchus mykiss. Fish Physiol. Biochem. 1992, 10, 123–132. [Google Scholar] [CrossRef] [PubMed]
  43. Kyprianou, T.D.; Portner, H.O.; Anestis, A.; Kostoglou, B.; Feidantsis, K.; Michaelidis, B. Metabolic and molecular stress responses of gilthead seam bream Sparus aurata during exposure to low ambient temperature: An analysis of mechanisms underlying the winter syndrome. J. Comp. Physiol. B 2010, 180, 1005–1018. [Google Scholar] [CrossRef] [PubMed]
  44. Crockett, E.L.; Sidell, B.D. Some pathways of energy metabolism are cold adapted in Antartic fishes. Physiol. Zool. 1990, 63, 472–488. [Google Scholar] [CrossRef]
  45. Lesser, M.P.; Kruse, V.A. Seasonal temperature compensation in the horse mussel, Modiolus modiolus: Metabolic enzymes, oxidative stress and heat shock proteins. Comp. Biochem. Physiol. A Mol. Integr. Physiol. 2004, 137, 495–504. [Google Scholar] [CrossRef]
  46. Rogers, K.D.; Thompson, M.B.; Seebacher, F. Beneficial acclimation: Sex specific thermal acclimation of metabolic capacity in the striped marsh frog (Limnodynastes peronii). J. Exp. Biol. 2007, 210, 2932–2938. [Google Scholar] [CrossRef] [Green Version]
  47. Gracey, A.Y.; Fraser, E.J.; Li, W.Z.; Fang, Y.X.; Taylor, R.R.; Rogers, J.; Brass, A.; Cossins, A.R. Coping with cold: An integrative, multitissue analysis of the transcriptome of a poikilothermic vertebrate. Proc. Nat. Acad. Sci. USA 2004, 101, 16970–16975. [Google Scholar] [CrossRef] [Green Version]
  48. Guderley, H.; Johnston, I.I. Plasticity of fish muscle mitochondria with thermal acclimation. J. Exp. Biol. 1996, 199, 1311–1317. [Google Scholar] [CrossRef]
  49. Hoskins, D.D.; Mackenzie, C.G. Solubilization and electron transfer flavoprtein requirement of mitochondrial sarcosine dehydrogenase and dimethylglycine dehydrogenase. J. Biol. Chem. 1961, 236, 177–183. [Google Scholar] [CrossRef]
  50. Ducker, G.S.; Rabinowitz, J.D. One-Carbon Metabolism in Health and Disease. Cell Metab. 2017, 25, 27–42. [Google Scholar] [CrossRef] [Green Version]
  51. Clare, C.E.; Brassington, A.H.; Kwong, W.Y.; Sinclair, K.D. One-Carbon Metabolism: Linking Nutritional Biochemistry to Epigenetic Programming of Long-Term Development. Annu. Rev. Anim. Biosci. 2019, 7, 263–287. [Google Scholar] [CrossRef] [PubMed]
  52. Abeles, R.H.; Frisell, W.R.; Mackenzie, C.G. A dual isotope effect in the enzymatic oxidation of deuteromethyl sarcosine. J. Biol. Chem. 1960, 235, 853–856. [Google Scholar] [CrossRef]
  53. Mackenzie, C.G.; Frisell, W.R. The metabolism of dimethylglycine by liver mitochondria. J. Biol. Chem. 1958, 232, 417–427. [Google Scholar] [CrossRef]
  54. Wittwer, A.J.; Wagner, C. Identification of the folate-binding proteins of rat liver mitochondria as dimethylglycine dehydrogenase and sarcosine dehydrogenase. Flavoprotein nature and enzymatic properties of the purified proteins. J. Biol. Chem. 1981, 256, 4109–4115. [Google Scholar] [CrossRef]
  55. Chung, D.J.; Sparagna, G.C.; Chicco, A.J.; Schulte, P.M. Patterns of mitochondrial membrane remodeling parallel functional adaptations to thermal stress. J. Exp. Biol 2018, 221, jeb174458. [Google Scholar] [CrossRef] [Green Version]
  56. Klingenberg, M. Localization of the glycerol-phosphate dehydrogenase in the outer phase of the mitochondrial inner membrane. Eur. J. Biochem. 1970, 13, 247–252. [Google Scholar] [CrossRef]
  57. Mráček, T.; Drahota, Z.; Houštěk, J. The function and the role of the mitochondrial glycerol-3-phosphate dehydrogenase in mammalian tissues. Biochim. Biophys. Acta 2013, 1827, 401–410. [Google Scholar] [CrossRef] [Green Version]
  58. Estabrook, R.W.; Sacktor, B. α-Glycerophosphate oxidase of flight muscle mitochondria. J. Biol. Chem. 1958, 233, 1014–1019. [Google Scholar] [CrossRef]
  59. Ghosh, R.K.; Ghosh, N.; De, S.; Ray, A.K.; Medda, A.K. Effect of l-triiodothyronine on the mitochondrial α-glycerophosphate dehydrogenase activity, mitochondrial and total protein contents of brain of Singi fish (Heteropneustes fossilis bloch). Neurochem. Int. 1983, 5, 635–640. [Google Scholar] [CrossRef]
  60. Medda, A.K.; Ghosh, R.K. Inhibitory influence of thiourea on brain of singi fish (Heteropneustes fossilis bloch) and subsequent recovery by l-triiodothyronine. Neurochem. Int. 1984, 6, 527–532. [Google Scholar] [CrossRef]
  61. Rigoulet, M.; Aguilaniu, H.; Avéret, N.; Bunoust, O.; Camougrand, N.; Grandier-Vazeille, X.; Larsson, C.; Pahlman, I.L.; Manon, S.; Gustafsson, L. Organization and regulation of the cytosolic NADH metabolism in the yeast Saccharomyces cerevisiae. Mol. Cell Biochem. 2004, 256–257, 73–81. [Google Scholar] [CrossRef] [PubMed]
  62. Hess, R.; Pearse, A.G. Histochemical and homogenization studies of mitochondrial alpha-glycerophosphate dehydrogenase in the nervous system. Nature 1961, 191, 718–719. [Google Scholar] [CrossRef] [PubMed]
  63. Lee, Y.P.; Lardy, H.A. Influence of thyroid hormones on L-alpha-glycerophosphate dehydrogenases and other dehydrogenases in various organs of the rat. J. Biol. Chem. 1965, 240, 1427–1436. [Google Scholar] [CrossRef]
  64. Salganicoff, L.; Fukami, M.H. Energy metabolism of blood platelets. I. Isolation and properties of platelet mitochondria. Arch. Biochem. Biophys. 1972, 153, 726–735. [Google Scholar] [CrossRef]
  65. Schenkman, J.B.; Richert, D.A.; Westerfeld, W.W. α-Glycerophosphate dehydrogenase activity in rat spermatozoa. Endocrinology 1965, 76, 1055–1061. [Google Scholar] [CrossRef]
  66. Swierczyński, J.; Scislowski, P.; Aleksandrowicz, Z. High activity of alpha-glycerophosphate oxidation by human placental mitochondria. Biochim. Biophys. Acta 1976, 429, 46–54. [Google Scholar] [CrossRef]
  67. Bissell, M.J.; Rambeck, W.A.; White, R.C.; Bassham, J.A. Glycerol phosphate shuttle in virus-transformed cells in culture. Science 1976, 191, 856–858. [Google Scholar] [CrossRef] [Green Version]
  68. McDonald, A.E.; Pichaud, N.; Darveau, C.A. Alternative fuels contributing to mitochondrial electron transport: Importance of non-classical pathways in the diversity of animal metabolism. Comp. Biochem, Physiol, B Biochem, Mol. Biol. 2018, 224, 185–194. [Google Scholar] [CrossRef]
  69. Yeh, J.I.; Chinte, U.; Du, S. Structure of glycerol-3-phosphate dehydrogenase, an essential monotopic membrane enzyme involved in respiration and metabolism. Proc. Natl. Acad. Sci. USA 2008, 105, 3280–3285. [Google Scholar] [CrossRef] [Green Version]
  70. Xue, L.L.; Chen, H.H.; Jiang, J.G. Implications of glycerol metabolism for lipid production. Prog. Lipid. Res. 2017, 68, 12–25. [Google Scholar] [CrossRef]
  71. Orr, A.L.; Ashok, D.; Sarantos, M.R.; Ng, R.; Shi, T.; Gerencser, A.A.; Hughes, R.E.; Brand, M.D. Novel inhibitors of mitochondrial sn-glycerol 3-phosphate dehydrogenase. PLoS ONE 2014, 9, e89938. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Brisson, D.; Vohl, M.C.; St-Pierre, J.; Hudson, T.J.; Gaudet, D. Glycerol: A neglected variable in metabolic processes? Bioessays 2001, 23, 534–542. [Google Scholar] [CrossRef] [PubMed]
  73. Soares, J.B.; Gaviraghi, A.; Oliveira, M.F. Mitochondrial physiology in the major arbovirus vector Aedes aegypti: Substrate preferences and sexual differences define respiratory capacity and superoxide production. PLoS ONE 2015, 10, e0120600. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Rodríguez, E.; Hakkou, M.; Hagen, T.M.; Lemieux, H.; Blier, P.U. Divergences in the control of mitochondrial respiration are associated with life-span variation in marine bivalves. J. Gerontol. A Biol. Med. Sci. 2021, 76, 796–804. [Google Scholar] [CrossRef]
  75. Bettinazzi, S.; Rodríguez, E.; Milani, L.; Blier, P.U.; Breton, S. Metabolic remodelling associated with mtDNA: Insights into the adaptive value of doubly uniparental inheritance of mitochondria. Proc. Biol. Sci. 2019, 286, 20182708. [Google Scholar] [CrossRef] [Green Version]
  76. Paget, C.M.; Schwartz, J.M.; Delneri, D. Environmental systems biology of cold-tolerant phenotype in Saccharomyces species adapted to grow at different temperatures. Mol. Ecol. 2014, 23, 5241–5257. [Google Scholar] [CrossRef] [Green Version]
  77. Jørgensen, L.B.; Overgaard, J.; Hunter-Manseau, F.; Pichaud, N. Dramatic changes in mitochondrial substrate use at critically high temperatures: A comparative study using Drosophila. J. Exp. Biol. 2021, 224, jeb240960. [Google Scholar] [CrossRef]
  78. Lavington, E.; Cogni, R.; Kuczynski, C.; Koury, S.; Behrman, E.L.; O’Brien, K.R.; Schmidt, P.S.; Eanes, W.F. A small system--high-resolution study of metabolic adaptation in the central metabolic pathway to temperate climates in Drosophila melanogaster. Mol. Biol. Evol. 2014, 31, 2032–2041. [Google Scholar] [CrossRef] [Green Version]
  79. Koza, R.A.; Kozak, U.C.; Brown, L.J.; Leiter, E.H.; MacDonald, M.J.; Kozak, L.P. Sequence and tissue-dependent RNA expression of mouse FAD-linked glycerol-3-phosphate dehydrogenase. Arch. Biochem. Biophys. 1996, 336, 97–104. [Google Scholar] [CrossRef]
  80. DosSantos, R.A.; Alfadda, A.; Eto, K.; Kadowaki, T.; Silva, J.E. Evidence for a compensated thermogenic defect in transgenic mice lacking the mitochondrial glycerol-3-phosphate dehydrogenase gene. Endocrinology 2003, 144, 5469–5479. [Google Scholar] [CrossRef] [Green Version]
  81. Jacobsson, A.; Stadler, U.; Glotzer, M.A.; Kozak, L.P. Mitochondrial uncoupling protein from mouse brown fat. Molecular cloning, genetic mapping, and mRNA expression. J. Biol. Chem. 1985, 260, 16250–16254. [Google Scholar] [CrossRef]
  82. Ratner, P.L.; Fisher, M.; Burkart, D.; Cook, J.R.; Kozak, L.P. The role of mRNA levels and cellular localization in controlling sn-glycerol-3-phosphate dehydrogenase expression in tissues of the mouse. J. Biol. Chem. 1981, 256, 3576–3579. [Google Scholar] [CrossRef]
  83. Panadero, J.; Pallotti, C.; Rodríguez-Vargas, S.; Randez-Gil, F.; Prieto, J.A. A downshift in temperature activates the high osmolarity glycerol (HOG) pathway, which determines freeze tolerance in Saccharomyces cerevisiae. J. Biol. Chem. 2006, 281, 4638–4645. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Bortz, W.M.; Paul, P.; Haff, A.C.; Holmes, W.L. Glycerol turnover and oxidation in man. J. Clin. Invest. 1972, 51, 1537–1546. [Google Scholar] [CrossRef] [PubMed]
  85. Owen, O.E.; Felig, P.; Morgan, A.P.; Wahren, J.; Cahill, G.F. Liver and kidney metabolism during prolonged starvation. J. Clin. Invest. 1969, 48, 574–583. [Google Scholar] [CrossRef]
  86. de la Roche, M.; Tessier, S.N.; Storey, K.B. Structural and functional properties of glycerol-3-phosphate dehydrogenase from a mammalian hibernator. Protein J. 2012, 31, 109–119. [Google Scholar] [CrossRef]
  87. Berrada, W.; Naya, A.; Ouafik, L.; Bourhim, N. Effect of hibernation, thyroid hormones and dexamethasone on cytosolic and mitochondrial glycerol-3-phosphate dehydrogenase from jerboa (Jaculus orientalis). Comp. Biochem. Physiol. B Biochem. Mol. Biol. 2000, 125, 439–449. [Google Scholar] [CrossRef]
  88. Ruberto, A.A.; Childers, C.L.; Storey, K.B. Purification and properties of glycerol-3-phosphate dehydrogenase from the liver of the hibernating ground squirrel, Urocitellus richardsonii. Comp. Biochem. Physiol. B Biochem. Mol. Biol. 2016, 202, 48–55. [Google Scholar] [CrossRef]
  89. Reis, R.A.G.; Calil, F.A.; Feliciano, P.R.; Pinheiro, M.P.; Nonato, M.C. The dihydroorotate dehydrogenases: Past and present. Arch. Biochem. Biophys. 2017, 632, 175–191. [Google Scholar] [CrossRef]
  90. Nara, T.; Hshimoto, T.; Aoki, T. Evolutionary implications of the mosaic pyrimidine-biosynthetic pathway in eukaryotes. Gene 2000, 257, 209–222. [Google Scholar] [CrossRef]
  91. Zhang, J.; Terán, G.; Popa, M.; Madapura, H.; Ladds, M.; Lianoudaki, D.; Grünler, J.; Arsenian-Henriksson, M.; McCormack, E.; Rottenberg, M.E.; et al. DHODH inhibition modulates glucose metabolism and circulating GDF15, and improves metabolic balance. iScience 2021, 24, 102494. [Google Scholar] [CrossRef] [PubMed]
  92. Rawls, J.; Knecht, W.; Diekert, K.; Lill, R.; Löffler, M. Requirements for the mitochondrial import and localization of dihydroorotate dehydrogenase. Eur. J. Biochem. 2000, 267, 2079–2087. [Google Scholar] [CrossRef] [PubMed]
  93. Khutornenko, A.A.; Roudko, V.V.; Chernyak, B.V.; Vartapetian, A.B.; Chumakov, P.M.; Evstafieva, A.G. Pyrimidine biosynthesis links mitochondrial respiration to the p53 pathway. Proc. Natl. Acad. Sci. USA 2010, 107, 12828–12833. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Miyazaki, Y.; Inaoka, D.K.; Shiba, T.; Saimoto, H.; Sakura, T.; Amalia, E.; Kido, Y.; Sakai, C.; Nakamura, M.; Moore, A.L.; et al. Selective Cytotoxicity of Dihydroorotate Dehydrogenase Inhibitors to Human Cancer Cells Under Hypoxia and Nutrient-Deprived Conditions. Front. Pharmacol. 2018, 9, 997. [Google Scholar] [CrossRef]
  95. Bader, B.; Knecht, W.; Fries, M.; Löffler, M. Expression, purification, and characterization of histidine-tagged rat and human flavoenzyme dihydroorotate dehydrogenase. Protein Expr. Purif. 1998, 13, 414–422. [Google Scholar] [CrossRef]
  96. Jones, S.W.; Penman, S.L.; French, N.S.; Park, B.K.; Chadwick, A.E. Investigating dihydroorotate dehydrogenase inhibitor mediated mitochondrial dysfunction in hepatic in vitro models. Toxicol In Vitro 2021, 72, 105096. [Google Scholar] [CrossRef]
  97. Sykes, D.B. The emergence of dihydroorotate dehydrogenase (DHODH) as a therapeutic target in acute myeloid leukemia. Expert Opin. Ther. Targets 2018, 22, 893–898. [Google Scholar] [CrossRef] [Green Version]
  98. Bajzikova, M.; Kovarova, J.; Coelho, A.R.; Boukalova, S.; Oh, S.; Rohlenova, K.; Svec, D.; Hubackova, S.; Endaya, B.; Judasova, K.; et al. Reactivation of dihydroorotate dehydrogenase-driven pyrimidine biosynthesis restores tumor growth of respiration-deficient cancer cells. Cell Metab. 2019, 29, 399–416. [Google Scholar] [CrossRef] [Green Version]
  99. Martínez-Reyes, I.; Cardona, L.R.; Kong, H.; Vasan, K.; McElroy, G.S.; Werner, M.; Kihshen, H.; Reczek, C.R.; Weinberg, S.E.; Gao, P.; et al. Mitochondrial ubiquinol oxidation is necessary for tumour growth. Nature 2020, 585, 288–292. [Google Scholar] [CrossRef]
  100. Yatsuga, S.; Fujita, Y.; Ishii, A.; Fukumoto, Y.; Arahata, H.; Kakuma, T.; Kojima, T.; Ito, M.; Tanaka, M.; Saiki, R.; et al. Growth differentiation factor 15 as a useful biomarker for mitochondrial disorders. Ann. Neurol. 2015, 78, 814–823. [Google Scholar] [CrossRef] [Green Version]
  101. Boukalova, S.; Hubackova, S.; Milosevic, M.; Ezrova, Z.; Neuzil, J.; Rohlena, J. Dihydroorotate dehydrogenase in oxidative phosphorylation and cancer. Biochim. Biophys. Acta Mol. Basis Dis. 2020, 1866, 165759. [Google Scholar] [CrossRef]
  102. Fang, J.; Uchiumi, T.; Yagi, M.; Matsumoto, S.; Amamoto, R.; Takazaki, S.; Yamaza, H.; Nonaka, K.; Kang, D. Dihydro-orotate dehydrogenase is physically associated with the respiratory complex and its loss leads to mitochondrial dysfunction. Biosci. Rep. 2013, 33, e00021. [Google Scholar] [CrossRef] [PubMed]
  103. Qiu, X.; Jiang, S.; Xiao, Y.; He, Y.; Ren, T.; Jiang, L.; Liu, R.; Chen, Q. SOX2-dependent expression of dihydroorotate dehydrogenase regulates oral squamous cell carcinoma cell proliferation. Int. J. Oral Sci. 2021, 13, 3. [Google Scholar] [CrossRef]
  104. Costeira-Paulo, J.; Gault, J.; Popova, G.; Ladds, M.; van Leeuwen, I.M.M.; Sarr, M.; Olsson, A.; Lane, D.P.; Laín, S.; Marklund, E.G.; et al. Lipids Shape the Electron Acceptor-Binding Site of the Peripheral Membrane Protein Dihydroorotate Dehydrogenase. Cell Chem Biol 2018, 25, 309–317.e4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Tsuge, H.; Nakano, Y.; Onishi, H.; Futamura, Y.; Ohashi, K. A novel purification and some properties of rat liver mitochondrial choline dehydrogenase. Biochim. Biophys. Acta 1980, 614, 274–284. [Google Scholar] [CrossRef]
  106. Gadda, G.; McAllister-Wilkins, E.E. Cloning, expression, and purification of choline dehydrogenase from the moderate halophile Halomonas elongata. Appl. Environ. Microbiol. 2003, 69, 2126–2132. [Google Scholar] [CrossRef] [Green Version]
  107. Grossman, E.B.; Hebert, S.C. Renal inner medullary choline dehydrogenase activity: Characterization and modulation. Am. J. Physiol. 1989, 256, F107–F112. [Google Scholar] [CrossRef]
  108. Miller, B.; Schmid, H.; Chen, T.J.; Schmolke, M.; Guder, W.G. Determination of choline dehydrogenase activity along the rat nephron. Biol. Chem. Hoppe Seyler 1996, 377, 129–137. [Google Scholar] [CrossRef]
  109. Boch, J.; Kempf, B.; Schmid, R.; Bremer, E. Synthesis of the osmoprotectant glycine betaine in Bacillus subtilis: Characterization of the gbsAB genes. J. Bacteriol. 1996, 178, 5121–5129. [Google Scholar] [CrossRef] [Green Version]
  110. Ikuta, S.; Imamura, S.; Misaki, H.; Horiuti, Y. Purification and characterization of choline oxidase from Arthrobacter globiformis. J. Biochem. 1977, 82, 1741–1749. [Google Scholar] [CrossRef]
  111. Lartillot, S. A simplified method of production of choline oxidase suitable for choline assay. Prep. Biochem. 1987, 17, 283–295. [Google Scholar] [PubMed]
  112. Rozwadowski, K.L.; Khachatourians, G.G.; Selvaraj, G. Choline oxidase, a catabolic enzyme in Arthrobacter pascens, facilitates adaptation to osmotic stress in Escherichia coli. J. Bacteriol. 1991, 173, 472–478. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Burnet, M.; Lafontaine, P.J.; Hanson, A.D. Assay, Purification, and Partial Characterization of Choline Monooxygenase from Spinach. Plant Physiol. 1995, 108, 581–588. [Google Scholar] [CrossRef] [PubMed]
  114. Rathinasabapathi, B.; Burnet, M.; Russell, B.L.; Gage, D.A.; Liao, P.C.; Nye, G.J.; Scott, P.; Golbeck, J.H.; Hanson, A.D. Choline monooxygenase, an unusual iron-sulfur enzyme catalyzing the first step of glycine betaine synthesis in plants: Prosthetic group characterization and cDNA cloning. Proc. Natl. Acad. Sci. USA 1997, 94, 3454–3458. [Google Scholar] [CrossRef] [Green Version]
  115. Figueroa-Soto, C.G.; Valenzuela-Soto, E.M. Glycine betaine rather than acting only as an osmolyte also plays a role as regulator in cellular metabolism. Biochimie 2018, 147, 89–97. [Google Scholar] [CrossRef]
  116. Pietruszko, R.; Chern, M. Betaine aldehyde dehydrogenase from rat liver mitochondrial matrix. Chem. Biol. Interact. 2001, 130–132, 193–199. [Google Scholar] [CrossRef]
  117. Wortmann, S.B.; Mayr, J.A. Choline-related-inherited metabolic diseases-A mini review. J. Inherit. Metab. Dis. 2019, 42, 237–242. [Google Scholar] [CrossRef] [Green Version]
  118. Rhodes, D.; Hanson, A.D. Quaternary ammonium and tertiary sulfonium compounds in higher plants. Annu. Rev. Plant Physiol. Plant Mol. Biol. 1993, 44, 357–384. [Google Scholar] [CrossRef]
  119. Kempf, B.; Bremer, E. Uptake and synthesis of compatible solutes as microbial stress responses to high-osmolality environments. Arch. Microbiol. 1998, 170, 319–330. [Google Scholar] [CrossRef]
  120. Eklund, M.; Bauer, E.; Wamatu, J.; Mosenthin, R. Potential nutritional and physiological functions of betaine in livestock. Nutr. Res. Rev. 2005, 18, 31–48. [Google Scholar] [CrossRef] [Green Version]
  121. Kuznetsov, V.V.; Shevyakova, N.I. Stress responses of tobacco cells to high temperature and salinity. Proline accumulation and phosphorylation of polypeptides. Physiol. Plant. 1997, 100, 320–326. [Google Scholar] [CrossRef]
  122. Alia; Hayashi, H.; Sakamoto, A.; Murata, N. Enhancement of the tolerance of Arabidopsis to high temperatures by genetic engineering of the synthesis of glycinebetaine. Plant J. 1998, 16, 155–161. [Google Scholar] [CrossRef] [PubMed]
  123. Hayashi, H.; Alia; Mustardy, L.; Deshnium, P.; Ida, M.; Murata, N. Transformation of Arabidopsis thaliana with the codA gene for choline oxidase; accumulation of glycinebetaine and enhanced tolerance to salt and cold stress. Plant J. 1997, 12, 133–142. [Google Scholar] [CrossRef] [PubMed]
  124. Sakamoto, A.; Valverde, R.; Alia; Chen, T.H.; Murata, N. Transformation of Arabidopsis with the codA gene for choline oxidase enhances freezing tolerance of plants. Plant J. 2000, 22, 449–453. [Google Scholar] [CrossRef]
  125. Alia; Hayashi, H.; Chen, T.H.H.; Murata, N. Transformation with a gene for choline oxidase enhances cold tolerance of Arabidopsis during germination and early growth. Plant Cell Environ. 1998, 21, 232–239. [Google Scholar] [CrossRef]
  126. Hayashi, H.; Alia; Sakamoto, A.; Nonaka, H.; Chen, T.H.H.; Murata, N. Enhanced germination under high-salt conditions of seeds of transgenic Arabidopsis with a bacterial gene (codA) for choline oxidase. J. Plant Res. 1998, 111, 357–362. [Google Scholar] [CrossRef]
  127. Bateman, J.B.; Evans, G.F.; Brown, P.R.; Gabriel, C.; Grant, E.H. Dielectric properties of the system bovine albumin: Urea: Betaine in aqueous solution. Phys. Med. Biol. 1992, 37, 175–182. [Google Scholar] [CrossRef]
  128. Xing, W.; Rajashekar, C.B. Glycine betaine involvement in freezing tolerance and water stress in Arabidopsis thaliana. Environ. Exp. Bot. 2001, 46, 21–28. [Google Scholar] [CrossRef]
  129. Allard, F.; Houde, M.; Krol, M.; Ivanov, A.; Huner, N.P.A.; Sarhan, F. Betaine improves freezing tolerance in wheat. Plant Cell Physiol. 1998, 39, 1194–1202. [Google Scholar] [CrossRef]
  130. Reaney, M.J.; Gusta, L.V. Factors Influencing the Induction of Freezing Tolerance by Abscisic Acid in Cell Suspension Cultures of Bromus inermis Leyss and Medicago sativa L. Plant Physiol. 1987, 83, 423–427. [Google Scholar] [CrossRef] [Green Version]
  131. Mantyla, E.; Lang, V.; Palva, E.T. Role of abscisic acid in drought-induced freezing tolerance, cold acclimation, and accumulation of LT178 and RAB18 proteins in Arabidopsis thaliana. Plant Physiol. 1995, 107, 141–148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Rivero, R.M.; Ruiz, J.M.; Romero, L.M. Importance of N source on heat stress tolerance due to the accumulation of proline and quaternary ammonium compounds in tomato plants. Plant Biol. 2004, 6, 702–707. [Google Scholar] [CrossRef] [PubMed]
  133. Naidu, B.P.; Paleg, L.G.; Aspinal, D.; Jennings, A.C.; Jones, G.P. Amino acid and glycine betaine accumulation in cold-stressed wheat seedlings. Phytochemistry 1991, 30, 407–409. [Google Scholar] [CrossRef]
  134. Che, T.M.; Aspinall, D.; Paleg, L.G. Stress metabolism. IV. Temperature stress and the accumulation of proline in barley and radish. J. Plant Physiol. 1974, 1, 87–97. [Google Scholar]
  135. Kuo, C.G.; Chen, H.M.; Ma, L.H. Effect of high temperature on proline content in tomato floral buds and leaves. J. Amer. Soc. Hort. Sci. 1986, 111, 746–750. [Google Scholar]
  136. Rathinasabapathi, B.; Sigua, C.; Ho, J.; Gage, D.A. Osmoprotectant beta-alanine betaine synthesis in the Plumbaginaceae: S-adenosyl-L-methionine dependent N-methylation of beta-alanine to its betaine is via N-methyl and N,N-dimethyl beta-alanines. Physiol. Plant. 2000, 109, 225–231. [Google Scholar] [CrossRef]
  137. Rathinasabapathi, B.; Gage, D.; Mackill, D.; Hanson, A. Cultivated and wild rices do not accumulate glycine-betaine due to deficiencies in two biosynthetic steps. Crop. Sci. 1993, 33, 534–538. [Google Scholar] [CrossRef]
  138. Park, S.; Choi, S.G.; Yoo, S.M.; Son, J.H.; Jung, Y.K. Choline dehydrogenase interacts with SQSTM1/p62 to recruit LC3 and stimulate mitophagy. Autophagy 2014, 10, 1906–1920. [Google Scholar] [CrossRef] [Green Version]
  139. Garza-Lombó, C.; Pappa, A.; Panayiotidis, M.I.; Franco, R. Redox homeostasis, oxidative stress and mitophagy. Mitochondrion 2020, 51, 105–117. [Google Scholar] [CrossRef]
  140. González Durán, E.; Cuaya, M.P.; Gutiérrez, M.V.; León, J.A. Effects of temperature and pH on the oxidative stress of benthic marine invertebrates. Biol. Bull. 2018, 45, 610–616. [Google Scholar] [CrossRef]
  141. Paital, B.; Chainy, G.B. Effects of temperature on complexes I and II mediated respiration, ROS generation and oxidative stress status in isolated gill mitochondria of the mud crab Scylla serrata. J. Therm. Biol. 2014, 41, 104–111. [Google Scholar] [CrossRef] [PubMed]
  142. Servet, C.; Ghelis, T.; Richard, L.; Zilberstein, A.; Savoure, A. Proline dehydrogenase: A key enzyme in controlling cellular homeostasis. Front. Biosci. Landmark Ed. 2012, 17, 607–620. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Hancock, C.N.; Liu, W.; Alvord, W.G.; Phang, J.M. Co-regulation of mitochondrial respiration by proline dehydrogenase/oxidase and succinate. Amino Acids 2016, 48, 859–872. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Pandhare, J.; Donald, S.P.; Cooper, S.K.; Phang, J.M. Regulation and function of proline oxidase under nutrient stress. J. Cell Biochem. 2009, 107, 759–768. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Hagedorn, C.H.; Phang, J.M. Transfer of reducing equivalents into mitochondria by the interconversions of proline and delta 1-pyrroline-5-carboxylate. Arch. Biochem. Biophys. 1983, 225, 95–101. [Google Scholar] [CrossRef]
  146. Phang, J.M. The regulatory functions of proline and pyrroline-5-carboxylic acid. Curr. Top. Cell Regul. 1985, 25, 91–132. [Google Scholar]
  147. Donald, S.P.; Sun, X.Y.; Hu, C.A.; Yu, J.; Mei, J.M.; Valle, D.; Phang, J.M. Proline oxidase, encoded by p53-induced gene-6, catalyzes the generation of proline-dependent reactive oxygen species. Cancer Res. 2001, 61, 1810–1815. [Google Scholar]
  148. Hu, C.A.; Donald, S.P.; Yu, J.; Lin, W.W.; Liu, Z.; Steel, G.; Obie, C.; Valle, D.; Phang, J.M. Overexpression of proline oxidase induces proline-dependent and mitochondria-mediated apoptosis. Mol. Cell Biochem. 2007, 295, 85–92. [Google Scholar] [CrossRef]
  149. Liu, Y.; Borchert, G.L.; Donald, S.P.; Surazynski, A.; Hu, C.A.; Weydert, C.J.; Oberley, L.W.; Phang, J.M. MnSOD inhibits proline oxidase-induced apoptosis in colorectal cancer cells. Carcinogenesis 2005, 26, 1335–1342. [Google Scholar] [CrossRef]
  150. Natarajan, S.K.; Zhu, W.; Liang, X.; Zhang, L.; Demers, A.J.; Zimmerman, M.C.; Simpson, M.A.; Becker, D.F. Proline dehydrogenase is essential for proline protection against hydrogen peroxide-induced cell death. Free Radic. Biol. Med. 2012, 53, 1181–1191. [Google Scholar] [CrossRef] [Green Version]
  151. Zarse, K.; Schmeisser, S.; Groth, M.; Priebe, S.; Beuster, G.; Kuhlow, D.; Guthke, R.; Platzer, M.; Kahn, C.R.; Ristow, M. Impaired insulin/IGF1 signaling extends life span by promoting mitochondrial L-proline catabolism to induce a transient ROS signal. Cell Metab. 2012, 15, 451–465. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Nishimura, A.; Yoshikawa, Y.; Ichikawa, K.; Takemoto, T.; Tanahashi, R.; Takagi, H. Longevity regulation by proline oxidation in yeast. Microorganisms 2021, 9, 1650. [Google Scholar] [CrossRef] [PubMed]
  153. Teulier, L.; Weber, J.M.; Crevier, J.; Darveau, C.A. Proline as a fuel for insect flight: Enhancing carbohydrate oxidation in hymenopterans. Proc. Biol. Sci. 2016, 283, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Olembo, N.K.; Pearson, D.J. Changes in the contents of intermediates of proline and carbohydrate metabolism in flight muscle of the tsetse fly Glossina morsitans and the fleshfly Sarcophaga tibialis. Insect Biochem. 1982, 12, 657–662. [Google Scholar] [CrossRef]
  155. Giulivi, C.; Ross-Inta, C.; Horton, A.A.; Luckhart, S. Metabolic pathways in Anopheles stephensi mitochondria. Biochem. J. 2008, 415, 309–316. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Goldstrohm, D.A.; Pennington, J.E.; Wells, M.A. The role of hemolymph proline as a nitrogen sink during blood meal digestion by the mosquito Aedes aegypti. J. Insect Physiol. 2003, 49, 115–121. [Google Scholar] [CrossRef]
  157. Scaraffia, P.Y.; Wells, M.A. Proline can be utilized as an energy substrate during flight of Aedes aegypti females. J. Insect Physiol. 2003, 49, 591–601. [Google Scholar] [CrossRef]
  158. Bursell, E. Substrates of oxidative metabolism in dipteran flight muscle. Comp. Biochem. Physiol. B 1975, 52, 235–238. [Google Scholar] [CrossRef]
  159. Bursell, E. Aspects of the metabolism of amino acids in the tsetse fly, Glossina (Diptera). J. Insect Physiol. 1963, 9, 439–452. [Google Scholar] [CrossRef]
  160. Ballantyne, J.S.; Storey, K.B. Mitochondria from the ventricle of the marine clam, Mercenaria mercenaria: Substrate preferences and effects of pH and salt concentration on proline oxidation. Comp. Biochem. Physiol. B Comp. Biochem. 1983, 76, 133–138. [Google Scholar] [CrossRef]
  161. Gäde, G.; Auerswald, L. Beetles’ choice–proline for energy output: Control by AKHs. Comp. Biochem. Physiol. B Biochem. Mol. Biol. 2002, 132, 117–129. [Google Scholar] [CrossRef]
  162. Storey, K.B.; Baust, J.G.; Storey, J.M. Intermediary metabolism during low temperature acclimation in the overwintering gall fly larva, Eurosta solidaginis. J. Comp. Physiol. B 1981, 144, 183–190. [Google Scholar] [CrossRef]
  163. Morgan, T.D.; Chippendale, G.M. Free amino acids of the haemolymph of the southwestern corn borer and the European corn borer in relation to their diapause. J. Insect Physiol. 1983, 29, 735–740. [Google Scholar] [CrossRef]
  164. Marchese, L.; Olavarria, K.; Mantilla, B.S.; Avila, C.C.; Souza, R.O.O.; Damasceno, F.S.; Elias, M.C.; Silber, A.M. Trypanosoma cruzi synthesizes proline via a Δ1-pyrroline-5-carboxylate reductase whose activity is fine-tuned by NADPH cytosolic pools. Biochem. J. 2020, 477, 1827–1845. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Magdaleno, A.; Ahn, I.Y.; Paes, L.S.; Silber, A.M. Actions of a proline analogue, L-thiazolidine-4-carboxylic acid (T4C), on Trypanosoma cruzi. PLoS ONE 2009, 4, e4534. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Terburgh, K.; Lindeque, Z.; Mason, S.; van der Westhuizen, F.; Louw, R. Metabolomics of Ndufs4(−/−) skeletal muscle: Adaptive mechanisms converge at the ubiquinone-cycle. Biochim. Biophys. Acta Mol. Basis Dis. 2019, 1865, 98–106. [Google Scholar] [CrossRef]
  167. Dionisio-Sese, M.L.; Shono, M.; Tobita, S. Effects of proline and betaine on heat inactivation of ribulose-1,5-bisphosphate carboxylase/oxygenase in crude extracts of rice seedlings. Photosynthetica 1999, 36, 557–563. [Google Scholar] [CrossRef]
  168. Chang, Y.C.; Lee, T.M. High temperature-induced free proline accumulation in Gracilaria tenuistipitata (Rhodophyta). Bot. Bull. Acad. Sin. 1999, 40, 289–294. [Google Scholar]
  169. Khan, A.; Ahmad, M.; Ahmed, M.; Iftikhar Hussain, M. Rising atmospheric temperature impact on wheat and thermotolerance strategies. Plants 2020, 10, 43. [Google Scholar] [CrossRef]
  170. Schobert, B.; Tschesche, H. Unusual solution properties of proline and its interaction with proteins. Biochim. Biophys. Acta 1978, 541, 270–277. [Google Scholar] [CrossRef]
  171. Paleg, L.G.; Douglas, T.J.; Van Daal, A.; Keech, D.B. Proline, betaine and other organic solutes protect enzymes against heat inactivation. Austr. J. Plant Physiol. 1981, 8, 107–114. [Google Scholar] [CrossRef]
  172. Charest, C.; Phan, C.T. Cold acclimation of wheat: Properties of enzymes involved in proline metabolism. Physiol. Plant. 1990, 80, 159–168. [Google Scholar] [CrossRef]
  173. Sarkar, D.; Bhowmik, P.C.; Kwon, Y.I.; Shetty, K. Clonal response to cold tolerance in creeping bentgrass and role of proline-associated pentose phosphate pathway. Bioresour. Technol. 2009, 100, 5332–5339. [Google Scholar] [CrossRef] [PubMed]
  174. Ismail; Hamayun, M.; Hussain, A.; Afzal Khan, S.; Iqbal, A.; Lee, I.J. Aspergillus flavus promoted the growth of soybean and sunflower seedlings at elevated temperature. BioMed Res. Int. 2019, 2019, 1–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Bouillaud, F.; Blachier, F. Mitochondria and sulfide: A very old story of poisoning, feeding, and signaling? Antioxid. Redox. Signal. 2011, 15, 379–391. [Google Scholar] [CrossRef]
  176. Evans, C.L. The toxicity of hydrogen sulphide and other sulphides. Q J. Exp. Physiol. Cogn. Med. Sci. 1967, 52, 231–248. [Google Scholar] [CrossRef]
  177. Nicholls, P. The effect of sulphide on cytochrome aa3. Isosteric and allosteric shifts of the reduced alpha-peak. Biochim. Biophys. Acta 1975, 396, 24–35. [Google Scholar] [CrossRef]
  178. Julian, D.; April, K.L.; Patel, S.; Stein, J.R.; Wohlgemuth, S.E. Mitochondrial depolarization following hydrogen sulfide exposure in erythrocytes from a sulfide-tolerant marine invertebrate. J. Exp. Biol. 2005, 208, 4109–4122. [Google Scholar] [CrossRef] [Green Version]
  179. Landry, A.P.; Ballou, D.P.; Banerjee, R. Hydrogen Sulfide Oxidation by Sulfide Quinone Oxidoreductase. Chembiochem 2021, 22, 949–960. [Google Scholar] [CrossRef]
  180. Kimura, H. Hydrogen sulfide: From brain to gut. Antioxid. Redox. Signal. 2010, 12, 1111–1123. [Google Scholar] [CrossRef]
  181. Módis, K.; Ju, Y.; Ahmad, A.; Untereiner, A.A.; Altaany, Z.; Wu, L.; Szabo, C.; Wang, R. S-sulfhydration of ATP synthase by hydrogen sulfide stimulates mitochondrial bioenergetics. Pharmacol. Res. 2016, 113, 116–124. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Sun, Y.; Wang, M.; Zhong, Z.; Chen, H.; Wang, H.; Zhou, L.; Cao, L.; Fu, L.; Zhang, H.; Lian, C.; et al. Adaption to hydrogen sulfide-rich environments: Strategies for active detoxification in deep-sea symbiotic mussels, Gigantidas platifrons. Sci. Total Environ. 2022, 804, 150054. [Google Scholar] [CrossRef]
  183. Icoglu Aksakal, F.; Ciltas, A. The impact of ultraviolet B (UV-B) radiation in combination with different temperatures in the early life stage of zebrafish (Danio rerio). Photochem. Photobiol. Sci. 2018, 17, 35–41. [Google Scholar] [CrossRef] [PubMed]
  184. Long, Y.; Li, L.; Li, Q.; He, X.; Cui, Z. Transcriptomic characterization of temperature stress responses in larval zebrafish. PLoS ONE 2012, 7, e37209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Long, Y.; Yan, J.; Song, G.; Li, X.; Li, X.; Li, Q.; Cui, Z. Transcriptional events co-regulated by hypoxia and cold stresses in zebrafish larvae. BMC Genom. 2015, 16, 385. [Google Scholar] [CrossRef] [Green Version]
  186. Feugere, L.; Scott, V.F.; Rodriguez-Barucg, Q.; Beltran-Alvarez, P.; Wollenberg Valero, K.C. Thermal stress induces a positive phenotypic and molecular feedback loop in zebrafish embryos. J. Therm. Biol. 2021, 102, 103114. [Google Scholar] [CrossRef]
  187. Lemieux, H.; Semsroth, S.; Antretter, H.; Höfer, D.; Gnaiger, E. Mitochondrial respiratory control and early defects of oxidative phosphorylation in the failing human heart. Int. J. Biochem. Cell Biol. 2011, 43, 1729–1738. [Google Scholar] [CrossRef]
  188. Lemieux, H.; Warren, B. An animal model to study human muscular diseases involving mitochondrial oxidative phosphorylation. J. Bioenerg. Biomembr. 2012, 44, 503–512. [Google Scholar] [CrossRef]
  189. Wiens, L.; Banh, S.; Sotiri, E.; Jastroch, M.; Block, B.A.; Brand, M.D.; Treberg, J.R. Comparison of mitochondrial reactive oxygen species production of ectothermic and endothermic fish muscle. Front. Physiol. 2017, 8, 704. [Google Scholar] [CrossRef] [Green Version]
  190. Munro, D.; Treberg, J.R. A radical shift in perspective: Mitochondria as regulators of reactive oxygen species. J. Exp. Biol. 2017, 220, 1170–1180. [Google Scholar] [CrossRef] [Green Version]
  191. Thoral, E.; Roussel, D.; Chinopoulos, C.; Teulier, L.; Salin, K. Low oxygen levels can help to prevent the detrimental effect of acute warming on mitochondrial efficiency in fish. Biol. Lett. 2021, 17, 20200759. [Google Scholar] [CrossRef] [PubMed]
  192. Salin, K.; Villasevil, E.M.; Anderson, G.J.; Selman, C.; Chinopoulos, C.; Metcalfe, N.B. The RCR and ATP/O indices can give contradictory messages about mitochondrial efficiency. Integr. Comp. Biol. 2018, 58, 486–494. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Electron entrance in oxidative phosphorylation (OXPHOS). The main entry points converging at the Q-junction (ubiquinol/ubiquinone) of the OXPHOS system are illustrated here and described as numbers 1 to 8. (1) The NADH pathway provides electrons to complex I of the electron transport system. These electrons can be provided—among others—by pyruvate (through pyruvate dehydrogenase), and glutamate (through glutamate dehydrogenase). (2) The succinate pathway uses succinate dehydrogenase to reduce FAD and feed complex II. (3) The electron-transferring flavoprotein (ETF) pathway (in light green) receives electrons from fatty acyl-CoA of various chain lengths. The carnitine shuttle system is composed of carnitine palmitoyltransferase 1 (CPT1), carnitine–acylcarnitine translocase (CACT), and carnitine palmitoyltransferase 2 (CPT2). Oxidation of the long-chain fatty acids (LCFAs) depends on the carnitine shuttle system, whereas the oxidation of short- (SCFAs) and medium- (MCFAs) chain fatty acids are largely independent of that system (reviewed by [12]). Once in the mitochondrial matrix, the fatty acids enter the process of long-chain, medium-chain, or short-chain fatty acid oxidation. In this process, NAD and FAD are reduced to NADH and FADH2; FADH2 is used as a substrate for electron-transferring flavoprotein (ETF), while NADH feeds complex I. Both NADH and FADH2 provide electrons into the Q-junction. The dashed line links LCFA, MCFA, and SCFA oxidation to the NADH pathway, which is not rate-limiting for the ETF pathway. Malate is also provided, with all fatty acid substrates, as a metabolite in the citric acid cycle to prevent the accumulation of acetyl-CoA and the concurrent inhibition of fatty acid oxidation. Dimethylglycine dehydrogenase (DMGDH) and sarcosine dehydrogenase (SARDH) can also feed electrons through the ETF pathway. (4) Glycerophosphate oxidation (in pink) occurs via mitochondrial glycerophosphate dehydrogenase (mGpDH) and shuttles electrons through FADH2 directly to the Q-junction. (5) Choline oxidation (in purple) occurs through choline dehydrogenase (ChoDH), which reduces FAD. (6) Proline dehydrogenase (ProDH, in blue) catalyzes proline oxidation and FAD reduction. (7) The dihydroorotate (Dho, in orange) is transformed into orotate by dihydroorotate dehydrogenase (DhoDH), releasing an electron from flavin mononucleotide directly into the Q-junction. (8) sulfide:quinone oxidoreductase (SQOR, in grey) oxidizes H2S and transfers an electron into the Q-junction. After convergence of electrons from diverse pathways at the Q-junction, they follow a linear segment through complexes III and IV before reducing molecular oxygen. Electron transfer is coupled with proton pumping into intermembrane space by complexes I, III, and IV, generating a proton motive force, and providing the energy required to phosphorylate ADP into ATP via ATP synthase (part of the phosphorylation system, also supported by the adenine nucleotide translocase, ANT, and the phosphate carrier, PiC). All components located in, or peripherally associated with, the mitochondrial membranes are in red; solid lines represent those associated with the inner membrane and the dashed line represents the one component associated with the outer mitochondrial membrane (i.e., CPT1).
Figure 1. Electron entrance in oxidative phosphorylation (OXPHOS). The main entry points converging at the Q-junction (ubiquinol/ubiquinone) of the OXPHOS system are illustrated here and described as numbers 1 to 8. (1) The NADH pathway provides electrons to complex I of the electron transport system. These electrons can be provided—among others—by pyruvate (through pyruvate dehydrogenase), and glutamate (through glutamate dehydrogenase). (2) The succinate pathway uses succinate dehydrogenase to reduce FAD and feed complex II. (3) The electron-transferring flavoprotein (ETF) pathway (in light green) receives electrons from fatty acyl-CoA of various chain lengths. The carnitine shuttle system is composed of carnitine palmitoyltransferase 1 (CPT1), carnitine–acylcarnitine translocase (CACT), and carnitine palmitoyltransferase 2 (CPT2). Oxidation of the long-chain fatty acids (LCFAs) depends on the carnitine shuttle system, whereas the oxidation of short- (SCFAs) and medium- (MCFAs) chain fatty acids are largely independent of that system (reviewed by [12]). Once in the mitochondrial matrix, the fatty acids enter the process of long-chain, medium-chain, or short-chain fatty acid oxidation. In this process, NAD and FAD are reduced to NADH and FADH2; FADH2 is used as a substrate for electron-transferring flavoprotein (ETF), while NADH feeds complex I. Both NADH and FADH2 provide electrons into the Q-junction. The dashed line links LCFA, MCFA, and SCFA oxidation to the NADH pathway, which is not rate-limiting for the ETF pathway. Malate is also provided, with all fatty acid substrates, as a metabolite in the citric acid cycle to prevent the accumulation of acetyl-CoA and the concurrent inhibition of fatty acid oxidation. Dimethylglycine dehydrogenase (DMGDH) and sarcosine dehydrogenase (SARDH) can also feed electrons through the ETF pathway. (4) Glycerophosphate oxidation (in pink) occurs via mitochondrial glycerophosphate dehydrogenase (mGpDH) and shuttles electrons through FADH2 directly to the Q-junction. (5) Choline oxidation (in purple) occurs through choline dehydrogenase (ChoDH), which reduces FAD. (6) Proline dehydrogenase (ProDH, in blue) catalyzes proline oxidation and FAD reduction. (7) The dihydroorotate (Dho, in orange) is transformed into orotate by dihydroorotate dehydrogenase (DhoDH), releasing an electron from flavin mononucleotide directly into the Q-junction. (8) sulfide:quinone oxidoreductase (SQOR, in grey) oxidizes H2S and transfers an electron into the Q-junction. After convergence of electrons from diverse pathways at the Q-junction, they follow a linear segment through complexes III and IV before reducing molecular oxygen. Electron transfer is coupled with proton pumping into intermembrane space by complexes I, III, and IV, generating a proton motive force, and providing the energy required to phosphorylate ADP into ATP via ATP synthase (part of the phosphorylation system, also supported by the adenine nucleotide translocase, ANT, and the phosphate carrier, PiC). All components located in, or peripherally associated with, the mitochondrial membranes are in red; solid lines represent those associated with the inner membrane and the dashed line represents the one component associated with the outer mitochondrial membrane (i.e., CPT1).
Metabolites 12 00360 g001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lemieux, H.; Blier, P.U. Exploring Thermal Sensitivities and Adaptations of Oxidative Phosphorylation Pathways. Metabolites 2022, 12, 360. https://doi.org/10.3390/metabo12040360

AMA Style

Lemieux H, Blier PU. Exploring Thermal Sensitivities and Adaptations of Oxidative Phosphorylation Pathways. Metabolites. 2022; 12(4):360. https://doi.org/10.3390/metabo12040360

Chicago/Turabian Style

Lemieux, Hélène, and Pierre U. Blier. 2022. "Exploring Thermal Sensitivities and Adaptations of Oxidative Phosphorylation Pathways" Metabolites 12, no. 4: 360. https://doi.org/10.3390/metabo12040360

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop