Next Article in Journal
The Role of Atypical Chemokine Receptors in Neuroinflammation and Neurodegenerative Disorders
Next Article in Special Issue
Reduced Levels of Lacrimal Glial Cell Line-Derived Neurotrophic Factor (GDNF) in Patients with Focal Epilepsy and Focal Epilepsy with Comorbid Depression: A Biomarker Candidate
Previous Article in Journal
Plasmonic and Photothermal Effects of CuS Nanoparticles Biosynthesized from Acid Mine Drainage with Potential Drug Delivery Applications
Previous Article in Special Issue
Astrocytes: Lessons Learned from the Cuprizone Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Noncoupled Mitochondrial Respiration as Therapeutic Approach for the Treatment of Metabolic Diseases: Focus on Transgenic Animal Models

by
Artem P. Gureev
1,2,
Alina A. Alimova
1,2,
Denis N. Silachev
3,* and
Egor Y. Plotnikov
3
1
Department of Genetics, Cytology and Bioengineering, Voronezh State University, 394018 Voronezh, Russia
2
Laboratory of Metagenomics and Food Biotechnology, Voronezh State University of Engineering Technologies, 394036 Voronezh, Russia
3
A.N. Belozersky Institute of Physico-Chemical Biology, Lomonosov Moscow State University, 119234 Moscow, Russia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(22), 16491; https://doi.org/10.3390/ijms242216491
Submission received: 10 October 2023 / Revised: 9 November 2023 / Accepted: 10 November 2023 / Published: 18 November 2023
(This article belongs to the Collection Feature Papers in Molecular Neurobiology)

Abstract

:
Mitochondrial dysfunction contributes to numerous chronic diseases, and mitochondria are targets for various toxins and xenobiotics. Therefore, the development of drugs or therapeutic strategies targeting mitochondria is an important task in modern medicine. It is well known that the primary, although not the sole, function of mitochondria is ATP generation, which is achieved by coupled respiration. However, a high membrane potential can lead to uncontrolled reactive oxygen species (ROS) production and associated dysfunction. For over 50 years, scientists have been studying various synthetic uncouplers, and for more than 30 years, uncoupling proteins that are responsible for uncoupled respiration in mitochondria. Additionally, the proteins of the mitochondrial alternative respiratory pathway exist in plant mitochondria, allowing noncoupled respiration, in which electron flow is not associated with membrane potential formation. Over the past two decades, advances in genetic engineering have facilitated the creation of various cellular and animal models that simulate the effects of uncoupled and noncoupled respiration in different tissues under various disease conditions. In this review, we summarize and discuss the findings obtained from these transgenic models. We focus on the advantages and limitations of transgenic organisms, the observed physiological and biochemical changes, and the therapeutic potential of uncoupled and noncoupled respiration.

1. Introduction

Mitochondrial respiration can be divided into three types according to its ability to form a membrane potential (∆μH+). Energy-coupled respiration allows the generation of ∆μH+ and it is utilized for ATP synthesis (Figure 1A). During uncoupled respiration, ∆μH+ is formed but immediately dissipated without ATP synthesis by specific proteins or in the presence of certain substances called uncouplers (Figure 1B). During noncoupled respiration, electron flow is not associated with ∆μH+ formation (Figure 1C) [1]. Energy-coupled respiration is facilitated by protein complexes that form the respiratory chain. The energy carried by electrons flowing through this electron transport chain is used to transport protons across the inner mitochondrial membrane. This generates potential energy in the form of an electrochemical gradient on the inner mitochondrial membrane. Subsequently, using this potential, F0F1-ATP synthase generates ATP from ADP and inorganic phosphate (Figure 1A) [2]. While oxidative phosphorylation is an extremely efficient process, a significant portion of the electron energy can be diverted towards the synthesis of free radicals, which generally have a negative impact on the structural integrity of membranes, proteins, and DNA [3]. For this reason, uncoupled and noncoupled respiration plays an important role in maintaining cellular redox homeostasis.
The physiological role of uncoupled respiration is multifaceted. Uncoupled respiration, primarily mediated by proteins from the UCP family, has been extensively studied. UCP1, found in brown adipose tissue, plays a crucial role in non-shivering adaptive thermogenesis. Through proton leak across the mitochondrial inner membrane, UCP1 dissipates the ∆μH+, leading to increased substrate oxidation and heat production independently of ATP synthesis [4]. The activation of UCP1 is facilitated by long-chain fatty acids (LCFAs). Initially, the LCFA anion binds to UCP1 on the cytosolic side of the membrane. Subsequently, H+ binds to the LCFA, triggering a conformational change that releases H+ on the opposite side of the inner mitochondrial membrane. The LCFA anion remains associated with UCP1 through hydrophobic interactions established by its carbon tail. The LCFA anion then returns for another cycle of H+ translocation (Figure 1B) [5]. Another mechanism of uncoupling oxidative phosphorylation involves specific small molecules that participate in proton translocation from the intermembrane space to the mitochondrial matrix. Protonophores are classic uncouplers that can directly transport protons across the inner mitochondrial membrane through their redox properties and lipophilic structure (Figure 1B) [6]. Non-protonophore uncouplers, on the other hand, typically act as agonists of other proteins which are involved in the regulation of ∆μH+ [7].
Noncoupled respiration is energetically similar to uncoupled respiration but has a different origin. Noncoupled respiration is facilitated by special respiratory complexes that participate in electron transport without forming ∆μH+. Classic examples of noncoupled respiration are found in alternative respiratory pathways in plant mitochondria [8]. Plant mitochondria contain at least five additional components in the electron transport chain. Four of these components catalyze the transfer of NADH or NADPH to ubiquinone, while the fifth component is an alternative oxidase (AOX) that directly catalyzes the transfer of electrons to molecular oxygen [9,10]. Alternative respiratory pathways do not involve the transport of protons across the inner mitochondrial membrane and, therefore, are not coupled with ATP synthesis (Figure 1C). With a few exceptions, animals, due to their active lifestyles, generally lack these alternative respiratory pathways [11]. Hence, animal mitochondria are highly susceptible to various poisons and xenobiotics that can inhibit electron flow and cause electron “leaks”, leading to the excessive production of reactive oxygen species (ROS) [12]. Mutations in mitochondrial DNA and nuclear DNA genes associated with the respiratory chain can also result in severe metabolic defects and pathologies [13].
Thus, the main, but not the only, positive effect of uncoupled and noncoupled respiration is the normalization of the redox state to reduce the production of free radicals, which is expected to significantly reduce the progression of various diseases [14]. In general, chemical uncouplers show promise as pharmacological agents for treating various metabolic disorders. However, it should be acknowledged that in recent years, there has been a waning of interest in chemical uncouplers, despite the growing interest in the investigation of mitochondrial pathologies (Figure 2). This is likely due to the fact that many classical and well-studied uncouplers have a very narrow therapeutic window. Perhaps this is because some uncouplers have been “discredited” due to a high number of side effects [15,16]. Therefore, while this review briefly addresses the topic of chemical uncouplers, its primary focus is on the comprehensive discussion of transgenic animal models pertaining to uncoupled and noncoupled respiration. Undoubtedly, genome editing as a therapeutic approach for human diseases is not yet considered a viable method. However, transgenic animal models allow for a better assessment of the effects of uncoupled and noncoupled respiration on cellular, tissue, and organismal states. These effects can be evaluated at different stages of organism development (from embryonic to aging stages) and under conditions of various induced diseases. A deeper understanding of these processes will help in developing new therapeutic approaches for treating metabolic disorders by regulating the degree of coupling in mitochondrial respiration.

2. Chemical Uncouplers

Mitochondrial uncouplers are synthetic compounds that belong to various classes of chemicals. These uncouplers exhibit multiple mechanisms of action, enabling them to be classified indirectly only. Based on their functionality, uncouplers can be divided into those that exhibit a protonophore effect, while others indirectly induce uncoupling by regulating the activity of uncoupling proteins or altering mitochondrial function [17]. Protonophores are typically hydrophobic aromatic compounds with a negative charge. These compounds have the ability to distribute negative charge among multiple atoms through π-orbitals, thus facilitating the delocalization of a proton’s charge upon its attachment to the molecule [18]. It allows them to easily penetrate through the lipid membrane and move in their neutral form along the concentration gradient. In addition to this, protonophores can interact with proteins within the inner membrane [6].
2,4-Dinitrophenol (2,4-DNP) is commonly considered a classical protonophore. As early as 1933, it was discovered that the use of 2,4-DNP leads to rapid weight loss by enhancing the basal metabolism [19], resulting in the accelerated metabolism of fats and carbohydrates [20]. However, at that time, the concept of protonophores and the mechanism of 2,4-DNP action were not yet established. By 1938, the sale of 2,4-DNP without a prescription was prohibited, and shortly thereafter, it was completely banned [21]. The rapid weight loss in patients was accompanied by side effects attributable to specific metabolic characteristics. These included the shifting of the electrochemical gradient and dissipation of potential energy as heat, leading to uncontrolled hyperthermia [22], the inhibition of mitochondrial inorganic phosphate uptake [23], the excessive stimulation of glycolysis [24], and the accumulation of Na+ and K+ [25,26] (Supplementary Table).
The term protonophore was introduced by Skulachev in 1970 [27]. By this time, other protonophores from the hydrazone class, carbonyl cyanide-p-trifluoromethoxyphenyl hydrazone (FCCP) and carbonyl cyanide m-chlorophenyl hydrazone (CCCP), had already been discovered [28]. However, they also exhibit non-specific effects on other organelles, including the cytoplasmic membrane [29]. Their effects can induce both the depolarization and hyperpolarization of the cytoplasmic membrane by influencing H+, Na+, K+, and Ca2+ channels [30,31]. Among other well-studied protonophores, it is worth mentioning FR58P1 [32], BAM15 [33], C12TPP [32], C12R1 [34], CDE [35], C4R1 [36], bupivacaine [37], catechin [38], fisetin [38], quercetin [38], apigenin [38], usnic acid [39], and others. Other compounds can induce uncoupling effects by targeting the mitochondrial membrane protein PTEN-induced kinase 1 (PINK1) (niclosamide [40], triclosan [41], sertraline [41]; the metabolic regulator AMP activated protein kinase (AMPK) (curcumin [42], sorafenib [43], SR4 [44], FR58P1a [32], FH535 [45]), as well as the uncoupling proteins of the UCP family (T3 [46]) (see Supplementary Materials). In addition, some antioxidants have an uncoupling effect, which makes them more promising for further use by compounds [47]. Some compounds are derivatives of previously studied substances, developed with the aim of enhancing their therapeutic properties and reducing toxicity (niclosamide piperazine [48] and DNPME [49]) (Supplementary Materials).
Chemical mitochondrial uncouplers are actively investigated in scientific research for the development of new approaches in the treatment of neurodegenerative diseases such as Alzheimer’s disease [50], Parkinson’s disease, traumatic brain injury and stroke [36,51], ischemic heart disease [52,53], liver diseases [54], kidney diseases [33], as well as various forms of obesity [55,56], diabetes [54], and cancer [57]. In the middle of the 20th century, niclosamide was used as an antihelmintic drug. But then, its other properties were discovered [58] (see Supplementary Materials). However, it should be noted that their use for medical purposes requires further research and testing (Supplementary Materials).
This chapter provides only brief information about the nature and medical applications of synthetic uncouplers, as there are already numerous reviews dedicated to this topic. The focus of the current review will now shift towards transgenic animal and cellular models that simulate intense uncoupled and noncoupled respiration without pharmacological intervention.

3. Transgenic Models Which Overexpress UCPs

The most well-characterized and studied member of the UCP gene family is UCP1 (Figure 3C). However, adult humans exhibit minimal UCP1 expression. It is specifically expressed in brown adipose tissue, which is abundant in newborns and infants during early childhood. For a long time, it was believed that brown adipose tissue was absent in adults [59]. However, at the beginning of the 21st century, it was discovered that accumulations of brown adipose tissue, larger than 4 mm in diameter, are present in 7.5% of women and 3.1% of men. There are cervical, supraclavicular, and upper mediastinal depots of brown adipose tissue. [60]. The relatively high expression of UCP1 is also observed in the adrenal gland [61] (Figure 3A). UCP1 is expressed to a lesser extent in white adipose tissue (Figure 3B).
Various studies have shown that UCP1 is mainly expressed in the back subcutaneous adipose tissue, perirenal adipose tissue, or visceral adipose tissue [62]. In inbred Lou/C rats, which are a transgenic model of obesity resistance, increased Ucp1 expression was observed in white adipose tissue, potentially contributing to resistance to diet-induced obesity [63]. A transgenic model that expresses Ucp1 in gastrocnemius muscle (MCK-UCP1-20) showed lower body weight and specifically decreased muscle mass despite consuming the same amount of food. Importantly, no cardiac muscle pathology was found in the MCK-UCP1-13 mouse strain, which also expressed Ucp1 [64]. Moreover, the MCK-UCP1-13 mice exhibited improved functional recovery after heart ischemia/reperfusion [65]. Another transgenic model, HSA-mUCP1, with increased Ucp1 expression in the muscles, exhibited increased respiratory quotient levels, indicating overall increased glucose oxidation [66]. Additionally, the expression of Ucp1 in skeletal muscle reduced the risk of reverse electron transfer in the mitochondrial respiratory chain and ROS production [67]. It is known that reverse electron transport to complex I of the respiratory chain is one of the factors contributing to the ROS hyperproduction, particularly in in vitro mitochondrial systems [68]. The impact of Ucp1 expression was also studied in kidney injury models, in which the viral-based overexpression of UCP1 reduced the mitochondrial ROS generation and apoptosis in hypoxia-treated tubular epithelial cells [69] (Table 1).
UCP2 is a well-studied uncoupling protein (Figure 3C) that is mainly expressed in organs and cells associated with the immune system [85] (Figure 3A). The analysis of its expression patterns also shows high expression levels in almost all sections of the digestive system, adipose tissue, smooth muscle, lungs, and gallbladder (Figure 3B and Figure 4A). UCP2 is expressed at a lower level (compared to UCP4) in the brain (Figure 4C), predominantly in axons and axon terminals. The heat generated by axonal UCP2 modulates neurotransmission in homeostatic centers, coordinating the activity of brain circuits that regulate energy balance and related autonomic and endocrine processes [86]. Drosophila models overexpressing human hUCP2 in the nervous system have shown increased lifespan, reduced oxidative damage, and enhanced resistance to paraquat and rotenone, both of which are commonly used to create Parkinson’s disease models [70,71]. Similarly, transgenic mice overexpressing Ucp2 in catecholaminergic neurons exhibited similar effects. These mice showed a twofold elevation in Ucp2 expression in dopaminergic neurons of the substantia nigra, resulting in increased mitochondrial uncoupling. When acutely exposed to 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP), a neurotoxin used in Parkinson’s disease models, transgenic mice demonstrated neuroprotection and retained locomotor functions [60] (Table 1).
The impact of Ucp2 overexpression has been studied in various neurological disease models. Contradictory results have been obtained in amyotrophic lateral sclerosis (ALS) models. On the one hand, it has been shown that the overexpression of hUCP2 increased the survival age of superoxide dismutase 2 knockdown (sod2−/−) mice and reduced ROS production and oxidative stress throughout the aging process [72]. Conversely, other studies conducted on the same transgenic model showed that hUCP2 overexpression worsens mitochondrial dysfunction and accelerates ALS progression [73]. The positive effects of UCP2 overexpression have been consistently observed in ischemic disease models. The overexpression of Ucp2 protected thalamic neurons following global ischemia [74] and attenuated the increase in IL-6 levels and decrease in Bcl2 levels following focal ischemia [75]. Ucp2-overexpressing mice demonstrated faster recovery rates after middle cerebral artery occlusion (MCAO)-induced stroke and traumatic brain injury [76]. Rat transgenic models overexpressing Ucp2 have not been created. However, the injection of a lentiviral vector encoding UCP2 (LV-UCP2) into stroke-prone spontaneously hypertensive rats (SHRSP), fed with a high-salt Japanese-style diet, resulted in the delayed onset of stroke and kidney injury [87]. In transgenic mice constitutively expressing Ucp2 in the hippocampus prior to epileptic seizure induction, a substantial reduction in cell death was observed [78]. Ucp2 expression in transgenic animals decreased retinal ganglion cell degeneration and death in a mouse model of glaucoma [77]. It is worth noting that the overexpression of UCP2 has a therapeutic effect not only in neurological and neurodegenerative diseases. It has been shown that the targeted expression of Ucp2 in mouse liver increases susceptibility to acute liver injury induced by lipopolysaccharide and galactosamine [79] (Table 1).
UCP3 is primarily expressed in skeletal muscles [88] (Figure 3B). The data from the Protein Atlas indicates that the high expression of UCP3 is observed in the tongue, which is not contradictory to the previous statement, as the tongue is a muscle organ (Figure 4A). Mice overexpressing human UCP3 in skeletal muscle (UCP-3tg) exhibited hyperphagia along with a significant reduction in adipose tissue mass [80] and increased β-oxidation through a thioesterase-1-dependent mechanism in the mitochondria [81]. However, the overexpression of Ucp3 in the skeletal muscle of transgenic mice was also accompanied by an increase in muscle mitochondrial inefficiency, as indicated by a reduction in the ratio of ATP synthesis to mitochondrial oxidation [82] (Table 1).
UCP4, or solute carrier family 25 member 27 (SLC25A27), is primarily expressed in the brain and in male and female reproductive systems (Figure 4B). However, it plays a lesser role in uncoupling oxidative phosphorylation, as it is preferentially localized in close vicinity to VDAC, presumably at the inner boundary membrane of neuronal mitochondria, whereas F0F1-ATP synthase is more centrally located at the cristae membrane. Therefore, due to the distinctive mitochondrial morphology, UCP4 is unlikely to function as a direct uncoupler of oxidative phosphorylation. However, this observation supports the possibility that UCP4 may instead play a role in dissipating the excessive proton gradient typically linked to ROS production [89]. No information was found regarding transgenic models that overexpress UCP4. However, targeted overexpression was induced using lentiviruses and vectors. The viral-induced overexpression of UCP4 improved neuronal survival in vitro in a mouse model of Alzheimer’s disease and prevented spatial memory impairments in vivo in 3xTg mice [90]. The lentiviral-induced overexpression of UCP4 in astrocytes was found to promote neuronal survival. The reduction in ATP production was effectively compensated by an enhancement of glycolysis, which resulted in nonoxidative energy production without deleterious H2O2 generation. It was observed that astrocytes exhibiting higher levels of UCP4 produced increased amounts of lactate, which served as an energy source for neurons and facilitated enhanced neuronal survival [91].
UCP5 (also known as BMCP1, brain mitochondrial carrier protein-1) is similarly expressed predominantly in the nervous system (Figure 4C), like UCP4. It is the least studied member of the UCP family (Figure 3C), but cellular cultures overexpressing UCP5 have been obtained. Neuronal (GT1-1) cell lines with the stable overexpression of UCP5 showed a lower mitochondrial ∆μH+, indicating the stronger uncoupling of mitochondria, as well as reduced ATP production [83]. The stable overexpression of UCP5 provided protection against 1-methyl-4-phenylpyridinium (MPP(+))- and dopamine-induced toxicity in SH-SY5Y neuroblastoma cells [84] (Table 1).

4. Transgenic Animal Models in Which Components of Alternative Respiratory Pathways Are Expressed

The discovery of gene-encoded proteins for alternative respiratory pathways in the genomes of some animals in 2004 has provided important background for the future development of transgenic animal models [11]. Ciona intestinalis, an ascidian, has been widely used as a donor for the alternative oxidase (AOX) in transgenic models in numerous studies [92,93,94,95,96,97,98,99]. Additionally, in two studies, C. intestinalis was utilized as the source of alternative NADH dehydrogenases [100,101]. Saccharomyces cerevisiae has been the primary source of alternative NADH dehydrogenases in most studies [102,103,104,105]. There has also been reported a study in which alternative NADH dehydrogenases from plants were transfected into human cells [106] (Table 2).
There are two main therapeutic effects resulting from the expression of alternative respiratory pathways in insect and mammalian cells. The first effect is associated with the restoration of the respiratory rate following inhibition or damage to the subunits of respiratory chain complexes. The expression of alternative NADH dehydrogenases restored the respiration rate in cells with defective complex I cells [106], in flies with the reduced expression of the complex I assembly factor [102], as well as during complex I inhibition by rotenone [101,104,105] and paraquat [105]. The expression of AOX led to the restoration of the respiratory rate upon the inhibition of complex III by antimycin [99], complex IV by cyanide [93,97,99], azide [99], cigarette smoke condensate [96], mutations in genes encoding complex IV subunits [92,95,107], and gene knockdown responsible for complex IV assembly [95] as well as in complex III-deficient mice [98].
The second effect is related to the modulation of ROS metabolism. The expression of alternative NADH dehydrogenases induced a decrease in the rate of ROS production [102,104,105] and suppressed the levels of oxidative stress markers [102,105,106]. AOX expression reduced the rate of ROS production [95,99], levels of oxidative stress markers [94], and the rate of superoxide production induced by cyanide [93], antimycin [92,93,97], menadione [100], and cigarette smoke condensate [96].
The therapeutic potential of alternative respiratory pathways has been demonstrated in models of Alzheimer’s disease [94], Parkinson’s disease [95], Leigh syndrome [95], and cardiomyopathy [92,98]. It has been shown that AOX expression may be associated with the activation of signaling pathways linked to cell survival and protection against oxidative stress, particularly the Nrf2/ARE signaling pathway [101].
However, it would be incorrect to assume that the expression of genes encoding components of alternative respiratory pathways is capable of resolving all mitochondrial dysfunctions. The tko25t mutant Drosophila, which carries a recessive point mutation in the gene for mitoribosomal protein S12, demonstrates a decreased abundance of mitoribosomal small subunits, multiple respiratory chain dysfunctions, and ATP synthase deficiency [108]. The expression of AOX from C. intestinalis does not rescue the tko25t phenotype. Additionally, the expression of Ndi1 by S. cerevisiae during development is lethal for tko25t [109]. Moreover, the expression of Ndi1 exacerbates the neuronal phenotype resulting from complex IV subunit knockdown [107]. The overexpression of monocyte chemoattractant protein 1 (Mcp1) in mice cardiomyocytes induces inflammatory cardiomyopathy, leading to death from heart failure at the age of 7–8 months. AOX is unable to rescue heart failure directly caused by complex IV deficiency in mice overexpressing Mcp1 [110]. Concerns have been raised that a drastic decrease in the rate of superoxide production, dependent on AOX, may impair the functioning of signaling pathways associated with ROS metabolism [93]. Catania et al. (2019) showed that the affinity of alternative NADH dehydrogenase from Arabidopsis thaliana to NADH is over 3-fold higher than the affinity of complex I for NADH in human fibroblasts. This could potentially have a negative impact on ATP production and the metabolic status of the entire organism [106].

5. Conclusions

With rare exceptions, we observe that transgenic models simulating uncoupled and noncoupled respiration have shown positive effects in various disease models. However, it is important to recognize that implementing this approach as a therapeutic strategy is currently challenging and not yet practical in clinical practice. In the context of uncoupled respiration, synthetic uncouplers serve as an alternative for transgenic models that theoretically could be implemented in clinical practice not only for eliminating helminths [58] but also for treating metabolic and neurodegenerative diseases in humans. However, the situation is more difficult when it comes to the analogs of noncoupled respiration, which is characteristic of plants and certain types of sessile animals. Currently, we are not aware of any specific compounds that prevent the formation of a membrane potential without inhibiting any respiratory complexes. However, in our opinion, the closest model for simulating noncoupled respiration is provided by methylene blue, which facilitates alternative electron transport [111]. Methylene blue can accept electrons from NADH, the succinate dehydrogenase complex, and the alpha-glycerophosphate dehydrogenase complex, and then transfer them to cytochrome c [112]. Methylene blue can act as a bypass for an inhibited or damaged complex I [113], similar to how plant alternative NADH dehydrogenases function. As a result, when using methylene blue, electrons from the reducing equivalents pass through fewer coupling complexes compared to fully coupled respiration. Today, the intravenous injection of methylene blue is approved by the Food and Drug Administration (FDA) (Accession Number: DB09241) and European Medicines Agency (EMA) (Agency product number: EMEA/H/C/002108) for the treatment of patients with acquired methemoglobinemia. Clinical trials are underway for its potential therapeutic use in Alzheimer’s disease (Accession Number: NCT03446001). Studies suggest that methylene blue may slow the progression of Parkinson’s disease [114], Huntington’s disease [115], amyotrophic lateral sclerosis [116], and cognitive decline associated with aging [117,118,119,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138]. Consequently, targeting the mitochondrial respiratory chain to reduce the tension on the inner mitochondrial membrane or bypass inhibited or damaged respiratory complexes represents a promising direction that requires further investigation.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms242216491/s1

Author Contributions

Conceptualization, D.N.S. and E.Y.P.; writing—original draft preparation, A.P.G. and A.A.A.; writing—review and editing, D.N.S. and E.Y.P.; visualization, A.P.G.; funding acquisition, A.P.G. and E.Y.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Ministry of Science and Higher Education of the Russian Federation under the State assignment for universities in the field of scientific activity for 2020–2022 (project FZGU-2023-0009) and by the President grant for the support of leading scientific schools (project NSh-1375.2022.5). The analysis of the metabolic alterations was supported by the Russian Science Foundation (project #21-75-30009).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Skulachev, V.P. Uncoupling of Respiration and Phosphorylation. In Frontiers of Cellular Bioenergetics; Papa, S., Guerrieri, F., Eds.; Springer: Boston, MA, USA, 1999; pp. 89–118. [Google Scholar] [CrossRef]
  2. Jonckheere, A.I.; Smeitink, J.A.M.; Rodenburg, R.J.T. Mitochondrial ATP synthase: Architecture, function and pathology. J. Inherit. Metab. Dis. 2012, 35, 211–225. [Google Scholar] [CrossRef]
  3. Orrenius, S.; Gogvadze, V.; Zhivotovsky, B. Mitochondrial oxidative stress: Implications for cell death. Annu. Rev. Pharmacol. Toxicol. 2007, 47, 143–183. [Google Scholar] [CrossRef]
  4. Rajman, L.; Chwalek, K.; Sinclair, D.A. Therapeutic Potential of NAD-Boosting Molecules: The in vivo Evidence. Cell Metab. 2018, 27, 529–547. [Google Scholar] [CrossRef]
  5. Fedorenko, A.; Lishko, P.V.; Kirichok, Y. Mechanism of fatty-acid-dependent UCP1 uncoupling in brown fat mitochondria. Cell 2012, 151, 400–413. [Google Scholar] [CrossRef] [PubMed]
  6. Kotova, E.A.; Antonenko, Y.N. Fifty Years of Research on Protonophores: Mitochondrial Uncoupling as a Basis for Therapeutic Action. Acta Nat. 2022, 14, 4–13. [Google Scholar] [CrossRef] [PubMed]
  7. Shrestha, R.; Johnson, E.; Byrne, F.L. Exploring the therapeutic potential of mitochondrial uncouplers in cancer. Mol. Metab. 2021, 51, 101222. [Google Scholar] [CrossRef]
  8. Eubel, H.; Heinemeyer, J.; Sunderhaus, S.; Braun, H.P. Respiratory chain supercomplexes in plant mitochondria. Plant Physiol. Biochem. 2004, 42, 937–942. [Google Scholar] [CrossRef]
  9. Rich, P.R.; Maréchal, A. The mitochondrial respiratory chain. Essays Biochem. 2010, 47, 1–23. [Google Scholar] [CrossRef]
  10. Sousa, J.S.; D′Imprima, E.; Vonck, J. Mitochondrial Respiratory Chain Complexes. Subcell. Biochem. 2018, 87, 167–227. [Google Scholar] [CrossRef]
  11. McDonald, A.; Vanlerberghe, G. Branched mitochondrial electron transport in the Animalia: Presence of alternative oxidase in several animal phyla. IUBMB Life 2004, 56, 333–341. [Google Scholar] [CrossRef]
  12. Vicente, J.A.; Peixoto, F.; Lopes, M.L.; Madeira, V.M. Differential sensitivities of plant and animal mitochondria to the herbicide paraquat. J. Biochem. Mol. Toxicol. 2001, 15, 322–330. [Google Scholar] [CrossRef]
  13. Fernandez-Vizarra, E.; Zeviani, M. Mitochondrial disorders of the OXPHOS system. FEBS Lett. 2021, 595, 1062–1106. [Google Scholar] [CrossRef] [PubMed]
  14. Brookes, P.S. Mitochondrial H(+) leak and ROS generation: An odd couple. Free Radic. Biol. Med. 2005, 38, 12–23. [Google Scholar] [CrossRef] [PubMed]
  15. Childress, E.S.; Alexopoulos, S.J.; Hoehn, K.L.; Santos, W.L. Small Molecule Mitochondrial Uncouplers and Their Therapeutic Potential. J. Med. Chem. 2018, 61, 4641–4655. [Google Scholar] [CrossRef] [PubMed]
  16. Grundlingh, J.; Dargan, P.I.; El-Zanfaly, M.; Wood, D.M. 2,4-Dinitrophenol (DNP): A Weight Loss Agent with Significant Acute Toxicity and Risk of Death. J. Med. Toxicol. 2011, 7, 205–212. [Google Scholar] [CrossRef]
  17. Samartsev, V.N.; Semenova, A.A.; Dubinin, M.V. A Comparative Study of the Action of Protonophore Uncouplers and Decoupling Agents as Inducers of Free Respiration in Mitochondria in States 3 and 4: Theoretical and Experimental Approaches. Cell Biochem. Biophys. 2020, 78, 203–216. [Google Scholar] [CrossRef]
  18. Nicholls, D.G.; Ferguson, S.J. 2—Ion Transport Across Energy-Conserving Membranes. In Bioenergetics, 4th ed.; Nicholls, D.G., Ferguson, S.J., Eds.; Academic Press: Cambridge, MA, USA, 2013; pp. 13–25. [Google Scholar] [CrossRef]
  19. Tainter, M.L.; Stockton, A.B.; Cutting, W.C. Use of dinitrophenol in obesity and related conditions: A progress report. JAMA 1933, 101, 1472–1475. [Google Scholar] [CrossRef]
  20. Tainter, M.L.; Cutting, W.C.; Hines, E. Effects of moderated doses of dinitrophenol on the energy exchange and nitrogen metabolism of patient under condition of restricted dietary. J. Pharmacol. Exp. Ther. 1935, 55, 326–353. [Google Scholar]
  21. Colman, E. Dinitrophenol and obesity: An early twentieth-century regulatory dilemma. Regul. Toxicol. Pharmacol. 2007, 48, 115–117. [Google Scholar] [CrossRef]
  22. Harper, J.A.; Dickinson, K.; Brand, M.D. Mitochondrial uncoupling as a target for drug development for the treatment of obesity. Obes. Rev. 2001, 2, 255–265. [Google Scholar] [CrossRef]
  23. Rognstad, R.; Katz, J. The effect of 2,4-dinitrophenol on adipose-tissue metabolism. Biochem. J. 1969, 111, 431–444. [Google Scholar] [CrossRef] [PubMed]
  24. El-Guindy, M.M.; Neder, A.C.; Gomes, C.B. 2,4-Dinitrophenol--mechanism of action. Cell Mol. Biol. Incl. Cyto Enzymol. 1981, 27, 399–402. [Google Scholar] [PubMed]
  25. Moffatt, E.J.; Miyamoto, M.D. Effect of sodium and calcium channel blockade on the increase in spontaneous transmitter release produced by the mitochondrial inhibitor, dinitrophenol. J. Pharmacol. Exp. Ther. 1988, 244, 613–618. [Google Scholar] [PubMed]
  26. Mudge, G.H. Electrolyte and water metabolism of rabbit kidney slices; effect of metabolic inhibitors. Am. J. Physiol. 1951, 167, 206–223. [Google Scholar] [CrossRef] [PubMed]
  27. Skulachev, V.P. Electric fields in coupling membranes. FEBS Lett. 1970, 11, 301–308. [Google Scholar] [CrossRef]
  28. Heytler, P.G.; Prichard, W.W. A new class of uncoupling agents—Carbonyl cyanide phenylhydrazones. Biochem. Biophys. Res. Commun. 1962, 7, 272–275. [Google Scholar] [CrossRef]
  29. Demine, S.; Renard, P.; Arnould, T. Mitochondrial Uncoupling: A Key Controller of Biological Processes in Physiology and Diseases. Cells 2019, 8, 795. [Google Scholar] [CrossRef]
  30. Duchen, M.R. Effects of metabolic inhibition on the membrane properties of isolated mouse primary sensory neurones. J. Physiol. 1990, 424, 387–409. [Google Scholar] [CrossRef]
  31. Park, K.S.; Jo, I.; Pak, K.; Bae, S.W.; Rhim, H.; Suh, S.H.; Park, S.; Zhu, M.; So, I.; Kim, K. FCCP depolarizes plasma membrane potential by activating proton and Na+ currents in bovine aortic endothelial cells. Pflugers Arch. 2002, 443, 344–352. [Google Scholar] [CrossRef]
  32. Urra, F.A.; Muñoz, F.; Córdova-Delgado, M.; Ramírez, M.P.; Peña-Ahumada, B.; Rios, M.; Cruz, P.; Ahumada-Castro, U.; Bustos, G.; Silva-Pavez, E.; et al. FR58P1a; a new uncoupler of OXPHOS that inhibits migration in triple-negative breast cancer cells via Sirt1/AMPK/β1-integrin pathway. Sci. Rep. 2018, 8, 13190. [Google Scholar] [CrossRef]
  33. Kenwood, B.M.; Weaver, J.L.; Bajwa, A.; Poon, I.K.; Byrne, F.L.; Murrow, B.A.; Calderone, J.A.; Huang, L.; Divakaruni, A.S.; Tomsig, J.L.; et al. Identification of a novel mitochondrial uncoupler that does not depolarize the plasma membrane. Mol. Metab. 2013, 3, 114–123. [Google Scholar] [CrossRef]
  34. Plotnikov, E.Y.; Silachev, D.N.; Jankauskas, S.S.; Rokitskaya, T.I.; Chupyrkina, A.A.; Pevzner, I.B.; Zorova, L.D.; Isaev, N.K.; Antonenko, Y.N.; Skulachev, V.P.; et al. Mild uncoupling of respiration and phosphorylation as a mechanism providing nephro- and neuroprotective effects of penetrating cations of the SkQ family. Biochemistry 2012, 77, 1029–1037. [Google Scholar] [CrossRef] [PubMed]
  35. Dalla Via, L.; García-Argáez, A.N.; Braga, A.; Martinez-Vazquez, M.; Grancara, S.; Martinis, P.; Agostinelli, E.; Toninello, A. An eudesman derivative from Verbesina persicifolia D.C. as a natural mild uncoupler in liver mitochondria. A new potential anti-obesity agent? Curr. Pharm. Des. 2014, 20, 253–261. [Google Scholar] [CrossRef] [PubMed]
  36. Khailova, L.S.; Silachev, D.N.; Rokitskaya, T.I.; Armine, V.; Avetisyan, A.V.; Lyamsaev, K.G.; Severina, I.I.; Il′yasova, T.M.; Gulyaev, M.V.; Dedukhova, V.I.; et al. A short-chain alkyl derivative of Rhodamine 19 acts as a mild uncoupler of mitochondria and a neuroprotector. Biochim. Biophys. Acta 2014, 1837, 1739–1747. [Google Scholar] [CrossRef] [PubMed]
  37. Sztark, F.; Ouhabi, R.; Dabadie, P.; Mazat, J.P. Effects of the local anesthetic bupivacaine on mitochondrial energy metabolism: Change from uncoupling to decoupling depending on the respiration state. Biochem. Mol. Biol. Int. 1997, 43, 997–1003. [Google Scholar] [CrossRef]
  38. Cho, I.; Song, H.O.; Cho, J.H. Flavonoids mitigate neurodegeneration in aged Caenorhabditis elegans by mitochondrial uncoupling. Food Sci. Nutr. 2020, 8, 6633–6642. [Google Scholar] [CrossRef]
  39. Han, D.; Matsumaru, K.; Rettori, D.; Kaplowitz, N. Usnic acid-induced necrosis of cultured mouse hepatocytes: Inhibition of mitochondrial function and oxidative stress. Biochem. Pharmacol. 2004, 67, 439–451. [Google Scholar] [CrossRef]
  40. Kadri, H.; Lambourne, O.A.; Mehellou, Y. Niclosamide, a Drug with Many (Re)purposes. ChemMedChem 2018, 13, 1088–1091. [Google Scholar] [CrossRef]
  41. Tjahjono, E.; Pei, J.; Revtovich, A.V.; Liu, T.J.E.; Swadi, A.; Hancu, M.C.; Tolar, J.G.; Kirienko, N.V. Mitochondria-affecting small molecules ameliorate proteostasis defects associated with neurodegenerative diseases. Sci. Rep. 2021, 11, 17733. [Google Scholar] [CrossRef]
  42. Lim, H.W.; Lim, H.Y.; Wong, K.P. Uncoupling of oxidative phosphorylation by curcumin: Implication of its cellular mechanism of action. Biochem. Biophys. Res. Commun. 2009, 389, 187–192. [Google Scholar] [CrossRef]
  43. Jian, C.; Fu, J.; Cheng, X.; Shen, L.J.; Ji, Y.X.; Wang, X.; Pan, S.; Tian, H.; Tian, S.; Liao, R.; et al. Low-Dose Sorafenib Acts as a Mitochondrial Uncoupler and Ameliorates Nonalcoholic Steatohepatitis. Cell Metab. 2020, 31, 892–908.e11. [Google Scholar] [CrossRef] [PubMed]
  44. Figarola, J.L.; Singhal, J.; Tompkins, J.D.; Rogers, G.W.; Warden, C.; Horne, D.; Riggs, A.D.; Awasthi, S.; Singhal, S.S. SR4 Uncouples Mitochondrial Oxidative Phosphorylation, Modulates AMP-dependent Kinase (AMPK)-Mammalian Target of Rapamycin (mTOR) Signaling, and Inhibits Proliferation of HepG2 Hepatocarcinoma Cells. J. Biol. Chem. 2015, 290, 30321–30341. [Google Scholar] [CrossRef]
  45. Zhang, W.; Sviripa, V.M.; Kril, L.M.; Yu, T.; Xie, Y.; Hubbard, W.B.; Sullivan, P.G.; Chen, X.; Zhan, C.G.; Yang-Hartwich, Y.; et al. An Underlying Mechanism of Dual Wnt Inhibition and AMPK Activation: Mitochondrial Uncouplers Masquerading as Wnt Inhibitors. J. Med. Chem. 2019, 62, 11348–11358. [Google Scholar] [CrossRef] [PubMed]
  46. Yau, W.W.; Singh, B.K.; Lesmana, R.; Zhou, J.; Sinha, R.A.; Wong, K.A.; Wu, Y.; Bay, B.H.; Sugii, S.; Sunet, L.; et al. Thyroid hormone (T3) stimulates brown adipose tissue activation via mitochondrial biogenesis and MTOR-mediated mitophagy. Autophagy 2019, 15, 131–150. [Google Scholar] [CrossRef] [PubMed]
  47. Anisimov, V.N.; Bakeeva, L.E.; Egormin, P.A.; Filenko, O.F.; Isakova, E.F.; Manskikh, V.N.; Mikhelson, V.M.; Panteleeva, A.A.; Pasyukova, E.G.; Pilipenko, D.I.; et al. Mitochondria-targeted plastoquinone derivatives as tools to interrupt execution of the aging program. 5. SkQ1 prolongs lifespan and prevents development of traits of senescence. Biochemistry 2008, 73, 1329–1342. [Google Scholar] [CrossRef]
  48. Guo, J.; Tao, H.; Alasadi, A.; Huang, Q.; Jin, S. Niclosamide piperazine prevents high-fat diet-induced obesity and diabetic symptoms in mice. Eat. Weight. Disord. 2019, 24, 91–96. [Google Scholar] [CrossRef]
  49. Perry, R.J.; Kim, T.; Zhang, X.M.; Lee, H.Y.; Pesta, D.; Popov, V.B.; Zhang, D.; Rahimi, Y.; Jurczak, M.J.; Cline, G.W.; et al. Reversal of hypertriglyceridemia, fatty liver disease, and insulin resistance by a liver-targeted mitochondrial uncoupler. Cell Metab. 2013, 18, 740–748. [Google Scholar] [CrossRef]
  50. Geisler, J.G.; Marosi, K.; Halpern, J.; Mattson, M.P. DNP, mitochondrial uncoupling, and neuroprotection: A little dab′ll do ya. Alzheimers Dement. 2017, 13, 582–591. [Google Scholar] [CrossRef]
  51. Silachev, D.N.; Khailova, L.S.; Babenko, V.A.; Gulyaev, M.V.; Kovalchuk, S.I.; Zorova, L.D.; Plotnikov, E.Y.; Antonenko, Y.N.; Zorov, D.B. Neuroprotective effect of glutamate-substituted analog of gramicidin A is mediated by the uncoupling of mitochondria. Biochim. Biophys. Acta 2014, 1840, 3434–3442. [Google Scholar] [CrossRef]
  52. Gao, J.L.; Zhao, J.; Zhu, H.B.; Peng, X.; Zhu, J.X.; Ma, M.H.; Fu, Y.; Hu, N.; Tai, Y.; Xuan, X.C.; et al. Characterizations of mitochondrial uncoupling induced by chemical mitochondrial uncouplers in cardiomyocytes. Free Radic. Biol. Med. 2018, 124, 288–298. [Google Scholar] [CrossRef]
  53. Minners, J.; van den Bos, E.J.; Yellon, D.M.; Schwalb, H.; Opie, L.H.; Sack, M.N. Dinitrophenol, cyclosporin A, and trimetazidine modulate preconditioning in the isolated rat heart: Support for a mitochondrial role in cardioprotection. Cardiovasc. Res. 2000, 47, 68–73. [Google Scholar] [CrossRef] [PubMed]
  54. Tao, H.; Zhang, Y.; Zeng, X.; Shulman, G.I.; Jin, S. Niclosamide ethanolamine-induced mild mitochondrial uncoupling improves diabetic symptoms in mice. Nat. Med. 2014, 20, 1263–1269. [Google Scholar] [CrossRef] [PubMed]
  55. Geisler, J.G. 2,4 Dinitrophenol as Medicine. Cells 2019, 8, 280. [Google Scholar] [CrossRef]
  56. Kalinovich, A.V.; Shabalina, I.G. Novel Mitochondrial Cationic Uncoupler C4R1 Is an Effective Treatment for Combating Obesity in Mice. Biochemistry 2015, 80, 620–628. [Google Scholar] [CrossRef] [PubMed]
  57. Singhal, S.S.; Figarola, J.; Singhal, J.; Leake, K.; Nagaprashantha, L.; Lincoln, C.; Gugiu, B.G.; Horne, D.; Jove, R.; Awasthi, S.; et al. 1,3-Bis(3,5-dichlorophenyl) urea compound ′COH-SR4′ inhibits proliferation and activates apoptosis in melanoma. Biochem. Pharmacol. 2012, 84, 1419–1427. [Google Scholar] [CrossRef]
  58. Pearson, R.D.; Hewlett, E.L. Niclosamide therapy for tapeworm infections. Ann. Intern. Med. 1985, 102, 550–551. [Google Scholar] [CrossRef]
  59. Cannon, B.; Nedergaard, J. Brown adipose tissue: Function and physiological significance. Physiol. Rev. 2004, 84, 277–359. [Google Scholar] [CrossRef]
  60. Conti, B.; Sugama, S.; Lucero, J.; Winsky-Sommerer, R.; Wirz, S.A.; Maher, P.; Andrews, Z.; Barr, A.M.; Morale, M.C.; Paneda, C.; et al. Uncoupling protein 2 protects dopaminergic neurons from acute 1,2,3,6-methyl-phenyl-tetrahydropyridine toxicity. J. Neurochem. 2005, 93, 493–501. [Google Scholar] [CrossRef]
  61. Bergman, J.; Botling, J.; Fagerberg, L.; Hallström, B.M.; Djureinovic, D.; Uhlén, M.; Pontén, F. The Human Adrenal Gland Proteome Defined by Transcriptomics and Antibody-Based Profiling. Endocrinology 2017, 158, 239–251. [Google Scholar] [CrossRef]
  62. Lim, J.; Park, H.S.; Kim, J.; Jang, Y.J.; Kim, J.-H.; Lee, Y.; Heo, Y. Depot-specific UCP1 expression in human white adipose tissue and its association with obesity-related markers. Int. J. Obes. 2020, 44, 697–706. [Google Scholar] [CrossRef]
  63. Poher, A.L.; Veyrat-Durebex, C.; Altirriba, J.; Montet, X.; Colin, D.J.; Caillon, A.; Lyautey, J.; Rohner-Jeanrenaud, F. Ectopic UCP1 Overexpression in White Adipose Tissue Improves Insulin Sensitivity in Lou/C Rats, a Model of Obesity Resistance. Diabetes 2015, 64, 3700–3712. [Google Scholar] [CrossRef]
  64. Couplan, E.; Gelly, C.; Goubern, M.; Fleury, C.; Quesson, B.; Silberberg, M.; Thiaudière, E.; Mateo, P.; Lonchampt, M.; Levens, N.; et al. High level of uncoupling protein 1 expression in muscle of transgenic mice selectively affects muscles at rest and decreases their IIb fiber content. J. Biol. Chem. 2002, 277, 43079–43088. [Google Scholar] [CrossRef] [PubMed]
  65. Hoerter, J.; Gonzalez-Barroso, M.D.; Couplan, E.; Mateo, P.; Gelly, C.; Cassard-Doulcier, A.-M.; Diolez, P.; Bouillaud, F. Mitochondrial uncoupling protein 1 expressed in the heart of transgenic mice protects against ischemic-reperfusion damage. Circulation 2004, 110, 528–533. [Google Scholar] [CrossRef] [PubMed]
  66. Klaus, S.; Rudolph, B.; Dohrmann, C.; Wehr, R. Expression of uncoupling protein 1 in skeletal muscle decreases muscle energy efficiency and affects thermoregulation and substrate oxidation. Physiol. Genomics 2005, 21, 193–200. [Google Scholar] [CrossRef] [PubMed]
  67. Keipert, S.; Klaus, S.; Heldmaier, G.; Jastroch, M. UCP1 ectopically expressed in murine muscle displays native function and mitigates mitochondrial superoxide production. Biochim. Biophys. Acta 2010, 1797, 324–330. [Google Scholar] [CrossRef]
  68. Andreyev, A.Y.; Kushnareva, Y.E.; Starkov, A.A. Mitochondrial metabolism of reactive oxygen species. Biochemistry 2005, 70, 200–214. [Google Scholar] [CrossRef] [PubMed]
  69. Jia, P.; Wu, X.; Pan, T.; Xu, S.; Hu, J.; Ding, X. Uncoupling protein 1 inhibits mitochondrial reactive oxygen species generation and alleviates acute kidney injury. EBioMedicine 2019, 49, 331–340. [Google Scholar] [CrossRef]
  70. Fridell, Y.W.; Sánchez-Blanco, A.; Silvia, B.A.; Helfand, S.L. Targeted expression of the human uncoupling protein 2 (hUCP2) to adult neurons extends life span in the fly. Cell Metab. 2005, 1, 145–152. [Google Scholar] [CrossRef] [PubMed]
  71. Islam, R.; Yang, L.; Sah, M.; Kannan, K.; Anamani, D.; Vijayan, C.; Kwok, J.; Cantino, M.E.; Beal, M.F.; Fridell, Y.-W.C. A neuroprotective role of the human uncoupling protein 2 (hUCP2) in a Drosophila Parkinson′s disease model. Neurobiol. Dis. 2012, 46, 137–146. [Google Scholar] [CrossRef]
  72. Andrews, Z.B.; Horvath, T.L. Uncoupling protein-2 regulates lifespan in mice. Am. J. Physiol. Endocrinol. Metab. 2009, 296, E621–E627. [Google Scholar] [CrossRef]
  73. Peixoto, P.M.; Kim, H.J.; Sider, B.; Starkov, A.; Horvath, T.L.; Manfredi, G. UCP2 overexpression worsens mitochondrial dysfunction and accelerates disease progression in a mouse model of amyotrophic lateral sclerosis. Mol. Cell Neurosci. 2013, 57, 104–110. [Google Scholar] [CrossRef] [PubMed]
  74. Deierborg, T.; Wieloch, T.; Diano, S.; Warden, C.H.; Horvath, T.L.; Mattiasson, G. Overexpression of UCP2 protects thalamic neurons following global ischemia in the mouse. J. Cereb. Blood Flow Metab. 2008, 28, 1186–1195. [Google Scholar] [CrossRef] [PubMed]
  75. Haines, B.; Li, P.A. Overexpression of mitochondrial uncoupling protein 2 inhibits inflammatory cytokines and activates cell survival factors after cerebral ischemia. PLoS ONE 2012, 7, e31739. [Google Scholar] [CrossRef]
  76. Mattiasson, G.; Shamloo, M.; Gido, G.; Mathi, K.; Tomasevic, G.; Yi, S.; Warden, C.H.; Castilho, R.F.; Melcher, T.; Gonzalez-Zulueta, M.; et al. Uncoupling protein-2 prevents neuronal death and diminishes brain dysfunction after stroke and brain trauma. Nat. Med. 2003, 9, 1062–1068. [Google Scholar] [CrossRef] [PubMed]
  77. Hass, D.T.; Barnstable, C.J. Cell Autonomous Neuroprotection by the Mitochondrial Uncoupling Protein 2 in a Mouse Model of Glaucoma. Front. Neurosci. 2019, 13, 201. [Google Scholar] [CrossRef]
  78. Diano, S.; Matthews, R.T.; Patrylo, P.; Yang, L.; Beal, M.F.; Barnstable, C.J.; Horvath, T.L. Uncoupling protein 2 prevents neuronal death including that occurring during seizures: A mechanism for preconditioning. Endocrinology 2003, 144, 5014–5021. [Google Scholar] [CrossRef]
  79. Shang, Y.; Liu, Y.; Du, L.; Wang, Y.; Cheng, X.; Xiao, W.; Wang, X.; Jin, H.; Yang, X.; Liu, S.; et al. Targeted expression of uncoupling protein 2 to mouse liver increases the susceptibility to lipopolysaccharide/galactosamine-induced acute liver injury. Hepatology 2009, 50, 1204–1216. [Google Scholar] [CrossRef]
  80. Li, H.; Wang, C.; Li, L.; Li, L. Skeletal muscle non-shivering thermogenesis as an attractive strategy to combat obesity. Life Sci. 2021, 269, 119024. [Google Scholar] [CrossRef]
  81. Moore, G.B.; Himms-Hagen, J.; Harper, M.E.; Clapham, J.C. Overexpression of UCP-3 in skeletal muscle of mice results in increased expression of mitochondrial thioesterase mRNA. Biochem. Biophys. Res. Commun. 2001, 283, 785–790. [Google Scholar] [CrossRef]
  82. Codella, R.; Alves, T.C.; Befroy, D.E.; Choi, C.S.; Luzi, L.; Rothman, D.L.; Kibbey, R.G.; Shulman, G.I. Overexpression of UCP3 decreases mitochondrial efficiency in mouse skeletal muscle in vivo. FEBS Lett. 2023, 597, 309–319. [Google Scholar] [CrossRef]
  83. Kim-Han, J.S.; Reichert, S.A.; Quick, K.L.; Dugan, L.L. BMCP1: A mitochondrial uncoupling protein in neurons which regulates mitochondrial function and oxidant production. J. Neurochem. 2001, 79, 658–668. [Google Scholar] [CrossRef]
  84. Kwok, K.H.; Ho, P.W.; Chu, A.C.; Ho, J.W.-M.; Liu, H.-F.; Yiu, D.C.-W.; Chan, K.-H.; Kung, M.H.-W.; Ramsden, D.B.; Ho, S.-L. Mitochondrial UCP5 is neuroprotective by preserving mitochondrial membrane potential, ATP levels, and reducing oxidative stress in MPP+ and dopamine toxicity. Free Radic. Biol. Med. 2010, 49, 1023–1035. [Google Scholar] [CrossRef] [PubMed]
  85. Rupprecht, A.; Bräuer, A.U.; Smorodchenko, A.; Goyn, J.; Hilse, K.E.; Shabalina, I.G.; Infante-Duarte, C.; Pohl, E.E. Quantification of uncoupling protein 2 reveals its main expression in immune cells and selective up-regulation during T-cell proliferation. PLoS ONE 2012, 7, e41406. [Google Scholar] [CrossRef] [PubMed]
  86. Horvath, T.L.; Warden, C.H.; Hajos, M.; Lombardi, A.; Goglia, F.; Diano, S. Brain uncoupling protein 2: Uncoupled neuronal mitochondria predict thermal synapses in homeostatic centers. J. Neurosci. 1999, 19, 10417–10427. [Google Scholar] [CrossRef] [PubMed]
  87. Busceti, C.L.; Cotugno, M.; Bianchi, F.; Forte, M.; Stanzione, R.; Marchitti, S.; Battaglia, G.; Nicoletti, F.; Fornai, F.; Rubattu, S. Brain Overexpression of Uncoupling Protein-2 (UCP2) Delays Renal Damage and Stroke Occurrence in Stroke-Prone Spontaneously Hypertensive Rats. Int. J. Mol. Sci. 2020, 21, 4289. [Google Scholar] [CrossRef] [PubMed]
  88. Clapham, J.C.; Arch, J.R.; Chapman, H.; Haynes, A.; Lister, C.; Moore, G.B.T.; Piercy, V.; Carter, S.A.; Lehner, I.; Smith, S.A.; et al. Mice overexpressing human uncoupling protein-3 in skeletal muscle are hyperphagic and lean. Nature 2000, 406, 415–418. [Google Scholar] [CrossRef] [PubMed]
  89. Klotzsch, E.; Smorodchenko, A.; Löfler, L.; Moldzio, R.; Parkinson, E.; Schütz, G.J.; Pohl, E.E. Superresolution microscopy reveals spatial separation of UCP4 and F0F1-ATP synthase in neuronal mitochondria. Proc. Natl. Acad. Sci. USA 2015, 112, 130–135. [Google Scholar] [CrossRef]
  90. Rosenberg, N.; Reva, M.; Binda, F.; Restivo, L.; Depierre, P.; Puyal, J.; Briquet, M.; Bernardinelli, Y.; Rocher, A.; Markram, H.; et al. Overexpression of UCP4 in astrocytic mitochondria prevents multilevel dysfunctions in a mouse model of Alzheimer’s disease. Glia 2023, 71, 957–973. [Google Scholar] [CrossRef]
  91. Perreten Lambert, H.; Zenger, M.; Azarias, G.; Chatton, J.Y.; Magistretti, P.J.; Lengacher, S. Control of mitochondrial pH by uncoupling protein 4 in astrocytes promotes neuronal survival. J. Biol. Chem. 2014, 289, 31014–31028. [Google Scholar] [CrossRef]
  92. Dassa, E.P.; Dufour, E.; Gonçalves, S.; Paupe, V.; Hakkaart, G.A.; Jacobs, H.T.; Rustin, P. Expression of the alternative oxidase complements cytochrome c oxidase deficiency in human cells. EMBO Mol. Med. 2009, 1, 30–36. [Google Scholar] [CrossRef]
  93. El-Khoury, R.; Dufour, E.; Rak, M.; Ramanantsoa, N.; Grandchamp, N.; Csaba, Z.; Duvillié, B.; Bénit, P.; Gallego, J.; Gressens, P.; et al. Alternative oxidase expression in the mouse enables bypassing cytochrome c oxidase blockade and limits mitochondrial ROS overproduction. PLoS Genet. 2013, 9, e1003182. [Google Scholar] [CrossRef]
  94. El-Khoury, R.; Kaulio, E.; Lassila, K.A.; Crowther, D.C.; Jacobs, H.T.; Rustin, P. Expression of the alternative oxidase mitigates beta-amyloid production and toxicity in model systems. Free Radic. Biol. Med. 2016, 96, 57–66. [Google Scholar] [CrossRef] [PubMed]
  95. Fernandez-Ayala, D.J.; Sanz, A.; Vartiainen, S.; Kemppainen, K.K.; Babusiak, M.; Mustalahti, E.; Costa, R.; Tuomela, T.; Zeviani, M.; Chung, J.; et al. Expression of the Ciona intestinalis alternative oxidase (AOX) in Drosophila complements defects in mitochondrial oxidative phosphorylation. Cell Metab. 2009, 9, 449–460. [Google Scholar] [CrossRef] [PubMed]
  96. Giordano, L.; Farnham, A.; Dhandapani, P.K.; Salminen, L.; Bhaskaran, J.; Voswinckel, R.; Rauschkolb, P.; Scheibe, S.; Sommer, N.; Beisswenger, C.; et al. Alternative Oxidase Attenuates Cigarette Smoke-induced Lung Dysfunction and Tissue Damage. Am. J. Respir. Cell Mol. Biol. 2019, 60, 515–522. [Google Scholar] [CrossRef]
  97. Hakkaart, G.A.; Dassa, E.P.; Jacobs, H.T.; Rustin, P. Allotopic expression of a mitochondrial alternative oxidase confers cyanide resistance to human cell respiration. EMBO Rep. 2006, 7, 341–345. [Google Scholar] [CrossRef]
  98. Rajendran, J.; Purhonen, J.; Tegelberg, S.; Smolander, O.P.; Mörgelin, M.; Rozman, J.; Gailus-Durner, V.; Fuchs, H.; de Angelis, M.H.; Auvinen, P.; et al. Alternative oxidase-mediated respiration prevents lethal mitochondrial cardiomyopathy. EMBO Mol. Med. 2019, 11, e9456. [Google Scholar] [CrossRef]
  99. Szibor, M.; Dhandapani, P.K.; Dufour, E.; Holmström, K.M.; Zhuang, Y.; Salwig, I.; Wittig, I.; Heidler, J.; Gizatullina, Z.; Gainutdinov, T.; et al. Broad AOX expression in a genetically tractable mouse model does not disturb normal physiology. Dis. Model. Mech. 2017, 10, 163–171. [Google Scholar] [CrossRef]
  100. Gospodaryov, D.V.; Lushchak, O.V.; Rovenko, B.M.; Perkhulyn, N.V.; Gerards, M.; Tuomela, T.; Jacobs, H.T. Ciona intestinalis NADH dehydrogenase NDX confers stress-resistance and extended lifespan on Drosophila. Biochim. Biophys. Acta 2014, 1837, 1861–1869. [Google Scholar] [CrossRef] [PubMed]
  101. Gospodaryov, D.V.; Strilbytska, O.M.; Semaniuk, U.V.; Perkhulyn, N.V.; Rovenko, B.M.; Yurkevych, I.S.; Barata, A.G.; Dick, T.P.; Lushchak, O.V.; Jacobs, H.T. Alternative NADH dehydrogenase extends lifespan and increases resistance to xenobiotics in Drosophila. Biogerontology 2020, 21, 155–171. [Google Scholar] [CrossRef]
  102. Bahadorani, S.; Cho, J.; Lo, T.; Contreras, H.; Lawal, H.O.; Krantz, D.E.; Bradley, T.J.; Walker, D.W. Neuronal expression of a single-subunit yeast NADH-ubiquinone oxidoreductase (Ndi1) extends Drosophila lifespan. Aging Cell 2010, 9, 191–202. [Google Scholar] [CrossRef]
  103. Cho, J.; Hur, J.H.; Graniel, J.; Benzer, S.; Walker, D.W. Expression of yeast NDI1 rescues a Drosophila complex I assembly defect. PLoS ONE 2012, 7, e50644. [Google Scholar] [CrossRef] [PubMed]
  104. Hur, J.H.; Bahadorani, S.; Graniel, J.; Koehler, C.L.; Ulgherait, M.; Rera, M.; Jones, D.L.; Walker, D.W. Increased longevity mediated by yeast NDI1 expression in Drosophila intestinal stem and progenitor cells. Aging 2013, 5, 662–681. [Google Scholar] [CrossRef] [PubMed]
  105. Sanz, A.; Soikkeli, M.; Portero-Otín, M.; Wilson, A.; Kemppainen, E.; McIlroy, G.; Ellilä, S.; Kemppainen, K.K.; Tuomela, T.; Lakanmaa, M.; et al. Expression of the yeast NADH dehydrogenase Ndi1 in Drosophila confers increased lifespan independently of dietary restriction. Proc. Natl. Acad. Sci. USA 2010, 107, 9105–9110. [Google Scholar] [CrossRef]
  106. Catania, A.; Iuso, A.; Bouchereau, J.; Kremer, L.S.; Paviolo, M.; Terrile, C.; Bénit, P.; Rasmusson, A.G.; Schwarzmayr, T.; Tiranti, V.; et al. Arabidopsis thaliana alternative dehydrogenases: A potential therapy for mitochondrial complex I deficiency? Perspectives and pitfalls. Orphanet J. Rare Dis. 2019, 14, 236. [Google Scholar] [CrossRef] [PubMed]
  107. Kemppainen, K.K.; Rinne, J.; Sriram, A.; Lakanmaa, M.; Zeb, A.; Tuomela, T.; Popplestone, A.; Singh, S.; Sanz, A.; Rustin, P.; et al. Expression of alternative oxidase in Drosophila ameliorates diverse phenotypes due to cytochrome oxidase deficiency. Hum. Mol. Genet. 2014, 23, 2078–2093. [Google Scholar] [CrossRef]
  108. Toivonen, J.M.; O′Dell, K.M.; Petit, N.; Irvine, S.C.; Knight, G.K.; Lehtonen, M.; Longmuir, M.; Luoto, K.; Touraille, S.; Wang, Z.; et al. Technical knockout, a Drosophila model of mitochondrial deafness. Genetics 2001, 159, 241–254. [Google Scholar] [CrossRef] [PubMed]
  109. Kemppainen, K.K.; Kemppainen, E.; Jacobs, H.T. The alternative oxidase AOX does not rescue the phenotype of tko25t mutant flies. G3 Genes Genomes Genet. 2014, 4, 2013–2021. [Google Scholar] [CrossRef]
  110. Dhandapani, P.K.; Begines-Moreno, I.M.; Brea-Calvo, G.; Gärtner, U.; Graeber, T.G.; Sanchez, G.J.; Morty, R.E.; Schönig, K.; Hoeve, J.T.; Wietelmann, A.; et al. Hyperoxia but not AOX expression mitigates pathological cardiac remodeling in a mouse model of inflammatory cardiomyopathy. Sci. Rep. 2019, 9, 12741. [Google Scholar] [CrossRef]
  111. Gureev, A.P.; Sadovnikova, I.S.; Popov, V.N. Molecular Mechanisms of the Neuroprotective Effect of Methylene Blue. Biochemistry 2022, 87, 940–956. [Google Scholar] [CrossRef]
  112. Tretter, L.; Horvath, G.; Hölgyesi, A.; Essek, F.; Adam-Vizi, V. Enhanced hydrogen peroxide generation accompanies the beneficial bioenergetic effects of methylene blue in isolated brain mitochondria. Free Radic. Biol. Med. 2014, 77, 317–330. [Google Scholar] [CrossRef]
  113. Gureev, A.P.; Shaforostova, E.A.; Popov, V.N.; Starkov, A.A. Methylene blue does not bypass Complex III antimycin block in mouse brain mitochondria. FEBS Lett. 2019, 593, 499–503. [Google Scholar] [CrossRef] [PubMed]
  114. Bariotto-Dos-Santos, K.; Padovan-Neto, F.E.; Bortolanza, M.; Dos-Santos-Pereira, M.; Raisman-Vozari, R.; Tumas, V.; Del Bel, E. Repurposing an established drug: An emerging role for methylene blue in L-DOPA-induced dyskinesia. Eur. J. Neurosci. 2019, 49, 869–882. [Google Scholar] [CrossRef] [PubMed]
  115. Heidari, R.; Monnier, V.; Martin, E.; Tricoire, H. Methylene Blue Partially Rescues Heart Defects in a Drosophila Model of Huntington′s Disease. J. Huntingtons Dis. 2015, 4, 173–186. [Google Scholar] [CrossRef] [PubMed]
  116. Dibaj, P.; Zschüntzsch, J.; Steffens, H.; Scheffel, J.; Göricke, B.; Weishaupt, J.H.; Le Meur, K.; Kirchhoff, F.; Hanisch, U.-K.; Schomburg, E.D.; et al. Influence of methylene blue on microglia-induced inflammation and motor neuron degeneration in the SOD1(G93A) model for ALS. PLoS ONE 2012, 7, e43963. [Google Scholar] [CrossRef] [PubMed]
  117. Sadovnikova, I.S.; Gureev, A.P.; Ignatyeva, D.A.; Gryaznova, M.V.; Chernyshova, E.V.; Krutskikh, E.P.; Novikova, A.G.; Popov, V.N. Nrf2/ARE Activators Improve Memory in Aged Mice via Maintaining of Mitochondrial Quality Control of Brain and the Modulation of Gut Microbiome. Pharmaceuticals 2021, 14, 607. [Google Scholar] [CrossRef]
  118. Alasadi, A.; Chen, M.; Swapna, G.V.T.; Tao, H.; Guo, J.; Collantes, J.; Jin, S. Effect of mitochondrial uncouplers niclosamide ethanolamine (NEN) and oxyclozanide on hepatic metastasis of colon cancer. Cell Death Dis. 2018, 9, 215. [Google Scholar] [CrossRef]
  119. Alexopoulos, S.J.; Chen, S.Y.; Brandon, A.E.; Salamoun, J.M.; Byrne, F.L.; Garcia, C.J.; Hoehn, K.L. Mitochondrial uncoupler BAM15 reverses diet-induced obesity and insulin resistance in mice. Nat. Commun. 2020, 11, 2397. [Google Scholar] [CrossRef]
  120. Amireddy, N.; Puttapaka, S.N.; Vinnakota, R.L.; Ravuri, H.G.; Thonda, S.; Kalivendi, S.V. The unintended mitochondrial uncoupling effects of the FDA-approved anti-helminth drug nitazoxanide mitigates experimental parkinsonism in mice. J. Biol. Chem. 2021, 292, 15731–15743, Corrected in J. Biol. Chem. 2021, 297, 100864.. [Google Scholar] [CrossRef]
  121. Chen, Y.; Du, F.; Tang, L.; Xu, J.; Zhao, Y.; Wu, X.; Xiao, Z. Carboranes as unique pharmacophores in antitumor medicinal chemistry. Mol. Ther. Oncolytics. 2022, 24, 400–416. [Google Scholar] [CrossRef]
  122. Dabadie, P.; Bendriss, P.; Erny, P.; Mazat, J.P. Uncoupling effects of local anesthetics on rat liver mitochondria. FEBS Lett. 1987, 226, 77–82. [Google Scholar] [CrossRef]
  123. Fu, Y.Y.; Zhang, M.; Turner, N.; Zhang, L.N.; Dong, T.C.; Gu, M.; Li, J. A novel chemical uncoupler ameliorates obesity and related phenotypes in mice with diet-induced obesity by modulating energy expenditure and food intake. Diabetologia. 2013, 56, 2297–2307. [Google Scholar] [CrossRef]
  124. Guimarães, E.L.; Best, J.; Dollé, L.; Najimi, M.; Sokal, E.; van Grunsven, L.A. Mitochondrial uncouplers inhibit hepatic stellate cell activation. BMC Gastroenterol. 2012, 12, 68. [Google Scholar] [CrossRef] [PubMed]
  125. Hägg, M.; Berndtsson, M.; Mandic, A.; Zhou, R.; Shoshan, M.C.; Linder, S. Induction of endoplasmic reticulum stress by ellipticine plant alkaloids. Mol. Cancer Ther. 2004, 3, 489–497. [Google Scholar] [CrossRef]
  126. Holtrup, F.; Bauer, A.; Fellenberg, K.; Hilger, R.A.; Wink, M.; Hoheisel, J.D. Microarray analysis of nemorosone-induced cytotoxic effects on pancreatic cancer cells reveals activation of the unfolded protein response (UPR). Br. J. Pharmacol. 2011, 162, 1045–1059. [Google Scholar] [CrossRef] [PubMed]
  127. Ilivicky, J.; Casida, J.E. Uncoupling action of 2,4-dinitrophenols, 2-trifluoromethylbenzimidazoles and certain other pesticide chemicals upon mitochondria from different sources and its relation to toxicity. Biochem. Pharmacol. 1969, 18, 1389–1401. [Google Scholar] [CrossRef] [PubMed]
  128. Kanemoto, N.; Okamoto, T.; Tanabe, K.; Shimada, T.; Minoshima, H.; Hidoh, Y.; Sato, S. Antidiabetic and cardiovascular beneficial effects of a liver-localized mitochondrial uncoupler. Nat. Commun. 2019, 10, 2172. [Google Scholar] [CrossRef] [PubMed]
  129. Nie, L.; Yuan, X.L.; Jiang, K.T.; Jiang, Y.H.; Yuan, J.; Luo, L.; Sun, C. Salsalate Activates Skeletal Muscle Thermogenesis and Protects Mice from High-Fat Diet Induced Metabolic Dysfunction. EBioMedicine. 2017, 23, 136–145. [Google Scholar] [CrossRef] [PubMed]
  130. Pardo-Andreu, G.L.; Nuñez-Figueredo, Y.; Tudella, V.G.; Cuesta-Rubio, O.; Rodrigues, F.P.; Pestana, C.R.; Curti, C. The anti-cancer agent guttiferone-A permeabilizes mitochondrial membrane: Ensuing energetic and oxidative stress implications. Toxicol. Appl. Pharmacol. 2011, 253, 282–289. [Google Scholar] [CrossRef]
  131. Rawling, T.; MacDermott-Opeskin, H.; Roseblade, A.; Pazderka, C.; Clarke, C.; Bourget, K.; Murray, M. Aryl urea substituted fatty acids: A new class of protonophoric mitochondrial uncoupler that utilises a synthetic anion transporter. Chem. Sci. 2020, 11, 12677–12685. [Google Scholar] [CrossRef]
  132. Rokitskaya, T.I.; Khailova, L.S.; Makarenkov, A.V.; Ol'shevskaya, V.A.; Kalinin, V.N.; Antonenko, Y.N. Weak C-H acids as protonophores can carry hydrogen ions through lipid membranes and mitochondria: A case of o-carborane. Phys Chem Chem Phys. 2016, 18, 16476–16482. [Google Scholar] [CrossRef]
  133. Sagara, Y.; Ishige, K.; Tsai, C.; Maher, P. Tyrphostins protect neuronal cells from oxidative stress. J. Biol. Chem. 2002, 277, 36204–36215. [Google Scholar] [CrossRef] [PubMed]
  134. Thomason, L.C.; Court, D.L. Evidence that bacteriophage λ lysogens may induce in response to the proton motive force uncoupler CCCP. FEMS Microbiol. Lett. 2016, 363, fnv244. [Google Scholar] [CrossRef] [PubMed]
  135. Williamson, R.L.; Metcalf, R.L. Salicylanilides: A new group of active uncouplers of oxidative phosphorylation. Science. 1967, 158, 1694–1695. [Google Scholar] [CrossRef] [PubMed]
  136. Wang, J.; He, H.; Xiang, C.; Fan, X.Y.; Yang, L.Y.; Yuan, L.; Jiang, F.L.; Liu, Y. Uncoupling Effect of F16 Is Responsible for Its Mitochondrial Toxicity and Anticancer Activity. Toxicol. Sci. 2018, 161, 431–442. [Google Scholar] [CrossRef]
  137. Marín-Prida, J.; Pardo Andreu, G.L.; Rossignoli, C.P.; Durruthy, M.G.; Rodríguez, E.O.; Reyes, Y.V.; Acosta, R.F.; Uyemura, S.A.; Alberici, L.C. The cytotoxic effects of VE-3N, a novel 1,4-dihydropyridine derivative, involve the mitochondrial bioenergetic disruption via uncoupling mechanisms. Toxicol. Vitr. 2017, 42, 21–30. [Google Scholar] [CrossRef]
  138. Dejonghe, W.; Kuenen, S.; Mylle, E.; Vasileva, M.; Keech, O.; Viotti, C.; Swerts, J.; Fendrych, M.; Ortiz-Morea, F.A.; Mishev, K.; et al. Mitochondrial uncouplers inhibit clathrin-mediated endocytosis largely through cytoplasmic acidification. Nat. Commun. 2016, 7, 11710. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the different types of mitochondrial respiration. (A) Coupled respiration. Electron flow from complex I and complex II through ubiquinone, cytochrome c complex III, and complex IV results in the transport of H+ from the mitochondrial matrix into the intermembrane space, generating a membrane potential. The membrane potential is then used for ATP synthesis by F0F1-ATP synthase. (B) Uncoupled respiration. Reverse leakage of H+ across the membrane occurs either through chemical uncouplers (represented by the classical example of 2,4-dinitrophenol) or through protein uncouplers (represented by UCP1). (C) Noncoupled respiration. This is characteristic of plants and certain sessile animals. There is no generation of a membrane potential because NAD(P)H oxidation is carried out by internal and external NADH dehydrogenases. Alternative oxidase (AOX) serves as a terminal complex utilizing oxygen, and electron flow through these proteins is not coupled to the transport of H+ across the inner mitochondrial membrane.
Figure 1. Schematic representation of the different types of mitochondrial respiration. (A) Coupled respiration. Electron flow from complex I and complex II through ubiquinone, cytochrome c complex III, and complex IV results in the transport of H+ from the mitochondrial matrix into the intermembrane space, generating a membrane potential. The membrane potential is then used for ATP synthesis by F0F1-ATP synthase. (B) Uncoupled respiration. Reverse leakage of H+ across the membrane occurs either through chemical uncouplers (represented by the classical example of 2,4-dinitrophenol) or through protein uncouplers (represented by UCP1). (C) Noncoupled respiration. This is characteristic of plants and certain sessile animals. There is no generation of a membrane potential because NAD(P)H oxidation is carried out by internal and external NADH dehydrogenases. Alternative oxidase (AOX) serves as a terminal complex utilizing oxygen, and electron flow through these proteins is not coupled to the transport of H+ across the inner mitochondrial membrane.
Ijms 24 16491 g001
Figure 2. Green line indicates the number of publications in PubMed database related to “mitochondrial uncouplers” (https://pubmed.ncbi.nlm.nih.gov/) (accessed 22 July 2023). It is noticeable that the number of publications dealing with mitochondrial uncouplers has decreased over the last decade. In contrast, interest in mitochondrial dysfunctions, represented by the blue line indicating the number of publications in the PubMed database on the topic of “mitochondrial dysfunctions”, is growing exponentially.
Figure 2. Green line indicates the number of publications in PubMed database related to “mitochondrial uncouplers” (https://pubmed.ncbi.nlm.nih.gov/) (accessed 22 July 2023). It is noticeable that the number of publications dealing with mitochondrial uncouplers has decreased over the last decade. In contrast, interest in mitochondrial dysfunctions, represented by the blue line indicating the number of publications in the PubMed database on the topic of “mitochondrial dysfunctions”, is growing exponentially.
Ijms 24 16491 g002
Figure 3. Normalized expression of UCP1-5 genes in the endocrine tissues and lymphoid system (A) and in different organs (B). Data were collected from the Human Protein Atlas (https://www.proteinatlas.org/) (accessed 22 July 2023). The number of publications for each query, “UCP1”, “UCP2”, “UCP3”, “UCP4”, “UCP5”, presented in the PubMed database (https://pubmed.ncbi.nlm.nih.gov/) (accessed 22 July 2023) (C).
Figure 3. Normalized expression of UCP1-5 genes in the endocrine tissues and lymphoid system (A) and in different organs (B). Data were collected from the Human Protein Atlas (https://www.proteinatlas.org/) (accessed 22 July 2023). The number of publications for each query, “UCP1”, “UCP2”, “UCP3”, “UCP4”, “UCP5”, presented in the PubMed database (https://pubmed.ncbi.nlm.nih.gov/) (accessed 22 July 2023) (C).
Ijms 24 16491 g003
Figure 4. Normalized expression of UCP1-5 genes in the gastrointestinal tract (A), in the male and female reproductive systems (B), in the different brain compartments (C). Data were obtained from the Human Protein Atlas (https://www.proteinatlas.org/) (accessed 22 July 2023).
Figure 4. Normalized expression of UCP1-5 genes in the gastrointestinal tract (A), in the male and female reproductive systems (B), in the different brain compartments (C). Data were obtained from the Human Protein Atlas (https://www.proteinatlas.org/) (accessed 22 July 2023).
Ijms 24 16491 g004
Table 1. Transgenic models in which UCPs are overexpressed.
Table 1. Transgenic models in which UCPs are overexpressed.
Overexpressed GeneTransgenic StrainsModel of DiseaseEffectReference
UCP1Transgenic strains expressing UCP1 in muscle and heart Body weight was reduced with the same food intake. The decrease in weight mainly occurred in muscle tissue. No changes were observed in cardiac muscle[64]
UCP1Transgenic strains expressing UCP1 in muscle and heartHeart ischemia/reperfusionFunctional recovery on reperfusion was improved [65]
UCP1HSA-mUCP1 mice expressing UCP1 in the skeletal muscles RQ level was increased, indicating an overall increase in glucose oxidation[66]
UCP1HSA-mUCP1 mice expressing UCP1 in the skeletal muscles UCP1 expression in skeletal muscle reduced the risk of reverse electron transfer and the production of reactive oxygen species[67]
UCP1Lou/C rats with UCP1 overexpression in WAT ObesityPrevented body weight gain, decreased fat mass, and improved insulin sensitivity[63]
UCP2Transgenic fly, UAS-hUCP2 Increased hUCP2 expression in the adult nervous system, extended life span[70]
UCP2Transgenic fly, UAS-hUCP2PD modelLess ROS accumulation, heightened resistance to rotenone-induced lethality, and extended life span[71]
UCP2Transgenic mice overexpressing
UCP2 in catecholaminergic neurons (TH-UCP2)
PD modelUpon acute exposure to MPTP, TH-UCP2 mice showed neuroprotection and retention of locomotor functions[60]
UCP2Transgenic mice overexpressing hUCP2ALS modelIncreased survival of sod2−/− mice[72]
UCP2Transgenic mice overexpressing hUCP2ALS modelWorsened mitochondrial dysfunction and accelerated disease progression of sod2−/− mice [73]
UCP2UCP2/3 transgenic overexpressing miceGlobal ischemiaOverexpression of UCP2 protects thalamic neurons following global ischemia[74]
UCP2UCP2/3 transgenic overexpressing mice Focal ischemiaOverexpression of UCP2 blunted the ischemia-induced increase in IL-6 and decrease in Bcl2.[75]
UCP2UCP2/3 transgenic overexpressing miceStroke
TBI
Overexpression of UCP2 enhased of neurological recovery [76]
UCP2Ucp2KIfl/fl miceGlaucomaDecreased glaucomatous cell death[77]
UCP2UCP2/3-overexpressing miceEpileptic seizuresIncreased mitochondrial number and ATP levels with a parallel decrease in free radical-induced damage[78]
UCP2Transgenic mice with targeted expression of UCP2 in the liverAcute liver injuryExpression of UCP2 in mouse liver increases susceptibility to acute liver injury induced by lipopolysaccharide and galactosamine [79]
UCP3Mice overexpressing human UCP-3 in skeletal muscle (UCP-3tg) Mice were hyperphagic but weighed less[80]
UCP3 Mice overexpressing human UCP-3 in skeletal muscle (UCP-3tg) UCP-3tg showed increase in β-oxidation in the MTE-1-dependent manner [81]
UCP3 Mice overexpressing human UCP-3 in skeletal muscle (UCP-3tg) UCP-3tg showed increase in muscle mitochondrial inefficiency and decrease in ATP synthesis [82]
UCP5UCP5-transfected cell lines of heart and kidney The mitochondrial ROS production was decreased [83]
UCP5SH-SY5Y neuroblastoma cells stably overexpressing human UCP5PD modelUCP5 overexpression protected against MPP(+)- and dopamine-induced toxicity[84]
Table 2. List of transgenic models in which component of alternative respiratory pathways was expressed.
Table 2. List of transgenic models in which component of alternative respiratory pathways was expressed.
Model ObjectGeneDonor OrganismComplex BypassROS EffectModel of DiseaseReference
Cell lines
Flp-InTM T-RExTM-293 cellsAOXC. intestinalisCIV dysfunction.
Cyanide-induced inhibition
Decrease in antimycin-induced superoxide overproduction [97]
COX10-depleted HEK-293-derived AOX-transgenic cells from Hakkaart et al., 2005AOXC. intestinalisCIV dysfunction.
shRNA targeted
against COX10
[92]
COX-defective fibroblastsAOXC. intestinalisCIV dysfunction.
Deleterious COX15 gene mutation
Reduction in the superoxide production in COX15 cells in the presence of antimycinHypertrophic cardiomyopathy[92]
HEK293 Flp-In cellsAOXC. intestinalis Alzheimer’s disease[94]
Complex I defective fibroblastsNDH-2Arabidopsis thalianaCI dysfunction.
CI-defective fibroblasts
The normalization
of SOD activity
[106]
Drosophila
w1118 DrosophilaAOXC. intestinalisCIV dysfunctions.
Partial knockdown of COXVIc and complex IV assembly factor Surf1
Leigh syndrome[95]
w1118 DrosophilaAOXC. intestinalis Reduction in the ROS productionParkinson’s disease[95]
w1118 Drosophila with knockdown of different complex IV subunitsAOXC. intestinalisCIV dysfunctions.
Knockdown of different CIV subunits
w1118 DrosophilaAOXC. intestinalis AOX abrogates induction of oxidative stress markers in a Drosophila AD modelAlzheimer’s disease[94]
w1118 DrosophilaNdi1S. cerevisiaeCI dysfunction.
Rotenone-induced inhibition,
paraquat-induced inhibition
Mitigation of mitochondrial ROS production, oxidative damage [105]
w1118 DrosophilaNdi1S. cerevisiaeCI dysfunction.
Rotenone-induced inhibition
Ndi1 expression in neurons, reducing ROS
levels
[102]
UAS-dCIA30 DrosophilaNdi1S. cerevisiaeCI dysfunction.
Reduced expression of CI assembly factor
[103]
w1118 DrosophilaNdi1Saccharomyces cerevisiaeCI dysfunction.
Rotenone-induced inhibition
Reduction in whole tissue ROS levels [104]
Drosophila lines 24,861 and 24,871NdxC. intestinalis Increased resistance to menadione-induced ROS production [100]
Mice
Mice (strain not specified)AOXC. intestinalisCIV dysfunction.
Cyanide-induced inhibition
Decrease in antimycin- and cyanide-induced superoxide overproduction [93]
AOXRosa26 mice,
C57Bl/6J strain background
AOXC. intestinalisCIV dysfunction.
Cyanide-induced inhibition
Azide-induced inhibition.
CIII dysfunction.
Antimicyn-induced inhibition
Decrease in H2O2 production in succinate-supported mitochondria [99]
AOXRosa26 mice,
C57Bl/6J strain background
AOXC. intestinalisCII dysfunction.
Cigarette smoke condensate-induced inhibition.
CIV dysfunction.
Cigarette smoke condensate-induced inhibition
Decreases
superoxide production
[96]
(cIII)-deficient Bcs1lp.S78G knock-in mice AOX backcross with transgenic miceAOXC. intestinalisCIII dysfunction.
(CIII)-deficient Bcs1lp.S78G knock-in mice
Lethal mitochondrial cardiomyopathy[98]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gureev, A.P.; Alimova, A.A.; Silachev, D.N.; Plotnikov, E.Y. Noncoupled Mitochondrial Respiration as Therapeutic Approach for the Treatment of Metabolic Diseases: Focus on Transgenic Animal Models. Int. J. Mol. Sci. 2023, 24, 16491. https://doi.org/10.3390/ijms242216491

AMA Style

Gureev AP, Alimova AA, Silachev DN, Plotnikov EY. Noncoupled Mitochondrial Respiration as Therapeutic Approach for the Treatment of Metabolic Diseases: Focus on Transgenic Animal Models. International Journal of Molecular Sciences. 2023; 24(22):16491. https://doi.org/10.3390/ijms242216491

Chicago/Turabian Style

Gureev, Artem P., Alina A. Alimova, Denis N. Silachev, and Egor Y. Plotnikov. 2023. "Noncoupled Mitochondrial Respiration as Therapeutic Approach for the Treatment of Metabolic Diseases: Focus on Transgenic Animal Models" International Journal of Molecular Sciences 24, no. 22: 16491. https://doi.org/10.3390/ijms242216491

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