Next Article in Journal
In Vitro Antioxidant Properties, Glucose-Diffusion Effects, α-Amylase Inhibitory Activity, and Antidiabetogenic Effects of C. Europaea Extracts in Experimental Animals
Next Article in Special Issue
Role of Oxidative Stress in Diabetic Cardiomyopathy
Previous Article in Journal
Glutathione Peroxidase in Stable Chronic Obstructive Pulmonary Disease: A Systematic Review and Meta-analysis
Previous Article in Special Issue
Melatonin Enhances the Mitochondrial Functionality of Brown Adipose Tissue in Obese—Diabetic Rats
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mitochondrial Uncoupling Proteins (UCPs) as Key Modulators of ROS Homeostasis: A Crosstalk between Diabesity and Male Infertility?

by
Bruno S. Monteiro
1,†,
Laís Freire-Brito
1,†,
David F. Carrageta
1,
Pedro F. Oliveira
2 and
Marco G. Alves
1,*
1
Department of Anatomy and Unit for Multidisciplinary Research in Biomedicine (UMIB), Institute of Biomedical Sciences Abel Salazar (ICBAS), University of Porto, 4050-313 Porto, Portugal
2
Química Orgânica, Produtos Naturais e Agroalimentares (QOPNA) & Laboratório Associado para a Química Verde—Rede de Química e Tecnologia (LAQV-REQUIMTE), Department of Chemistry, University of Aveiro, 3810-193 Aveiro, Portugal
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Antioxidants 2021, 10(11), 1746; https://doi.org/10.3390/antiox10111746
Submission received: 29 September 2021 / Revised: 27 October 2021 / Accepted: 28 October 2021 / Published: 30 October 2021

Abstract

:
Uncoupling proteins (UCPs) are transmembrane proteins members of the mitochondrial anion transporter family present in the mitochondrial inner membrane. Currently, six homologs have been identified (UCP1-6) in mammals, with ubiquitous tissue distribution and multiple physiological functions. UCPs are regulators of key events for cellular bioenergetic metabolism, such as membrane potential, metabolic efficiency, and energy dissipation also functioning as pivotal modulators of ROS production and general cellular redox state. UCPs can act as proton channels, leading to proton re-entry the mitochondrial matrix from the intermembrane space and thus collapsing the proton gradient and decreasing the membrane potential. Each homolog exhibits its specific functions, from thermogenesis to regulation of ROS production. The expression and function of UCPs are intimately linked to diabesity, with their dysregulation/dysfunction not only associated to diabesity onset, but also by exacerbating oxidative stress-related damage. Male infertility is one of the most overlooked diabesity-related comorbidities, where high oxidative stress takes a major role. In this review, we discuss in detail the expression and function of the different UCP homologs. In addition, the role of UCPs as key regulators of ROS production and redox homeostasis, as well as their influence on the pathophysiology of diabesity and potential role on diabesity-induced male infertility is debated.

1. Introduction

Obesity and diabetes mellitus are perilous health issues that also constitute a huge economic burden worldwide. It is currently estimated that 9.3% of the world population suffers from diabetes mellitus, whereas 39% of adults aged 18 years and over are overweight, and 13% are obese [1,2]. Obesity and diabetes mellitus are intrinsically related to each other, leading to the term “diabesity”. Both conditions are regarded as metabolic diseases due to the impairment of metabolism and augmented levels of reactive oxygen species (ROS), that result in oxidative stress [3]. The biggest source of ROS is the mitochondrial oxidative phosphorylation, which is critical to produce the high quantities of ATP essential for cell survival. However, oxidative phosphorylation is not perfectly coupled to ATP synthesis and some of the energy present in the electrochemical force is dissipated due to the re-entry of protons in the mitochondrial matrix independently of ATP synthase. Indeed, some electrons escape to oxygen (0.2–2%), essentially in complexes I and III, leading to the appearance of reactive oxygen species (ROS) [4,5,6,7]. Although ROS, at physiological levels, are important cell signaling agents, elevated concentrations of these radicals are extremely noxious to cells. Therefore, it is imperative to control their generation [4,5,6,8]. The movement of protons to the mitochondrial matrix without passing through the ATP synthase results in an enormous decrease in ROS production while having a minimal effect on ATP synthesis (a reduced attenuation in the potential difference in the inner membrane of the mitochondria can restrain the formation of H2O2 by 70%) [7,9,10]. Proton transport to the mitochondrial matrix independent of ATP synthase occurs through two processes: basal proton leak and inducible proton leak (Figure 1). Basal proton leak is not finely regulated and depends only on fatty acid composition of the inner mitochondrial membrane and on the abundance of adenine nucleotide translocase (ANT) [6]. On the other hand, inducible proton leak is finely regulated, with uncoupling proteins (UCPs) playing a crucial role [6].
Herein we discuss in detail the expression and function of the different UCP homologs in human, mouse, and rat. In addition, the role of UCPs as key regulators of ROS production and redox homeostasis and their potential role on the pathophysiology of diabesity and on diabesity-induced male infertility will be discussed.

2. Mitochondrial Uncoupling Proteins (UCPs) Expression and Function

UCPs are integral transmembrane proteins of the inner mitochondrial membrane and members of the mitochondrial anionic transporter family SLC25 [7,9,11,12]. UCPs are constituted by three repetition domains, each composed of two α-helix regions [7,12,13,14,15]. The carboxy and amino-terminal regions are on the intermembrane space while the α-helix regions are linked by long loops that reside on the matrix side of mitochondria (Figure 2A) [7,16,17]. There are two main hypotheses concerning the proton transport mechanism by UCPs: the flip-flop model and the cofactor model (Figure 2B) [18,19]. The flip-flop model occurs in the presence of high concentrations of fatty acids, in which the protonated form of the fatty acid freely crosses the membrane towards the matrix and dissociates the proton in the matrix due to the pH difference (higher in the matrix). The anionic fatty acid returns to the intermembrane space with the help of the UCP, completing the cycle. In the cofactor model, protons can be transported independently or, at a higher rate, linked to the carboxyl group of fatty acids, which operate as prosthetic groups.
UCPs play a central role in regulating the potential of mitochondrial membrane, exhibiting several distinct functions from thermogenesis (dissipating energy in the form of heat) to oxidative phosphorylation or ROS levels regulation [13,18,20]. Currently, six homologs of mammalian UCPs (UCP1-6) have been identified (Table 1), which will be discussed in detail on the following topics.

2.1. UCP1

UCP1 is predominantly expressed in brown adipose tissue (BAT), where it constitutes up to 10% of the mitochondrial proteins in this tissue [13,79]. Humans present UCP1 expression almost exclusively in brown adipocytes [21], however, it has also been found to be expressed in white adipose tissue (WAT), on brite or beige adipocytes [22,80] and skin [23]. This expression pattern is shared by mice and rats, where UCP1 expression is almost confined to BAT [24,27], with some studies describing its presence in WAT (including epididymal WAT) [25,28], and mice adrenal gland [26]. Moreover, UCP1 has been identified in smooth muscle cells from mice (including those from the female and male reproductive tracts) [81], and mouse and rat thymocytes [82]. However, these findings have been disputed on later publications [38,83]. Furthermore, the existence of UCP1 mRNA in human ductus deferens, testis, epididymis, seminal vesicle, and prostate [84] has been predicted.
The main function attributed to UCP1 is the production of heat through thermogenesis [85]. BAT mainly consists of brown adipocytes, cells specialized in thermogenesis that have high expression of thermogenic genes, including UCP1. Exposure to cold leads to adrenergic activation of brown adipocytes and a higher expression of Ucp1 [86]. As a result, there is an increase in the number and size of mitochondria and the rate of lipolysis is also increased, leading to a higher content of free fatty acids [87]. These free fatty acids activate UCP1, instigating the uncoupling of the respiratory chain and the production of heat [20]. In addition to the cold exposure and adrenergic regulation, including triiodothyronine (T3) and norepinephrine [88], UCP1 activity and expression is also stimulated by superoxide [89], and compounds with a 2-alkenal functional group, including trans-retinoic acid, trans-retinal, trans-2-nonenal, and retinoic acid through a non-adrenergic pathway [90]. In addition, UCP1 also responds to the nutritional status of the organism. Gong et al. reported that Ucp1 mRNA levels drop with starvation and recover after feeding [91]. UCP1 is also regulated by nucleotides. Interestingly, the binding site of fatty acids to the protein is different from that of nucleotides. Through a FTIR spectra analysis, Chan et al. observed that the addition of nucleotides causes considerable changes in the secondary, and probably tertiary, structure of UCP1 of hamster BAT [92]. In contrast, no secondary or tertiary structure changes were detected upon fatty acid binding.

2.2. UCP2

UCP2, which is 59% identical to UCP1, is probably the most widely expressed UCP homolog [35,93]. Unlike UCP1, UCP2 is presumably ubiquitously expressed throughout the body. In humans, UCP2 expression has been detected in kidneys [29], placenta [31], spermatozoa [32], and brain [35]. In skin [23], lungs [35], WAT, BAT, skeletal muscle [33], and pancreatic β-cells [34], only the mRNA expression has been reported in humans. Widespread UCP2 expression is also observed in rodents, where in mice expression has been identified in thymus [36], kidney [37], mammary gland, stomach [38], macrophages [13], B lymphocytes, T lymphocytes, dendritic cells, neutrophils [45], spleen [39], heart [41], lungs [46], testis [42], and pancreatic α-cells [43], while Ucp2 mRNA was reported in pancreatic β-cells [34], skin [37], female reproductive tract [38], brain [40], WAT, BAT, skeletal muscle [35], and liver [44]. In rats, UCP2 presence has been described in kidneys [47], neurons [50], microglia [49], heart [51], testis [53], and pancreatic β-cells [54], and its mRNA in liver [48], skeletal muscle [52], WAT, BAT [33], lungs [55], spleen [55], and thymus [56]. Additionally, the existence of UCP2 mRNA has been predicted in human ductus deferens, testis, epididymis, seminal vesicle, and prostate [84].
The main function attributed to UCP2 comprises the regulation of ROS production and protection against oxidative damage [94,95]. Thus, UCP2 has been studied for its protective effects on several diseases associated with increased oxidative stress, including Alzheimer’s disease, Parkinson’s disease, arteriosclerosis, diabetes mellitus [96,97,98,99]. In addition, UCP2 is reported to have a protective role against stroke and trauma [100], ethanol intoxication [101], brain and cardiac ischemia [102,103], and a regulatory role on inflammation [98,104] and glucose-stimulated insulin secretion [105]. Some studies also report that UCP2 may transport inorganic anions such as aspartate, malate, phosphate, or oxaloacetate [106]. However, it is not known whether the anion transport can be part of the uncoupling cycle or if it is just a side reaction [96]. Interestingly, UCP2 appears to mediate the transport of protons only when specifically activated [96]. UCP2 is activated by high ROS levels, fatty acids, and molecules containing the 2-alkenal functional group [89,96,107]. Ucp2 mRNA expression is reported to be regulated by glutamine and leptin [108,109]. On the other hand, UCP2 is inhibited by purine nucleotides [89,107].

2.3. UCP3

UCP3 shares an 73% identity with UCP2 and 57% with UCP1 [33,60]. In humans, UCP3 is predominantly expressed in skeletal muscle [59] and, to a lesser extent, skin [23], and pancreatic β-cells [58]. An equivalent UCP3 expression pattern is observed in mice which is preferentially expressed in skeletal muscle and BAT [60], and, in a smaller degree, in heart, WAT [61], spleen, and thymus [62]. Likewise, rats have predominant UCP3 expression in skeletal muscle [52], and smaller amounts in heart [64], spleen, and thymus [57]. mRNA has been detected in human heart [33], spleen [57], thymus [57], keratinocytes [23], thyroid, and bone marrow [60], in mice brain, kidney, colon, liver [60], and circulating leucocytes [63], and in rat BAT [52], WAT and kidney [33]. The existence of UCP3 mRNA was predicted in human ductus deferens, testis, epididymis, seminal vesicle, and prostate [84].
Similar to UCP2, UCP3 also plays a role on the regulation of ROS production and protection against oxidative damage [110,111,112]. In addition, UCP3 is hypothesized to play a protective role against cardiac ischemia [103]. As UCP2, UCP3 is positively regulated by ROS levels [113], fatty acids, molecules containing 2-alkenal functional group [96], and leptin [114], while it is inhibited by purine nucleotides [107]. However, UCP3 also has several resemblances with UCP1. Although it is not normally thermogenic or involved in the adaptive response to cold, UCP3 may be significantly thermogenic under specific conditions, including glucocorticoids or T3 stimulation, and cold exposure [91,96,115]. In addition, UCP3 mRNA is also regulated by fasting in both human and rat [116,117].

2.4. UCP4

UCP4, along with UCP5 and UCP6, are less identical to the other three homologs [15]. UCP4 is predominantly expressed in the central nervous system either from human [7], rat [65] or mouse [66], but its expression has also been detected in other tissues and cells such as Merkel cells, modiolus ear, organ of Corti [67], and chondrocytes [65] of rat, and spinal cord [66], and mast cells [70] of mouse. In human cartilage [65], rat vestibular ganglion [68], heart, lungs, skeletal muscle, kidney, and liver [55] and mouse spiral ganglion, vestibular ganglion, kidney [71], heart, liver, skeletal muscle, WAT [55], and BAT [72], only the mRNA expression was reported. Furthermore, the existence of UCP4 mRNA in human ductus deferens, testis, epididymis, seminal vesicle, and prostate [84] has been predicted.
The activation of UCP4 attenuates mitochondrial ROS levels [9]. Since the apoptosis of neurons is related to the action of ROS, UCP4 is hypothesized to play neuroprotective role in Alzheimer’s and Parkinson’s disease [118,119]. This uncoupling protein may also be responsible for the observed mitochondrial proton leakage in the brain [120], acting as a thermoregulator in response to cold [69]. Another suggested role for UCP4 in neurons is the regulation of mitochondrial Ca2+ signalling [9]. Analogously to the other homologs, UCP4 is inhibited by purine nucleotides and activated by ROS and fatty acids [19]. Calorie restriction and cold induce Ucp4 mRNA expression in the rat brain as well [20].

2.5. UCP5

UCP5 is mainly present in human [73], rat [75] and mouse [74] central nervous system. It is also expressed in other tissues, organs and cells as kidney, heart, lungs, stomach, liver, spleen, skeletal muscle, spinal cord, uterus [73], pituitary, and testis [14] in humans, heart, lungs, adrenals, kidney, gonadal WAT, ovary [14], skeletal muscle, and liver [55] in rat, and spiral ganglion, vestibular ganglion [71], hepatocytes [44], BAT, spleen, intestine, lungs, testis, uterus, and periovarian [14] fat in mouse. Its protein has been detected in mouse brain, heart, kidney, skeletal muscle, WAT, and spinal cord [74]. Additionally, the existence of UCP5 mRNA has been predicted in human ductus deferens, epididymis, seminal vesicle, and prostate [84].
As UCP4, UCP5 may have a neuroprotective effect against damage related to oxidative stress [74]. This can have implications in several pathological conditions, including Alzheimer’s disease [118], and cortical spreading depression [121]. It is also hypothesized that UCP5 is involved in metabolic rate changes associated with starvation and a high-fat diet [73]. In addition, Ucp5 mRNA levels decrease during brain hypoxia and increase under hyperoxic conditions [122]. UCP5 is also hypothesized to play a role in thermoregulation induced by cold exposure [73]. Interestingly, UCP5 transports mainly sulphate and thiosulfate [78]. These compounds are products of the degradation of H2S, so UCP5 is a potential regulator of the mitochondrial levels of this important signalling molecule. UCP5 also efficiently transports other inorganic anions, including sulphite, phosphate and, to a lesser extent, maleate, oxalate, malonate, malate, citramalate, aspartate, and glutamate [78].

2.6. UCP6

Ucp6 mRNA was found to be mainly expressed in mouse kidneys [30]. It is also considerably expressed in the testis of mice. Its expression has also been detected in other mouse tissues and organs such as WAT, BAT, brain, heart, muscle, liver, lungs, and spleen. The existence of UCP6 mRNA was predicted in human ductus deferens, testis, epididymis, seminal vesicle, and prostate [84].
UCP6 is upregulated in response to oxidative stress [30]. Although UCP6 seems to have no uncoupling capacity [30], its antioxidant role is hypothesized to be due to the regulation of H2S concentration, which at low concentrations inhibits ROS production in mitochondria [123]. In fact, Ucp6 was found to be upregulated during cancer [124] or the regenerative phase after renal tubular damage [30], which are conditions characterized by high oxidative damage. Moreover, it is assumed that increased Ucp6 expression is associated with increased mitochondrial activity [30]. Similar to UCP5, UCP6 primarily transports sulphate and thiosulfate, which supports its role as regulator of H2S levels [78]. UCP6 is also able to transport the same inorganic anions as UCP5, except for glutamate [78].

3. UCPs Are Key Regulators of ROS Production and Redox Homeostasis

ROS are an essential part of the cell function, participating in numerous processes such as immune response, cellular proliferation and differentiation, angiogenesis, aging, and several signaling pathways. They are a natural byproduct of the mitochondria’s oxidative phosphorylation, arising when a single unpaired electron reduces O2 at the oxidative chain level, which results in formation of the superoxide anion. However, an unregulated increase in the production of these free oxygen radicals renders the cellular antioxidant system incapable of mitigating their harmful effects, causing a shift in the homeostasis of the redox balance towards oxidative stress, a damaging state that participates on the pathophysiology of several diseases and capable of inducing macromolecular damage towards apoptosis/necrosis [125,126]. It has been proposed that mild uncoupling of the mitochondrial oxidative phosphorylation system might prevent the oversupply of electrons to the various protein complexes of the electron transport chain, diminishing the possibility of electron leaks and interaction with oxygen. Hence, controlled proton conductance via the mitochondrial inner membrane has been suggested as a key regulator of ROS production, through membrane potential modulation [126,127,128,129].
All UCPs homologs are well known for regulating ROS homeostasis, although UCP1 role as an antioxidant remains debatable. Given the extensively described uncoupling activity of UCP1, it would be expected that this protein has an active role in the control of mitochondrial ROS production. However, there are some conflicting data. For instance, Dlasková et al. observed that BAT mitochondria from both wild-type C57BL/6J mice treated with guanosine diphosphate (GDP), an UCP1 inhibitor, and Ucp1 knockout mice exhibited greater H2O2 production and higher mitochondrial membrane potential [130]. In accordance with these results, ablation of Ucp1 in the BAT mitochondria of mice exposed to cold stress leads to higher superoxide production, and subsequently oxidative stress, when compared to wild-type mice [131,132]. Besides, ROS have been shown to be capable of modulating UCP1 activity via sulfonylation of the Cys253 residue, increasing thermogenesis [133]. Conversely, Shabalina et al. reported that UCP1-dependent proton leak is not affected by the oxidative stress byproduct 4-hydroxy-2-nonenal (HNE), and Ucp1 deletion in mice does not predispose them to higher oxidative damage [134]. In a later study, also using mitochondria from BAT of wild-type and Ucp1 knockout mice, they determined that UCP1 activity bears no weight in the reduction of ROS emission, except when succinate was endogenously added. Interestingly, the authors also inferred that low membrane potential may not necessarily correlate with lower oxidative stress, challenging the claim that mild uncoupling protects mitochondria against oxidative damage [135].
UCP2 and UCP3 role as modulators of ROS production is extensively described. UCP2 and UCP3 do not display a clear uncoupling activity, so it is unlikely that they regulate superoxide production in a mitochondrial uncoupling-dependent manner. The redox balance mediated by UCP2 was first demonstrated by Nègre-Salvayre et al., where it was shown that mitochondria fractions containing high levels of UCP2 treated with the inhibitor GDP presented a significant increase in H2O2 generation, whereas the treatment in mitochondria with low expression of UCP2 led to no changes in the oxidative state [56]. Other studies focusing on the overexpression of UCP2, using an adenoviral vector demonstrated that UCP2 plays a protective role against oxidative stress, shielding both cardiomyocytes and endothelial cells from ROS-induced cell death [51,97]. UCP2 neuroprotective role against oxidative stress has also been reported [136,137,138]. Compelling evidence highlights that ablation of Ucp2 in multiple mice strains induces chronic high levels of ROS in pancreatic cells [139,140,141]. Elevated ROS levels in Ucp2-deficient mice also results in higher levels of oxidative stress markers in the liver and delayed hepatic regeneration [95], as well as aggravated atherosclerosis injury [142]. On another note, Arsenijevic et al. observed that Ucp2 knockout mice infected with Toxoplasma gondii presented a stronger infection fighting capacity, when compared to wild type, due to higher ROS emission in macrophages [13]. Due to its ROS suppressing activity, targeted UCP2 inhibition has been proposed as a potential novel treatment for cancer [143,144,145].
Akin to UCP2, it was hypothesized that the overproduction of ROS activates UCP3 through a negative feedback mechanism, diminishing oxidative damage [146]. The presence of superoxide from both endogenous [111,146,147] and exogenous sources [89,148], along with byproducts of lipid peroxidation [149,150] have been shown to induce UCP3-mediated proton leak. In addition, Ucp3 knockout mice display higher rates of ROS production and are more vulnerable to oxidative damage, including increased vulnerability to ischemia-reperfusion injury [77,110,151,152]. Toime and Brand studied isolated energized skeletal muscle mitochondria from wild-type and Ucp3 knockout mice and reported that UCP3 decreases the rate of ROS production [151]. Interestingly, these authors treated the isolated mitochondria with FCCP, a well-known chemical uncoupler, which mirrored the UCP3 antioxidant activity. However, inhibition with GDP was ineffective at reducing UCP3-dependent ROS suppression and no differences in the mitochondrial membrane potential were observed, which suggests that UCP3 can modulate ROS production through a membrane potential-independent mechanism. In support to these results, MacLellan et al. demonstrated that moderate overexpression of UCP3 in L6 myocytes does not interfere with basal cellular oxygen consumption or membrane potential despite potentiating its antioxidant activity [153]. UCP3 is also reported to be upregulated in myocytes upon muscle/fiber contraction, suggesting a role on protection against physical exercise-related ROS production [146,154]. Overexpression of UCP3 has also been associated with a protective effect against age-related rise in mitochondrial oxidative damage [155]. On a final note, UCP3-mediated translocation of lipid peroxides to the extramitochondrial medium has also been reported to be a competent strategy in minimizing oxidative stress [156].
Although fewer studies focus on UCP4-6, compelling evidence highlights their role on the regulation of ROS production. Increased levels of Ucp4 and Ucp5 mRNA and respective proteins were found to protect neurons against oxidative stress damage [118]. In addition, other studies shown that UCP5 expression decreased the production of superoxide radicals in mouse neurons [74], whereas it protected cells against oxidative stress due to hyperoxic conditions [122]. UCP6 represents an intriguing case. Although UCP6 does not display an uncoupling function, this homolog can be upregulated in response to oxidative stress [30]. It was hypothesized that its antioxidant effect must be due to the regulation of H2S levels. Although H2S exerts cytotoxic effects when in high concentrations, at low concentrations it exhibits beneficial and protective effects [123]. One of the beneficial effects of low H2S concentrations is the inhibition of ROS production in mitochondria through the stimulation of cellular antioxidants [78,123]. Hence, UCP6 seems to contribute to the maintenance of low levels of ROS by keeping the concentration of H2S low.

4. UCPs Dysregulation Leads to Diabesity

UCPs have a significant influence in mitochondrial bioenergetics, thermogenesis, and control of ROS production, which are all key processes known to progressively become impaired with the onset of metabolic disorders (including obesity and diabetes mellitus). Obesity and diabetes mellitus are closely linked, where a prolonged and persistent positive intake of energy lead to chronic hyperglycemia and subsequent insulin resistance [157]. Even though physical activity accounts for a large percentage of total energy expenditure, resting metabolic rate (RMR) and adaptive thermogenesis are fundamental mechanisms [158,159]. For instance, it is estimated that proton leak through the inner membrane of mitochondria is responsible for 20–30% of the RMR in rats [160]. In fact, alterations in proton conductance and UCPs expression have been implicated in the development of obesity and are considered potential targets for its treatment [161].
UCP1-mediated thermogenesis can increase 12–20% the total daily energy expenditure [162]. However, early studies in Ucp1 knockout mice were incapable of establish a clear connection between UCP1 deficiency and obesity. Enerbäck et al. were the first to report the absence of an obesogenic effect, as Ucp1-deficient mice fed a standard or high-fat diet did not become obese [163]. These findings were later supported by other studies [164,165]. Notably, Feldmann et al. suggested that the housing temperature at which the mice were kept might have confounded some of the results from these studies [166]. Mice housed at a standard temperature (18–22 °C) are under chronic cold stress, which requires an increase in their metabolic rate of 50% to 60% just to sustain body temperature. This increase in energy expenditure might have masked any possible effect of UCP1 in the energy metabolism. In this same study, the authors demonstrated that when Ucp1 knockout mice were housed at thermoneutral temperature (i.e., 30 °C), they do indeed develop obesity even when fed standard diet. In accordance with these results, more recent studies have also described the obesogenic role of Ucp1 ablation in mice kept at ~30 °C [167,168], even for notoriously obesity resistant 129 mice strain [169]. Interestingly, it was observed that this strain had markedly higher metabolic rate with increased energy expenditure when compared to other obesity prone strain, the C57BL/6 mice. Raised RMR was explained by higher expression of UCP1 and uncoupling in mitochondria from muscle. After histological analysis it was possible to determine that higher expression of UCP1 correlates with the presence of intramuscular depots of BAT [170]. In addition, Winn et al. reported that Ucp1 knockout mice fed Western diet and kept at 25 °C developed insulin tolerance, although no significant differences in body weight, visceral adiposity, and energy expenditure in comparison with wild-type mice were observed [171]. Loss of Ucp1 also causes whitening of BAT [171], decreasing its protective role against the development of insulin resistance, type 2 diabetes, and obesity [172,173]. In fact, an UCP1-dependent role for BAT in glucose homeostasis has been previously described. Ucp1-deficient mice lack BAT glucose uptake after norepinephrine administration, suggesting that sympathetic stimulation leads to UCP1 mediated glucose disposal by BAT [174]. Notably, BAT transplant to Ucp1 knockout mice was able to eliminate glucose intolerance [175]. Furthermore, stimulation of UCP1 dependent BAT thermogenesis, either through cold or diet, has been shown to increase energy expenditure via catabolic acceleration. When active, BAT oxidizes significant amounts of metabolic substrates such as stored fatty acids, as well as plasma glucose and triglycerides. Thus, through reduction of the circulating glucose and lipids in order to supply the energy costly thermogenesis, BAT is a tissue of great interest to treat metabolic disorders such as obesity and type 2 diabetes mellitus [87,167,168,171,176]. In accordance with these results, Vázquez et al. reported that Zücker diabetic fatty rats, an experimental model for obesity and type 2 diabetes mellitus, exhibited significantly lower levels of UCP1 protein and mRNA expression in BAT when compared to their lean littermates. Interestingly, treatment with melatonin, which has been described has having an anti-obesity effect, was able to restore not only UCP1 expression but also its functionality in obese rats. This increase in UCP1 expression and function resulted in higher thermogenic activity, which was linked with enhanced energy expenditure and overall improved metabolic state of treated rats [177]. Further evidence for the UCP1 antidiabetic and anti-obesity effect comes from studies in transgenic mice models overexpressing UCP1. Transgenic mice with upregulation of UCP1 in WAT showed resistance to diet-induced obesity even when fed a high-fat diet [178,179]. Moreover, several studies have linked ectopic overexpression of UCP1 in the skeletal muscle with higher energy expenditure, lean phenotype, lower levels of glucose, insulin, cholesterol, and adiposity, enhanced lipid metabolism and glucose transport, and improved responsiveness to insulin [180,181,182,183,184].
UCP2 has also been implicated in the pathogenesis of diabesity, almost since its discovery. A study by Zhang et al. observed that Ucp2 knockout mice had improved glucose-stimulated insulin secretion (GSIS), indicating that UCP2 activity negatively regulates insulin secretion [105]. The authors also reported that UCP2 was significantly upregulated in isolated pancreatic islets of a model of obesity-induced diabetic mice (ob/ob mice), when compared to control. Interestingly, ob/ob mice lacking Ucp2 had enhanced β-cell responsiveness to glucose, decreased levels of glycemia and restored first phase insulin secretion. Overall, these results seem to indicate that obesity-dependent induction of UCP2 in pancreatic β-cells has a pathogenic effect towards the development of insulin resistance and in turn diabetes mellitus. In accordance with this hypothesis, overexpression of UCP2 in isolated β-cells was reported to impair GSIS [34,54,185], while inhibition with genipin, an UCP2 inhibitor, improved glucose secretion [186]. Desouza et al. observed that partial inhibition of UCP2 with an antisense oligonucleotide in two mice models of type 2 diabetes mellitus, ob/ob mice and diet-induced obese and diabetic mice, led not only to improvements in insulin secretion in β-cell but also enhanced insulin action in peripheral tissues (eg. adipose tissue) [187]. Saleh et al. shown that acute knockdown of UCP2 in isolated pancreatic islets from lean mice increased in vitro insulin secretion but failed to improve GSIS in isolated pancreatic islets from ob/ob mice [188]. Nevertheless, dissenting results have highlighted the possibility of the confounding effects of genetic background and systemic absence of protein activity in studies with Ucp2-ablated mice (detailedly reviewed in [189]). In addition, early reports negatively correlated UCP2 induction with β-cell ATP levels [34,54,105,186,188]. A study by Affourtit and Brand using acute knockdown of Ucp2 in clonal β-cells, INS-1E cells, reported that UCP2 is responsible for the high mitochondrial proton conductance observed in this model [190]. ATP is a powerful stimulus for insulin secretion, which led the authors to speculate that UCP2 modulates GSIS in pancreatic islets through the coupling efficiency of oxidative phosphorylation. In fact, this hypothesis is shared by several other authors [54,105,187]. Growing evidence also has revealed the pivotal role of ROS production in β-cell function. Glucose-dependent ROS production and overall changes in the redox status of pancreatic islets have been shown to stimulate insulin production and secretion, which in turn can be suppressed through antioxidative activity. Therefore, a role for ROS in controlling β-cell function under pathological conditions, such as obesity and diabetes mellitus, has been proposed [191,192,193]. In support, several studies focused on the potential role of UCP2 in GSIS through the control of mitochondrial ROS production and protection against oxidative stress. For instance, Ucp2 ablation in INS-1E insulinoma cells resulted in improved GSIS, which can be mimicked using the cell-permeative antioxidant MnTMPyP [194]. A study using a refined model of β-cell specific Ucp2 knockout reported that Ucp2-ablated pancreatic islets had increased ROS levels correlating with increased GSIS [140]. Interestingly, while no changes in the uncoupling respiration rates and ATP/ADP ratio were observed, Ucp2-deficient islets exhibit mildly increased mitochondrial membrane potential [140]. Taken together, these findings suggest that Ucp2 knockout increases intracellular ROS and thus stimulate insulin secretion. Thus, UCP2 might be seen as a double-edged sword: decreasing its activity seems to improve GSIS, however it leaves insulin-secretion β-cell susceptible to oxidative damage, which suggests a finely regulated activity.
UCP3 plays an active role in ROS production and fatty acid oxidation, which makes it a candidate for better understanding the pathophysiology of metabolic disorders such as obesity and type 2 diabetes mellitus. Experimental evidence from multiple studies have proposed UCP3 as a modulator of energy metabolism and a possible protective role in obesity. For instance, transgenic mice overexpressing UCP3 fed a fatty diet exhibit lower body weight, decreased adipose tissue, and triglyceride accumulation [195,196,197]. Furthermore, moderate upregulation of UCP3 can mimic the beneficial consequences of physical exercise, promoting fatty acid oxidation, increasing energy expenditure and consequentially increase weight loss [198]. In contrast, studies using Ucp3−/− mice failed to demonstrate an obesogenic effect for Ucp3 ablation [77,199]. More surprisingly, a recent study by Lomax et al. linked Ucp3 deficiency with protection from diet-induced obesity [200]. However, it is important to highlight that housing of animals in conditions of chronic thermal stress may have conflicting results. In regard to diabetes mellitus, overexpression of UCP3 improves glucose tolerance and protected the animals against fat-induced insulin resistance [201,202]. Studies focusing in Ucp3 ablation have conflicting evidence as Ucp3−/− animals fed a fatty diet exhibited either enhanced [200], decrease [203] or unaltered [77] insulin resistance. In humans, high glucose exposure decreased UCP3 protein expression in isolated islets, whereas overexpression of UCP3, contrary to UCP2, improves GSIS [58]. Concurrently, clinical studies have reported lower expression of UCP3 mRNA and protein in prediabetic and diabetic individuals, when compared to healthy individuals [204,205,206]. Interestingly, both 8-week treatment with rosiglitazone and physical exercise training were able to restore UCP3 levels in diabetic patients [205].

5. The Potential Role of UCPs on Diabesity-Induced Male Infertility

The link between oxidative stress and male factor infertility is well documented since it was first described by MacLeod in 1943, where a connection between H2O2 production and loss of spermatozoa mobility was observed [207]. In fact, clinical studies have reported that 30% to 80% of infertile men exhibit supraphysiological levels of seminal ROS [208,209,210,211,212,213]. Furthermore, a study comparing ROS levels in semen from proven donor (those who resulted in a pregnancy) and infertile men, using chemiluminescence assay, reported that fertile men possessed reduced ROS levels [214]. Several studies have already established a correlation between oxidative stress and poor sperm parameters, including reduced sperm concentrations and motility, aberrant morphology, as well as increased DNA fragmentation [215,216,217,218,219].
Interestingly, ROS have been described as a double-edged sword for spermatozoa. ROS have a crucial influence in sperm physiology and spermatogenesis, as well as in the fertilization process. Seminal ROS participate as a secondary messenger in several pathways necessary for spermatozoa hyperactivation, capacitation, sperm–oocyte fusion, chemotaxis, acrosome reaction, and sperm head decondensation [220,221,222,223]. In addition to endogenous sources, ROS may also arise from exposure to pollutants, radiation, and drugs [224]. An imbalance in the redox homeostasis, as a result of increased ROS production or a deficient antioxidant system (or both), result in pathological oxidative damage to spermatozoa. Interestingly, it has been recently shown that antioxidant mechanisms, as is the case of Sperm Nuclear Basic Proteins (SNBP), when in contact with certain pollutants such as heavy metals can overturn their protective role and contribute to oxidative damage, which highlights how vulnerable spermatozoa are to oxidative stress [225,226]. Even though infertile men present lower seminal plasma antioxidant capacity [79,227], oxidative damage in spermatozoa is mostly associated with exacerbated levels of ROS as opposed to impaired superoxide scavenging activity [211]. Loss of motility is one of the major consequences of oxidative stress in spermatozoa [228], arising from a combination of reduced membrane fluidity, due to dysregulated lipid peroxidation [229], and defective axonemal phosphorylation [230,231,232]. Furthermore, oxidative stress has been associated with spermatozoa DNA damage. Disruption in sperm genomic integrity clinically relates to augmented teratozoospermia, impaired oocyte fertilization, abnormal embryonic development, and increased miscarriage rates [233,234,235].
As previously mentioned, individuals who suffer from obesity or diabetes mellitus present augmented levels of ROS, which results in oxidative stress and ultimately subfertility or even infertility. In fact, the deleterious effects of diabetes mellitus and obesity in male reproductive function are well established. Patients suffering from diabetes and/or obesity exhibit poor sperm parameter, decreased motility, abnormal morphology, reduced density, and higher damage to both nuclear and mitochondrial sperm DNA [236,237]. It was also reported that obese men have a higher mean scrotal temperature than healthy individuals, which can be related to male infertility [238,239]. Besides, hormonal dysfunctions that decrease steroidogenesis in Leydig cells, resulting in hypogonadism, and alterations in the body energy metabolism disturb the normal function of testis [157,237]. On the other hand, weight loss and improved lifestyle has been associated with better sperm parameters [240], although a series of studies by Crisóstomo et al. demonstrated that the consumption of a high fat diet during childhood permanently alters testicular metabolism which led to an irreversible decline in sperm quality [241,242]. Glucose metabolism is of great importance in the testis as sperm cells are highly sensitive to glucose concentrations. Accordingly, glucose transport through the blood-germ barrier is a highly controlled process. Hyperglycemia is particularly harmful to spermatogenesis being correlated with poor reproductive outcomes [243]. Sertoli and Leydig cells as well as spermatozoa have been shown to be sensitive to insulin, therefore, glucose homeostasis is necessary for male reproductive function [244]. Furthermore, both obese and diabetic patients have been reported to exhibit elevated ROS in testis and seminal fluid, which is, as described above, extremely detrimental [245,246].
Further understanding of the nefarious impact of ROS production and antioxidant mechanism in male related infertility will surely lead to more focused treatment and accurate diagnostic, as well as better outcomes for subfertile and infertile men. Hence, UCPs by virtue of their antioxidant activity may represent a clinically relevant target to improve reproductive outcomes. UCPs seem to play an important part in mediating the deleterious effects of diabesity in male infertility and subfertility, however little is known about their function in the male reproductive tract. Given the extensively described role of UCPs as key regulators of the mitochondria redox state, it should not come as a surprise that most of the experimental data comes from studies on the activity of UCPs in the testis/sperm ROS production, more precisely UCP2. Zhang et al. reported that UCP2 protects germ cells from hyperthermia-derived ROS production. Furthermore, Ucp2 mRNA expression was increased in the mouse testis of experimental cryptorchidism. Hyperthermia in the testis arises from multiple pathological conditions such as varicocele, cryptorchidism, or obesity, and correlates with exacerbated spermatozoa apoptosis [42]. In another study, it was found that ROS production is the main mechanism behind hyperthermia-induced sperm cell apoptosis. Acute exposure of mouse testicles to heat (43 °C for 5 min) was sufficient to induce a 6-fold increase in UCP2 expression. Furthermore, overexpression of UCP2 in mouse GC-2 cell line, a model for germ cell differentiating, improved the survival rate after exposure to the ROS-inducing agent menadione [42]. Moreover, older mice exhibit increased UCP2 activity and proton leak in the testis. Testicular age involves higher rates of mitochondrial dysfunction which correlates with increased oxidative stress, therefore, UCP2 upregulation in older testis might be a protective mechanism [53]. In humans, UCP2 expression was found to be significantly increased in spermatozoa from donors with normal sperm motility when compared to asthenospermic men [32]. The link between UCP2 and ROS production was further highlighted through treatment of spermatozoa with genipin, which resulted in elevated levels of mROS and decreased sperm motility. It should be noted that it was recently reported by Kreiter et al. that genipin is incapable of specifically inhibit UCP2, but rather is a nonspecific inhibitor of UCPs activity [247]. Additionally, treatment of spermatozoa with H2O2 was shown to induce UCP2 expression [32]. Although the experimental data on the role of UCPs in the male reproductive function is scarce, given their presence in testis and their established function in other tissues it should be expected that they are important modulators of testicular bioenergetics and redox state. Therefore, they are enticing targets for further research in the crosstalk between diabetes mellitus and obesity and their secondary effects in male reproduction.

6. Conclusions and Future Perspectives

UCPs are widely expressed throughout the organism and play an important role on ROS production and homeostasis. UCP1 is responsible for thermogenesis, but the other homologs display broader functions, including insulin release and regulation of oxidative phosphorylation. Due to their functions, UCPs take part in the pathophysiology of diabesity, where its altered expression and/or dysfunction is associated with the onset of the disease. Studying UCPs expression and function will lead to a better understanding of the pathophysiology of metabolic diseases and potentially to novel therapeutic targets. As mentioned above, diabesity is positively associated with high oxidative stress and, consequently, to male subfertility or even infertility, which is one of the most overlooked comorbidities. Working towards a better understanding of the relationship of UCPs activity with male reproductive function could lead to the development of novel fertility therapies and better reproductive outcomes. However, not much is known concerning the expression and function of UCPs in the male reproductive tract. Further studies are needed to disclosure the role of UCPs in male fertility and redox signaling and homeostasis during spermatogenesis and spermatozoa physiology.

Author Contributions

Conceptualization, D.F.C., P.F.O. and M.G.A.; investigation, B.S.M. and L.F.-B.; resources, M.G.A.; writing—original draft preparation, B.S.M. and L.F.-B.; writing—review and editing, D.F.C., P.F.O. and M.G.A.; supervision, P.F.O. and M.G.A.; project administration, M.G.A.; funding acquisition, M.G.A. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Portuguese Foundation for Science and Technology: D.F. Carrageta (SFRH/BD/136779/2018), M.G. Alves (IFCT2015 and PTDC/MEC-AND/28691/2017), UMIB (UIDB/00215/2020 and UIDP/00215/2020), ITR (LA/P/0064/2020) and QOPNA (UID/QUI/00062/2019) co-funded by FEDER funds (POCI/COMPETE 2020). Pedro F. Oliveira is funded by national funds through FCT—Fundação para a Ciência e a Tecnologia, I.P., under the Scientific Employment Stimulus—Institutional Call—reference CEECINST/00026/2018.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Saeedi, P.; Petersohn, I.; Salpea, P.; Malanda, B.; Karuranga, S.; Unwin, N.; Colagiuri, S.; Guariguata, L.; Motala, A.A.; Ogurtsova, K.; et al. Global and regional diabetes prevalence estimates for 2019 and projections for 2030 and 2045: Results from the International Diabetes Federation Diabetes Atlas, 9(th) edition. Diabetes Res. Clin. Pract. 2019, 157, 107843. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Nuertey, B.D.; Alhassan, A.I.; Nuertey, A.D.; Mensah, I.A.; Adongo, V.; Kabutey, C.; Addai, J.; Biritwum, R.B. Prevalence of obesity and overweight and its associated factors among registered pensioners in Ghana; A cross sectional studies. BMC Obes. 2017, 4, 26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Vincent, H.K.; Taylor, A.G. Biomarkers and potential mechanisms of obesity-induced oxidant stress in humans. Int. J. Obes. 2006, 30, 400–418. [Google Scholar] [CrossRef] [Green Version]
  4. Ferramosca, A.; Zara, V. Bioenergetics of mammalian sperm capacitation. Biomed. Res. Int. 2014, 2014, 902953. [Google Scholar] [CrossRef] [Green Version]
  5. Carrageta, D.F.; Guerra-Carvalho, B.; Sousa, M.; Barros, A.; Oliveira, P.F.; Monteiro, M.P.; Alves, M.G. Mitochondrial Activation and Reactive Oxygen-Species Overproduction during Sperm Capacitation are Independent of Glucose Stimuli. Antioxidants 2020, 9, 750. [Google Scholar] [CrossRef]
  6. Zhao, R.Z.; Jiang, S.; Zhang, L.; Yu, Z.B. Mitochondrial electron transport chain, ROS generation and uncoupling (Review). Int. J. Mol. Med. 2019, 44, 3–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Ramsden, D.B.; Ho, P.W.; Ho, J.W.; Liu, H.F.; So, D.H.; Tse, H.M.; Chan, K.H.; Ho, S.L. Human neuronal uncoupling proteins 4 and 5 (UCP4 and UCP5): Structural properties, regulation, and physiological role in protection against oxidative stress and mitochondrial dysfunction. Brain Behav. 2012, 2, 468–478. [Google Scholar] [CrossRef]
  8. Rhee, S.G. Cell signaling. H2O2, a necessary evil for cell signaling. Science 2006, 312, 1882–1883. [Google Scholar] [CrossRef]
  9. Chan, S.L.; Liu, D.; Kyriazis, G.A.; Bagsiyao, P.; Ouyang, X.; Mattson, M.P. Mitochondrial uncoupling protein-4 regulates calcium homeostasis and sensitivity to store depletion-induced apoptosis in neural cells. J. Biol. Chem. 2006, 281, 37391–37403. [Google Scholar] [CrossRef] [Green Version]
  10. Miwa, S.; Brand, M.D. Mitochondrial matrix reactive oxygen species production is very sensitive to mild uncoupling. Biochem. Soc. Trans. 2003, 31, 1300–1301. [Google Scholar] [CrossRef]
  11. Pecqueur, C.; Alves-Guerra, M.C.; Gelly, C.; Levi-Meyrueis, C.; Couplan, E.; Collins, S.; Ricquier, D.; Bouillaud, F.; Miroux, B. Uncoupling protein 2, in vivo distribution, induction upon oxidative stress, and evidence for translational regulation. J. Biol. Chem. 2001, 276, 8705–8712. [Google Scholar] [CrossRef] [Green Version]
  12. Hinz, W.; Gruninger, S.; de Pover, A.; Chiesi, M. Properties of the human long and short isoforms of the uncoupling protein-3 expressed in yeast cells. FEBS Lett. 1999, 462, 411–415. [Google Scholar] [CrossRef] [Green Version]
  13. Arsenijevic, D.; Onuma, H.; Pecqueur, C.; Raimbault, S.; Manning, B.S.; Miroux, B.; Couplan, E.; Alves-Guerra, M.C.; Goubern, M.; Surwit, R.; et al. Disruption of the uncoupling protein-2 gene in mice reveals a role in immunity and reactive oxygen species production. Nat. Genet. 2000, 26, 435–439. [Google Scholar] [CrossRef]
  14. Sanchis, D.; Fleury, C.; Chomiki, N.; Goubern, M.; Huang, Q.; Neverova, M.; Gregoire, F.; Easlick, J.; Raimbault, S.; Levi-Meyrueis, C.; et al. BMCP1, a novel mitochondrial carrier with high expression in the central nervous system of humans and rodents, and respiration uncoupling activity in recombinant yeast. J. Biol. Chem. 1998, 273, 34611–34615. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Mao, W.; Yu, X.X.; Zhong, A.; Li, W.; Brush, J.; Sherwood, S.W.; Adams, S.H.; Pan, G. UCP4, a novel brain-specific mitochondrial protein that reduces membrane potential in mammalian cells. FEBS Lett. 1999, 443, 326–330. [Google Scholar] [CrossRef] [Green Version]
  16. Eckerskorn, C.; Klingenberg, M. In the uncoupling protein from brown adipose tissue the C-terminus protrudes to the c-side of the membrane as shown by tryptic cleavage. FEBS Lett. 1987, 226, 166–170. [Google Scholar] [CrossRef] [Green Version]
  17. Miroux, B.; Frossard, V.; Raimbault, S.; Ricquier, D.; Bouillaud, F. The topology of the brown adipose tissue mitochondrial uncoupling protein determined with antibodies against its antigenic sites revealed by a library of fusion proteins. EMBO J. 1993, 12, 3739–3745. [Google Scholar] [CrossRef]
  18. Ricquier, D.; Bouillaud, F. The uncoupling protein homologues: UCP1, UCP2, UCP3, StUCP and AtUCP. Biochem. J. 2000, 345, 161–179. [Google Scholar] [CrossRef] [PubMed]
  19. Hoang, T.; Smith, M.D.; Jelokhani-Niaraki, M. Toward understanding the mechanism of ion transport activity of neuronal uncoupling proteins UCP2, UCP4, and UCP5. Biochemistry 2012, 51, 4004–4014. [Google Scholar] [CrossRef]
  20. Liu, D.; Chan, S.L.; de Souza-Pinto, N.C.; Slevin, J.R.; Wersto, R.P.; Zhan, M.; Mustafa, K.; de Cabo, R.; Mattson, M.P. Mitochondrial UCP4 mediates an adaptive shift in energy metabolism and increases the resistance of neurons to metabolic and oxidative stress. Neuromol. Med. 2006, 8, 389–414. [Google Scholar] [CrossRef] [Green Version]
  21. Virtanen, K.A.; Lidell, M.E.; Orava, J.; Heglind, M.; Westergren, R.; Niemi, T.; Taittonen, M.; Laine, J.; Savisto, N.J.; Enerback, S.; et al. Functional brown adipose tissue in healthy adults. N. Engl. J. Med. 2009, 360, 1518–1525. [Google Scholar] [CrossRef] [PubMed]
  22. Tiraby, C.; Tavernier, G.; Lefort, C.; Larrouy, D.; Bouillaud, F.; Ricquier, D.; Langin, D. Acquirement of brown fat cell features by human white adipocytes. J. Biol. Chem. 2003, 278, 33370–33376. [Google Scholar] [CrossRef] [Green Version]
  23. Mori, S.; Yoshizuka, N.; Takizawa, M.; Takema, Y.; Murase, T.; Tokimitsu, I.; Saito, M. Expression of uncoupling proteins in human skin and skin-derived cells. J. Investig. Dermatol. 2008, 128, 1894–1900. [Google Scholar] [CrossRef]
  24. Matthias, A.; Ohlson, K.B.; Fredriksson, J.M.; Jacobsson, A.; Nedergaard, J.; Cannon, B. Thermogenic responses in brown fat cells are fully UCP1-dependent. UCP2 or UCP3 do not substitute for UCP1 in adrenergically or fatty scid-induced thermogenesis. J. Biol. Chem. 2000, 275, 25073–25081. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Shabalina, I.G.; Petrovic, N.; de Jong, J.M.; Kalinovich, A.V.; Cannon, B.; Nedergaard, J. UCP1 in brite/beige adipose tissue mitochondria is functionally thermogenic. Cell Rep. 2013, 5, 1196–1203. [Google Scholar] [CrossRef] [Green Version]
  26. Fujita, H.; Habuta, M.; Hattori, T.; Kubota, S.; Kumon, H.; Ohuchi, H. UCP1 expression in the mouse adrenal gland is not upregulated by thermogenic conditions. Biochem. Biophys. Res. Commun. 2021, 566, 184–189. [Google Scholar] [CrossRef] [PubMed]
  27. Jakus, P.B.; Sipos, K.; Kispal, G.; Sandor, A. Opposite regulation of uncoupling protein 1 and uncoupling protein 3 in vivo in brown adipose tissue of cold-exposed rats. FEBS Lett. 2002, 519, 210–214. [Google Scholar] [CrossRef] [Green Version]
  28. Sepa-Kishi, D.M.; Jani, S.; da Eira, D.; Ceddia, R.B. Cold acclimation enhances UCP1 content, lipolysis, and triacylglycerol resynthesis, but not mitochondrial uncoupling and fat oxidation, in rat white adipocytes. Am. J. Physiol. Cell Physiol. 2019, 316, C365–C376. [Google Scholar] [CrossRef]
  29. Nigro, M.; de Sanctis, C.; Formisano, P.; Stanzione, R.; Forte, M.; Capasso, G.; Gigliotti, G.; Rubattu, S.; Viggiano, D. Cellular and subcellular localization of uncoupling protein 2 in the human kidney. J. Mol. Histol. 2018, 49, 437–445. [Google Scholar] [CrossRef]
  30. Haguenauer, A.; Raimbault, S.; Masscheleyn, S.; Gonzalez-Barroso Mdel, M.; Criscuolo, F.; Plamondon, J.; Miroux, B.; Ricquier, D.; Richard, D.; Bouillaud, F.; et al. A new renal mitochondrial carrier, KMCP1, is up-regulated during tubular cell regeneration and induction of antioxidant enzymes. J. Biol. Chem. 2005, 280, 22036–22043. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Stark, M.J.; Hodyl, N.A.; Butler, M.; Clifton, V.L. Localisation and characterisation of uncoupling protein-2 (UCP2) in the human preterm placenta. Placenta 2012, 33, 1020–1025. [Google Scholar] [CrossRef]
  32. Wang, X.; Qian, H.; Huang, X.; Li, J.; Zhang, J.; Zhu, N.; Chen, H.; Zhu, C.; Wang, J.; Zhang, P.; et al. UCP2 Mitigates the Loss of Human Spermatozoa Motility by Promoting mROS Elimination. Cell Physiol. Biochem. 2018, 50, 952–962. [Google Scholar] [CrossRef]
  33. Boss, O.; Samec, S.; Paoloni-Giacobino, A.; Rossier, C.; Dulloo, A.; Seydoux, J.; Muzzin, P.; Giacobino, J.P. Uncoupling protein-3: A new member of the mitochondrial carrier family with tissue-specific expression. FEBS Lett. 1997, 408, 39–42. [Google Scholar] [CrossRef] [Green Version]
  34. Chan, C.B.; de Leo, D.; Joseph, J.W.; McQuaid, T.S.; Ha, X.F.; Xu, F.; Tsushima, R.G.; Pennefather, P.S.; Salapatek, A.M.; Wheeler, M.B. Increased uncoupling protein-2 levels in beta-cells are associated with impaired glucose-stimulated insulin secretion: Mechanism of action. Diabetes 2001, 50, 1302–1310. [Google Scholar] [CrossRef] [Green Version]
  35. Fleury, C.; Neverova, M.; Collins, S.; Raimbault, S.; Champigny, O.; Levi-Meyrueis, C.; Bouillaud, F.; Seldin, M.F.; Surwit, R.S.; Ricquier, D.; et al. Uncoupling protein-2: A novel gene linked to obesity and hyperinsulinemia. Nat. Genet. 1997, 15, 269–272. [Google Scholar] [CrossRef]
  36. Krauss, S.; Zhang, C.Y.; Lowell, B.B. A significant portion of mitochondrial proton leak in intact thymocytes depends on expression of UCP2. Proc. Natl. Acad. Sci. USA 2002, 99, 118–122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Jiang, L.; Qiu, W.; Zhou, Y.; Wen, P.; Fang, L.; Cao, H.; Zen, K.; He, W.; Zhang, C.; Dai, C.; et al. A microRNA-30e/mitochondrial uncoupling protein 2 axis mediates TGF-beta1-induced tubular epithelial cell extracellular matrix production and kidney fibrosis. Kidney Int. 2013, 84, 285–296. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Rousset, S.; Alves-Guerra, M.C.; Ouadghiri-Bencherif, S.; Kozak, L.P.; Miroux, B.; Richard, D.; Bouillaud, F.; Ricquier, D.; Cassard-Doulcier, A.M. Uncoupling protein 2, but not uncoupling protein 1, is expressed in the female mouse reproductive tract. J. Biol. Chem. 2003, 278, 45843–45847. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Cao, T.; Dong, Y.; Tang, R.; Chen, J.; Zhang, C.Y.; Zen, K. Mitochondrial uncoupling protein 2 protects splenocytes from oxidative stress-induced apoptosis during pathogen activation. Cell Immunol. 2013, 286, 39–44. [Google Scholar] [CrossRef] [PubMed]
  40. Arsenijevic, D.; Clavel, S.; Sanchis, D.; Plamondon, J.; Huang, Q.; Ricquier, D.; Rouger, L.; Richard, D. Induction of Ucp2 expression in brain phagocytes and neurons following murine toxoplasmosis: An essential role of IFN-gamma and an association with negative energy balance. J. Neuroimmunol. 2007, 186, 121–132. [Google Scholar] [CrossRef]
  41. Motloch, L.J.; Larbig, R.; Gebing, T.; Reda, S.; Schwaiger, A.; Leitner, J.; Wolny, M.; Eckardt, L.; Hoppe, U.C. By Regulating Mitochondrial Ca2+-Uptake UCP2 Modulates Intracellular Ca2+. PLoS ONE 2016, 11, e0148359. [Google Scholar] [CrossRef]
  42. Zhang, K.; Shang, Y.; Liao, S.; Zhang, W.; Nian, H.; Liu, Y.; Chen, Q.; Han, C. Uncoupling protein 2 protects testicular germ cells from hyperthermia-induced apoptosis. Biochem. Biophys. Res. Commun. 2007, 360, 327–332. [Google Scholar] [CrossRef]
  43. Diao, J.; Allister, E.M.; Koshkin, V.; Lee, S.C.; Bhattacharjee, A.; Tang, C.; Giacca, A.; Chan, C.B.; Wheeler, M.B. UCP2 is highly expressed in pancreatic alpha-cells and influences secretion and survival. Proc. Natl. Acad. Sci. USA 2008, 105, 12057–12062. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Yu, X.X.; Barger, J.L.; Boyer, B.B.; Brand, M.D.; Pan, G.; Adams, S.H. Impact of endotoxin on UCP homolog mRNA abundance, thermoregulation, and mitochondrial proton leak kinetics. Am. J. Physiol. Endocrinol. Metab. 2000, 279, E433–E446. [Google Scholar] [CrossRef]
  45. Rousset, S.; Emre, Y.; Join-Lambert, O.; Hurtaud, C.; Ricquier, D.; Cassard-Doulcier, A.M. The uncoupling protein 2 modulates the cytokine balance in innate immunity. Cytokine 2006, 35, 135–142. [Google Scholar] [CrossRef] [PubMed]
  46. Chen, X.; Wang, K.; Chen, J.; Guo, J.; Yin, Y.; Cai, X.; Guo, X.; Wang, G.; Yang, R.; Zhu, L.; et al. In vitro evidence suggests that miR-133a-mediated regulation of uncoupling protein 2 (UCP2) is an indispensable step in myogenic differentiation. J. Biol. Chem. 2009, 284, 5362–5369. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Friederich, M.; Fasching, A.; Hansell, P.; Nordquist, L.; Palm, F. Diabetes-induced up-regulation of uncoupling protein-2 results in increased mitochondrial uncoupling in kidney proximal tubular cells. Biochim. Biophys. Acta 2008, 1777, 935–940. [Google Scholar] [CrossRef] [Green Version]
  48. Larrouy, D.; Laharrague, P.; Carrera, G.; Viguerie-Bascands, N.; Levi-Meyrueis, C.; Fleury, C.; Pecqueur, C.; Nibbelink, M.; Andre, M.; Casteilla, L.; et al. Kupffer cells are a dominant site of uncoupling protein 2 expression in rat liver. Biochem. Biophys. Res. Commun. 1997, 235, 760–764. [Google Scholar] [CrossRef]
  49. De Simone, R.; Ajmone-Cat, M.A.; Pandolfi, M.; Bernardo, A.; de Nuccio, C.; Minghetti, L.; Visentin, S. The mitochondrial uncoupling protein-2 is a master regulator of both M1 and M2 microglial responses. J. Neurochem. 2015, 135, 147–156. [Google Scholar] [CrossRef]
  50. 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]
  51. Teshima, Y.; Akao, M.; Jones, S.P.; Marban, E. Uncoupling protein-2 overexpression inhibits mitochondrial death pathway in cardiomyocytes. Circ. Res. 2003, 93, 192–200. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Cadenas, S.; Buckingham, J.A.; Samec, S.; Seydoux, J.; Din, N.; Dulloo, A.G.; Brand, M.D. UCP2 and UCP3 rise in starved rat skeletal muscle but mitochondrial proton conductance is unchanged. FEBS Lett. 1999, 462, 257–260. [Google Scholar] [CrossRef] [Green Version]
  53. Amaral, S.; Mota, P.; Rodrigues, A.S.; Martins, L.; Oliveira, P.J.; Ramalho-Santos, J. Testicular aging involves mitochondrial dysfunction as well as an increase in UCP2 levels and proton leak. FEBS Lett. 2008, 582, 4191–4196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Chan, C.B.; MacDonald, P.E.; Saleh, M.C.; Johns, D.C.; Marban, E.; Wheeler, M.B. Overexpression of uncoupling protein 2 inhibits glucose-stimulated insulin secretion from rat islets. Diabetes 1999, 48, 1482–1486. [Google Scholar] [CrossRef] [PubMed]
  55. Alan, L.; Smolkova, K.; Kronusova, E.; Santorova, J.; Jezek, P. Absolute levels of transcripts for mitochondrial uncoupling proteins UCP2, UCP3, UCP4, and UCP5 show different patterns in rat and mice tissues. J. Bioenerg. Biomembr. 2009, 41, 71–78. [Google Scholar] [CrossRef] [PubMed]
  56. Negre-Salvayre, A.; Hirtz, C.; Carrera, G.; Cazenave, R.; Troly, M.; Salvayre, R.; Penicaud, L.; Casteilla, L. A role for uncoupling protein-2 as a regulator of mitochondrial hydrogen peroxide generation. FASEB J. 1997, 11, 809–815. [Google Scholar] [CrossRef] [PubMed]
  57. Carroll, A.M.; Porter, R.K. Starvation-sensitive UCP 3 protein expression in thymus and spleen mitochondria. Biochim. Biophys. Acta 2004, 1700, 145–150. [Google Scholar] [CrossRef] [PubMed]
  58. Li, Y.; Maedler, K.; Shu, L.; Haataja, L. UCP-2 and UCP-3 proteins are differentially regulated in pancreatic beta-cells. PLoS ONE 2008, 3, e1397. [Google Scholar] [CrossRef] [PubMed]
  59. Hesselink, M.K.; Greenhaff, P.L.; Constantin-Teodosiu, D.; Hultman, E.; Saris, W.H.; Nieuwlaat, R.; Schaart, G.; Kornips, E.; Schrauwen, P. Increased uncoupling protein 3 content does not affect mitochondrial function in human skeletal muscle in vivo. J. Clin. Investig. 2003, 111, 479–486. [Google Scholar] [CrossRef] [PubMed]
  60. Vidal-Puig, A.; Solanes, G.; Grujic, D.; Flier, J.S.; Lowell, B.B. UCP3: An uncoupling protein homologue expressed preferentially and abundantly in skeletal muscle and brown adipose tissue. Biochem. Biophys. Res. Commun. 1997, 235, 79–82. [Google Scholar] [CrossRef] [PubMed]
  61. Hilse, K.E.; Kalinovich, A.V.; Rupprecht, A.; Smorodchenko, A.; Zeitz, U.; Staniek, K.; Erben, R.G.; Pohl, E.E. The expression of UCP3 directly correlates to UCP1 abundance in brown adipose tissue. Biochim. Biophys. Acta 2016, 1857, 72–78. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Kelly, O.M.; Porter, R.K. Absence of mitochondrial uncoupling protein 3: Effect on thymus and spleen in the fed and fasted mice. Biochim. Biophys. Acta 2011, 1807, 1064–1074. [Google Scholar] [CrossRef] [Green Version]
  63. O’Connor, E.B.; Munoz-Wolf, N.; Leon, G.; Lavelle, E.C.; Mills, K.H.G.; Walsh, P.T.; Porter, R.K. UCP3 reciprocally controls CD4+ Th17 and Treg cell differentiation. PLoS ONE 2020, 15, e0239713. [Google Scholar] [CrossRef]
  64. Boudina, S.; Han, Y.H.; Pei, S.; Tidwell, T.J.; Henrie, B.; Tuinei, J.; Olsen, C.; Sena, S.; Abel, E.D. UCP3 regulates cardiac efficiency and mitochondrial coupling in high fat-fed mice but not in leptin-deficient mice. Diabetes 2012, 61, 3260–3269. [Google Scholar] [CrossRef] [Green Version]
  65. Huang, Z.; Li, J.; Du, S.; Chen, G.; Qi, Y.; Huang, L.; Xiao, L.; Tong, P. Effects of UCP4 on the Proliferation and Apoptosis of Chondrocytes: Its Possible Involvement and Regulation in Osteoarthritis. PLoS ONE 2016, 11, e0150684. [Google Scholar] [CrossRef] [PubMed]
  66. Smorodchenko, A.; Rupprecht, A.; Sarilova, I.; Ninnemann, O.; Brauer, A.U.; Franke, K.; Schumacher, S.; Techritz, S.; Nitsch, R.; Schuelke, M.; et al. Comparative analysis of uncoupling protein 4 distribution in various tissues under physiological conditions and during development. Biochim. Biophys. Acta 2009, 1788, 2309–2319. [Google Scholar] [CrossRef] [Green Version]
  67. Smorodchenko, A.; Rupprecht, A.; Fuchs, J.; Gross, J.; Pohl, E.E. Role of mitochondrial uncoupling protein 4 in rat inner ear. Mol. Cell Neurosci. 2011, 47, 244–253. [Google Scholar] [CrossRef]
  68. Kitahara, T.; Li, H.S.; Balaban, C.D. Localization of the mitochondrial uncoupling protein family in the rat inner ear. Hear. Res. 2004, 196, 39–48. [Google Scholar] [CrossRef]
  69. Donhoffer, S.; Szegvari, G.; Jarai, I.; Farkas, M. Thermoregulatory heat production in the brain. Nature 1959, 184 (Suppl. 13), 993–994. [Google Scholar] [CrossRef]
  70. Kempuraj, D.; Thangavel, R.; Fattal, R.; Pattani, S.; Yang, E.; Zaheer, S.; Santillan, D.A.; Santillan, M.K.; Zaheer, A. Mast Cells Release Chemokine CCL2 in Response to Parkinsonian Toxin 1-Methyl-4-Phenyl-Pyridinium (MPP(+)). Neurochem. Res. 2016, 41, 1042–1049. [Google Scholar] [CrossRef] [Green Version]
  71. Kitahara, T.; Li-Korotky, H.S.; Balaban, C.D. Regulation of mitochondrial uncoupling proteins in mouse inner ear ganglion cells in response to systemic kanamycin challenge. Neuroscience 2005, 135, 639–653. [Google Scholar] [CrossRef] [PubMed]
  72. Lengacher, S.; Magistretti, P.J.; Pellerin, L. Quantitative rt-PCR analysis of uncoupling protein isoforms in mouse brain cortex: Methodological optimization and comparison of expression with brown adipose tissue and skeletal muscle. J. Cereb. Blood Flow Metab. 2004, 24, 780–788. [Google Scholar] [CrossRef] [Green Version]
  73. Yu, X.X.; Mao, W.; Zhong, A.; Schow, P.; Brush, J.; Sherwood, S.W.; Adams, S.H.; Pan, G. Characterization of novel UCP5/BMCP1 isoforms and differential regulation of UCP4 and UCP5 expression through dietary or temperature manipulation. FASEB J. 2000, 14, 1611–1618. [Google Scholar] [CrossRef] [PubMed]
  74. 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] [Green Version]
  75. Erden, Y.; Tekin, S.; Sandal, S.; Onalan, E.E.; Tektemur, A.; Kirbag, S. Effects of central irisin administration on the uncoupling proteins in rat brain. Neurosci. Lett. 2016, 618, 6–13. [Google Scholar] [CrossRef]
  76. Huang, P.S.; Son, J.H.; Abbott, L.C.; Winzer-Serhan, U.H. Regulated expression of neuronal SIRT1 and related genes by aging and neuronal beta2-containing nicotinic cholinergic receptors. Neuroscience 2011, 196, 189–202. [Google Scholar] [CrossRef]
  77. Vidal-Puig, A.J.; Grujic, D.; Zhang, C.Y.; Hagen, T.; Boss, O.; Ido, Y.; Szczepanik, A.; Wade, J.; Mootha, V.; Cortright, R.; et al. Energy metabolism in uncoupling protein 3 gene knockout mice. J. Biol. Chem. 2000, 275, 16258–16266. [Google Scholar] [CrossRef] [Green Version]
  78. Gorgoglione, R.; Porcelli, V.; Santoro, A.; Daddabbo, L.; Vozza, A.; Monne, M.; di Noia, M.A.; Palmieri, L.; Fiermonte, G.; Palmieri, F. The human uncoupling proteins 5 and 6 (UCP5/SLC25A14 and UCP6/SLC25A30) transport sulfur oxyanions, phosphate and dicarboxylates. Biochim. Biophys. Acta 2019, 1860, 724–733. [Google Scholar] [CrossRef]
  79. Ko, E.Y.; Sabanegh, E.S., Jr.; Agarwal, A. Male infertility testing: Reactive oxygen species and antioxidant capacity. Fertil. Steril. 2014, 102, 1518–1527. [Google Scholar] [CrossRef] [PubMed]
  80. Okamatsu-Ogura, Y.; Fukano, K.; Tsubota, A.; Uozumi, A.; Terao, A.; Kimura, K.; Saito, M. Thermogenic ability of uncoupling protein 1 in beige adipocytes in mice. PLoS ONE 2013, 8, e84229. [Google Scholar] [CrossRef] [Green Version]
  81. Nibbelink, M.; Moulin, K.; Arnaud, E.; Duval, C.; Penicaud, L.; Casteilla, L. Brown fat UCP1 is specifically expressed in uterine longitudinal smooth muscle cells. J. Biol. Chem. 2001, 276, 47291–47295. [Google Scholar] [CrossRef] [Green Version]
  82. Carroll, A.M.; Haines, L.R.; Pearson, T.W.; Fallon, P.G.; Walsh, C.M.; Brennan, C.M.; Breen, E.P.; Porter, R.K. Identification of a functioning mitochondrial uncoupling protein 1 in thymus. J. Biol. Chem. 2005, 280, 15534–15543. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Frontini, A.; Rousset, S.; Cassard-Doulcier, A.M.; Zingaretti, C.; Ricquier, D.; Cinti, S. Thymus uncoupling protein 1 is exclusive to typical brown adipocytes and is not found in thymocytes. J. Histochem. Cytochem. 2007, 55, 183–189. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Lizio, M.; Harshbarger, J.; Shimoji, H.; Severin, J.; Kasukawa, T.; Sahin, S.; Abugessaisa, I.; Fukuda, S.; Hori, F.; Ishikawa-Kato, S.; et al. Gateways to the FANTOM5 promoter level mammalian expression atlas. Genome Biol. 2015, 16, 22. [Google Scholar] [CrossRef] [Green Version]
  85. Darley-Usmar, V. The powerhouse takes control of the cell; The role of mitochondria in signal transduction. Free Radic. Biol. Med. 2004, 37, 753–754. [Google Scholar] [CrossRef] [PubMed]
  86. Ricquier, D.; Bouillaud, F.; Toumelin, P.; Mory, G.; Bazin, R.; Arch, J.; Penicaud, L. Expression of uncoupling protein mRNA in thermogenic or weakly thermogenic brown adipose tissue. Evidence for a rapid beta-adrenoreceptor-mediated and transcriptionally regulated step during activation of thermogenesis. J. Biol. Chem. 1986, 261, 13905–13910. [Google Scholar] [CrossRef]
  87. Busiello, R.A.; Savarese, S.; Lombardi, A. Mitochondrial uncoupling proteins and energy metabolism. Front. Physiol. 2015, 6, 36. [Google Scholar] [CrossRef] [Green Version]
  88. Rabelo, R.; Schifman, A.; Rubio, A.; Sheng, X.; Silva, J.E. Delineation of thyroid hormone-responsive sequences within a critical enhancer in the rat uncoupling protein gene. Endocrinology 1995, 136, 1003–1013. [Google Scholar] [CrossRef]
  89. Echtay, K.S.; Roussel, D.; St-Pierre, J.; Jekabsons, M.B.; Cadenas, S.; Stuart, J.A.; Harper, J.A.; Roebuck, S.J.; Morrison, A.; Pickering, S.; et al. Superoxide activates mitochondrial uncoupling proteins. Nature 2002, 415, 96–99. [Google Scholar] [CrossRef]
  90. Alvarez, R.; de Andres, J.; Yubero, P.; Vinas, O.; Mampel, T.; Iglesias, R.; Giralt, M.; Villarroya, F. A novel regulatory pathway of brown fat thermogenesis. Retinoic acid is a transcriptional activator of the mitochondrial uncoupling protein gene. J. Biol. Chem. 1995, 270, 5666–5673. [Google Scholar] [CrossRef] [Green Version]
  91. Gong, D.W.; He, Y.; Karas, M.; Reitman, M. Uncoupling protein-3 is a mediator of thermogenesis regulated by thyroid hormone, beta3-adrenergic agonists, and leptin. J. Biol. Chem. 1997, 272, 24129–24132. [Google Scholar] [CrossRef] [Green Version]
  92. Rial, E.; Muga, A.; Valpuesta, J.M.; Arrondo, J.L.; Goni, F.M. Infrared spectroscopic studies of detergent-solubilized uncoupling protein from brown-adipose-tissue mitochondria. Eur. J. Biochem. 1990, 188, 83–89. [Google Scholar] [CrossRef] [PubMed]
  93. Gimeno, R.E.; Dembski, M.; Weng, X.; Deng, N.; Shyjan, A.W.; Gimeno, C.J.; Iris, F.; Ellis, S.J.; Woolf, E.A.; Tartaglia, L.A. Cloning and characterization of an uncoupling protein homolog: A potential molecular mediator of human thermogenesis. Diabetes 1997, 46, 900–906. [Google Scholar] [CrossRef] [Green Version]
  94. Krauss, S.; Zhang, C.Y.; Scorrano, L.; Dalgaard, L.T.; St-Pierre, J.; Grey, S.T.; Lowell, B.B. Superoxide-mediated activation of uncoupling protein 2 causes pancreatic beta cell dysfunction. J. Clin. Investig. 2003, 112, 1831–1842. [Google Scholar] [CrossRef]
  95. Horimoto, M.; Fulop, P.; Derdak, Z.; Wands, J.R.; Baffy, G. Uncoupling protein-2 deficiency promotes oxidant stress and delays liver regeneration in mice. Hepatology 2004, 39, 386–392. [Google Scholar] [CrossRef]
  96. Brand, M.D.; Esteves, T.C. Physiological functions of the mitochondrial uncoupling proteins UCP2 and UCP3. Cell Metab. 2005, 2, 85–93. [Google Scholar] [CrossRef] [Green Version]
  97. Lee, K.U.; Lee, I.K.; Han, J.; Song, D.K.; Kim, Y.M.; Song, H.S.; Kim, H.S.; Lee, W.J.; Koh, E.H.; Song, K.H.; et al. Effects of recombinant adenovirus-mediated uncoupling protein 2 overexpression on endothelial function and apoptosis. Circ. Res. 2005, 96, 1200–1207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Alves-Guerra, M.C.; Rousset, S.; Pecqueur, C.; Mallat, Z.; Blanc, J.; Tedgui, A.; Bouillaud, F.; Cassard-Doulcier, A.M.; Ricquier, D.; Miroux, B. Bone marrow transplantation reveals the in vivo expression of the mitochondrial uncoupling protein 2 in immune and nonimmune cells during inflammation. J. Biol. Chem. 2003, 278, 42307–42312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Nowak, W.N.; Deng, J.; Ruan, X.Z.; Xu, Q. Reactive Oxygen Species Generation and Atherosclerosis. Arterioscler. Thromb. Vasc. Biol. 2017, 37, e41–e52. [Google Scholar] [CrossRef] [PubMed]
  100. 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]
  101. Horvath, B.; Spies, C.; Horvath, G.; Kox, W.J.; Miyamoto, S.; Barry, S.; Warden, C.H.; Bechmann, I.; Diano, S.; Heemskerk, J.; et al. Uncoupling protein 2 (UCP2) lowers alcohol sensitivity and pain threshold. Biochem. Pharmacol. 2002, 64, 369–374. [Google Scholar] [CrossRef]
  102. Nakase, T.; Yoshida, Y.; Nagata, K. Amplified expression of uncoupling proteins in human brain ischemic lesions. Neuropathology 2007, 27, 442–447. [Google Scholar] [CrossRef] [PubMed]
  103. McLeod, C.J.; Aziz, A.; Hoyt, R.F., Jr.; McCoy, J.P., Jr.; Sack, M.N. Uncoupling proteins 2 and 3 function in concert to augment tolerance to cardiac ischemia. J. Biol. Chem. 2005, 280, 33470–33476. [Google Scholar] [CrossRef] [Green Version]
  104. Kizaki, T.; Suzuki, K.; Hitomi, Y.; Taniguchi, N.; Saitoh, D.; Watanabe, K.; Onoe, K.; Day, N.K.; Good, R.A.; Ohno, H. Uncoupling protein 2 plays an important role in nitric oxide production of lipopolysaccharide-stimulated macrophages. Proc. Natl. Acad. Sci. USA 2002, 99, 9392–9397. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Zhang, C.Y.; Baffy, G.; Perret, P.; Krauss, S.; Peroni, O.; Grujic, D.; Hagen, T.; Vidal-Puig, A.J.; Boss, O.; Kim, Y.B.; et al. Uncoupling protein-2 negatively regulates insulin secretion and is a major link between obesity, beta cell dysfunction, and type 2 diabetes. Cell 2001, 105, 745–755. [Google Scholar] [CrossRef] [Green Version]
  106. Vozza, A.; Parisi, G.; de Leonardis, F.; Lasorsa, F.M.; Castegna, A.; Amorese, D.; Marmo, R.; Calcagnile, V.M.; Palmieri, L.; Ricquier, D.; et al. UCP2 transports C4 metabolites out of mitochondria, regulating glucose and glutamine oxidation. Proc. Natl. Acad. Sci. USA 2014, 111, 960–965. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Azzu, V.; Brand, M.D. The on-off switches of the mitochondrial uncoupling proteins. Trends Biochem. Sci. 2010, 35, 298–307. [Google Scholar] [CrossRef] [Green Version]
  108. Hurtaud, C.; Gelly, C.; Chen, Z.; Levi-Meyrueis, C.; Bouillaud, F. Glutamine stimulates translation of uncoupling protein 2mRNA. Cell. Mol. Life Sci. 2007, 64, 1853–1860. [Google Scholar] [CrossRef] [PubMed]
  109. Zhou, Y.T.; Shimabukuro, M.; Koyama, K.; Lee, Y.; Wang, M.Y.; Trieu, F.; Newgard, C.B.; Unger, R.H. Induction by leptin of uncoupling protein-2 and enzymes of fatty acid oxidation. Proc. Natl. Acad. Sci. USA 1997, 94, 6386–6390. [Google Scholar] [CrossRef] [Green Version]
  110. Brand, M.D.; Pamplona, R.; Portero-Otin, M.; Requena, J.R.; Roebuck, S.J.; Buckingham, J.A.; Clapham, J.C.; Cadenas, S. Oxidative damage and phospholipid fatty acyl composition in skeletal muscle mitochondria from mice underexpressing or overexpressing uncoupling protein 3. Biochem. J. 2002, 368, 597–603. [Google Scholar] [CrossRef]
  111. Talbot, D.A.; Lambert, A.J.; Brand, M.D. Production of endogenous matrix superoxide from mitochondrial complex I leads to activation of uncoupling protein 3. FEBS Lett. 2004, 556, 111–115. [Google Scholar] [CrossRef] [Green Version]
  112. Vincent, A.M.; Olzmann, J.A.; Brownlee, M.; Sivitz, W.I.; Russell, J.W. Uncoupling proteins prevent glucose-induced neuronal oxidative stress and programmed cell death. Diabetes 2004, 53, 726–734. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Cadenas, S.; Echtay, K.S.; Harper, J.A.; Jekabsons, M.B.; Buckingham, J.A.; Grau, E.; Abuin, A.; Chapman, H.; Clapham, J.C.; Brand, M.D. The basal proton conductance of skeletal muscle mitochondria from transgenic mice overexpressing or lacking uncoupling protein-3. J. Biol. Chem. 2002, 277, 2773–2778. [Google Scholar] [CrossRef] [Green Version]
  114. Liu, Q.; Bai, C.; Chen, F.; Wang, R.; MacDonald, T.; Gu, M.; Zhang, Q.; Morsy, M.A.; Caskey, C.T. Uncoupling protein-3: A muscle-specific gene upregulated by leptin in ob/ob mice. Gene 1998, 207, 1–7. [Google Scholar] [CrossRef]
  115. Larkin, S.; Mull, E.; Miao, W.; Pittner, R.; Albrandt, K.; Moore, C.; Young, A.; Denaro, M.; Beaumont, K. Regulation of the third member of the uncoupling protein family, UCP3, by cold and thyroid hormone. Biochem. Biophys. Res. Commun. 1997, 240, 222–227. [Google Scholar] [CrossRef]
  116. Boss, O.; Samec, S.; Kuhne, F.; Bijlenga, P.; Assimacopoulos-Jeannet, F.; Seydoux, J.; Giacobino, J.P.; Muzzin, P. Uncoupling protein-3 expression in rodent skeletal muscle is modulated by food intake but not by changes in environmental temperature. J. Biol. Chem. 1998, 273, 5–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Millet, L.; Vidal, H.; Andreelli, F.; Larrouy, D.; Riou, J.P.; Ricquier, D.; Laville, M.; Langin, D. Increased uncoupling protein-2 and -3 mRNA expression during fasting in obese and lean humans. J. Clin. Investig. 1997, 100, 2665–2670. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Huang, J.L.; Jing, X.; Tian, X.; Qin, M.C.; Xu, Z.H.; Wu, D.P.; Zhong, Z.G. Neuroprotective Properties of Panax notoginseng Saponins via Preventing Oxidative Stress Injury in SAMP8 Mice. Evid Based Complement. Alternat. Med. 2017, 2017, 8713561. [Google Scholar] [CrossRef] [Green Version]
  119. Xu, S.; Yang, X.; Qian, Y.; Xiao, Q. Parkinson’s disease-related DJ-1 modulates the expression of uncoupling protein 4 against oxidative stress. J. Neurochem. 2018, 145, 312–322. [Google Scholar] [CrossRef] [Green Version]
  120. Xu, Y.; Peng, S.; Cao, X.; Qian, S.; Shen, S.; Luo, J.; Zhang, X.; Sun, H.; Shen, W.L.; Jia, W.; et al. High doses of butyrate induce a reversible body temperature drop through transient proton leak in mitochondria of brain neurons. Life Sci. 2021, 278, 119614. [Google Scholar] [CrossRef]
  121. Viggiano, E.; Monda, V.; Messina, A.; Moscatelli, F.; Valenzano, A.; Tafuri, D.; Cibelli, G.; de Luca, B.; Messina, G.; Monda, M. Cortical spreading depression produces a neuroprotective effect activating mitochondrial uncoupling protein-5. Neuropsychiatr. Dis. Treat. 2016, 12, 1705–1710. [Google Scholar] [CrossRef] [Green Version]
  122. Pichiule, P.; Chavez, J.C.; LaManna, J.C. Oxygen and oxidative stress modulate the expression of uncoupling protein-5 in vitro and in vivo. Adv. Exp. Med. Biol. 2003, 540, 103–107. [Google Scholar] [CrossRef]
  123. Xie, Z.Z.; Liu, Y.; Bian, J.S. Hydrogen Sulfide and Cellular Redox Homeostasis. Oxid. Med. Cell. Longev. 2016, 2016, 6043038. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Nohara, K.; Tateishi, Y.; Suzuki, T.; Okamura, K.; Murai, H.; Takumi, S.; Maekawa, F.; Nishimura, N.; Kobori, M.; Ito, T. Late-onset increases in oxidative stress and other tumorigenic activities and tumors with a Ha-ras mutation in the liver of adult male C3H mice gestationally exposed to arsenic. Toxicol. Sci. 2012, 129, 293–304. [Google Scholar] [CrossRef] [Green Version]
  125. Mydin, R.; Okekpa, S.I. Reactive oxygen species, cellular redox homeostasis and cancer. In Homeostasis-An Integrated Vision; IntechOpen: London, UK, 2019. [Google Scholar] [CrossRef] [Green Version]
  126. Papa, S.; Skulachev, V.P. Reactive oxygen species, mitochondria, apoptosis and aging. Mol. Cell. Biochem. 1997, 174, 305–319. [Google Scholar] [CrossRef]
  127. Cadenas, S. Mitochondrial uncoupling, ROS generation and cardioprotection. Biochim. Biophys. Acta 2018, 1859, 940–950. [Google Scholar] [CrossRef] [PubMed]
  128. Mailloux, R.J.; Harper, M.E. Uncoupling proteins and the control of mitochondrial reactive oxygen species production. Free Radic. Biol. Med. 2011, 51, 1106–1115. [Google Scholar] [CrossRef]
  129. Azzu, V.; Jastroch, M.; Divakaruni, A.S.; Brand, M.D. The regulation and turnover of mitochondrial uncoupling proteins. Biochim. Biophys. Acta 2010, 1797, 785–791. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Dlaskova, A.; Clarke, K.J.; Porter, R.K. The role of UCP 1 in production of reactive oxygen species by mitochondria isolated from brown adipose tissue. Biochim. Biophys. Acta 2010, 1797, 1470–1476. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Oelkrug, R.; Kutschke, M.; Meyer, C.W.; Heldmaier, G.; Jastroch, M. Uncoupling protein 1 decreases superoxide production in brown adipose tissue mitochondria. J. Biol. Chem. 2010, 285, 21961–21968. [Google Scholar] [CrossRef] [Green Version]
  132. Stier, A.; Bize, P.; Habold, C.; Bouillaud, F.; Massemin, S.; Criscuolo, F. Mitochondrial uncoupling prevents cold-induced oxidative stress: A case study using UCP1 knockout mice. J. Exp. Biol. 2014, 217, 624–630. [Google Scholar] [CrossRef] [Green Version]
  133. Chouchani, E.T.; Kazak, L.; Jedrychowski, M.P.; Lu, G.Z.; Erickson, B.K.; Szpyt, J.; Pierce, K.A.; Laznik-Bogoslavski, D.; Vetrivelan, R.; Clish, C.B.; et al. Mitochondrial ROS regulate thermogenic energy expenditure and sulfenylation of UCP1. Nature 2016, 532, 112–116. [Google Scholar] [CrossRef] [Green Version]
  134. Shabalina, I.G.; Petrovic, N.; Kramarova, T.V.; Hoeks, J.; Cannon, B.; Nedergaard, J. UCP1 and defense against oxidative stress. 4-Hydroxy-2-nonenal effects on brown fat mitochondria are uncoupling protein 1-independent. J. Biol. Chem. 2006, 281, 13882–13893. [Google Scholar] [CrossRef] [Green Version]
  135. Shabalina, I.G.; Vrbacky, M.; Pecinova, A.; Kalinovich, A.V.; Drahota, Z.; Houstek, J.; Mracek, T.; Cannon, B.; Nedergaard, J. ROS production in brown adipose tissue mitochondria: The question of UCP1-dependence. Biochim. Biophys. Acta 2014, 1837, 2017–2030. [Google Scholar] [CrossRef] [Green Version]
  136. Jun, Z.; Ibrahim, M.M.; Dezheng, G.; Bo, Y.; Qiong, W.; Yuan, Z. UCP2 protects against amyloid beta toxicity and oxidative stress in primary neuronal culture. Biomed. Pharmacother. 2015, 74, 211–214. [Google Scholar] [CrossRef] [PubMed]
  137. Liu, Y.; Chen, L.; Xu, X.; Vicaut, E.; Sercombe, R. Both ischemic preconditioning and ghrelin administration protect hippocampus from ischemia/reperfusion and upregulate uncoupling protein-2. BMC Physiol. 2009, 9, 17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Chan, S.H.; Wu, C.A.; Wu, K.L.; Ho, Y.H.; Chang, A.Y.; Chan, J.Y. Transcriptional upregulation of mitochondrial uncoupling protein 2 protects against oxidative stress-associated neurogenic hypertension. Circ. Res. 2009, 105, 886–896. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Lee, S.C.; Robson-Doucette, C.A.; Wheeler, M.B. Uncoupling protein 2 regulates reactive oxygen species formation in islets and influences susceptibility to diabetogenic action of streptozotocin. J. Endocrinol. 2009, 203, 33–43. [Google Scholar] [CrossRef] [Green Version]
  140. Robson-Doucette, C.A.; Sultan, S.; Allister, E.M.; Wikstrom, J.D.; Koshkin, V.; Bhattacharjee, A.; Prentice, K.J.; Sereda, S.B.; Shirihai, O.S.; Wheeler, M.B. Beta-cell uncoupling protein 2 regulates reactive oxygen species production, which influences both insulin and glucagon secretion. Diabetes 2011, 60, 2710–2719. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  141. Pi, J.; Bai, Y.; Daniel, K.W.; Liu, D.; Lyght, O.; Edelstein, D.; Brownlee, M.; Corkey, B.E.; Collins, S. Persistent oxidative stress due to absence of uncoupling protein 2 associated with impaired pancreatic beta-cell function. Endocrinology 2009, 150, 3040–3048. [Google Scholar] [CrossRef] [Green Version]
  142. Blanc, J.; Alves-Guerra, M.C.; Esposito, B.; Rousset, S.; Gourdy, P.; Ricquier, D.; Tedgui, A.; Miroux, B.; Mallat, Z. Protective role of uncoupling protein 2 in atherosclerosis. Circulation 2003, 107, 388–390. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Derdak, Z.; Mark, N.M.; Beldi, G.; Robson, S.C.; Wands, J.R.; Baffy, G. The mitochondrial uncoupling protein-2 promotes chemoresistance in cancer cells. Cancer Res. 2008, 68, 2813–2819. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Baffy, G. Uncoupling protein-2 and cancer. Mitochondrion 2010, 10, 243–252. [Google Scholar] [CrossRef] [PubMed]
  145. Mailloux, R.J.; Adjeitey, C.N.; Harper, M.E. Genipin-induced inhibition of uncoupling protein-2 sensitizes drug-resistant cancer cells to cytotoxic agents. PLoS ONE 2010, 5, e13289. [Google Scholar] [CrossRef] [Green Version]
  146. Anderson, E.J.; Yamazaki, H.; Neufer, P.D. Induction of endogenous uncoupling protein 3 suppresses mitochondrial oxidant emission during fatty acid-supported respiration. J. Biol. Chem. 2007, 282, 31257–31266. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Talbot, D.A.; Brand, M.D. Uncoupling protein 3 protects aconitase against inactivation in isolated skeletal muscle mitochondria. Biochim. Biophys. Acta 2005, 1709, 150–156. [Google Scholar] [CrossRef] [Green Version]
  148. Echtay, K.S.; Murphy, M.P.; Smith, R.A.; Talbot, D.A.; Brand, M.D. Superoxide activates mitochondrial uncoupling protein 2 from the matrix side. Studies using targeted antioxidants. J. Biol. Chem. 2002, 277, 47129–47135. [Google Scholar] [CrossRef] [Green Version]
  149. Echtay, K.S.; Pakay, J.L.; Esteves, T.C.; Brand, M.D. Hydroxynonenal and uncoupling proteins: A model for protection against oxidative damage. Biofactors 2005, 24, 119–130. [Google Scholar] [CrossRef]
  150. Murphy, M.P.; Echtay, K.S.; Blaikie, F.H.; Asin-Cayuela, J.; Cocheme, H.M.; Green, K.; Buckingham, J.A.; Taylor, E.R.; Hurrell, F.; Hughes, G.; et al. Superoxide activates uncoupling proteins by generating carbon-centered radicals and initiating lipid peroxidation: Studies using a mitochondria-targeted spin trap derived from alpha-phenyl-N-tert-butylnitrone. J. Biol. Chem. 2003, 278, 48534–48545. [Google Scholar] [CrossRef] [Green Version]
  151. Toime, L.J.; Brand, M.D. Uncoupling protein-3 lowers reactive oxygen species production in isolated mitochondria. Free Radic. Biol. Med. 2010, 49, 606–611. [Google Scholar] [CrossRef] [Green Version]
  152. Ozcan, C.; Palmeri, M.; Horvath, T.L.; Russell, K.S.; Russell, R.R., 3rd. Role of uncoupling protein 3 in ischemia-reperfusion injury, arrhythmias, and preconditioning. Am. J. Physiol. Heart Circ. Physiol. 2013, 304, H1192–H1200. [Google Scholar] [CrossRef] [Green Version]
  153. MacLellan, J.D.; Gerrits, M.F.; Gowing, A.; Smith, P.J.; Wheeler, M.B.; Harper, M.E. Physiological increases in uncoupling protein 3 augment fatty acid oxidation and decrease reactive oxygen species production without uncoupling respiration in muscle cells. Diabetes 2005, 54, 2343–2350. [Google Scholar] [CrossRef] [Green Version]
  154. Jiang, N.; Zhang, G.; Bo, H.; Qu, J.; Ma, G.; Cao, D.; Wen, L.; Liu, S.; Ji, L.L.; Zhang, Y. Upregulation of uncoupling protein-3 in skeletal muscle during exercise: A potential antioxidant function. Free Radic. Biol. Med. 2009, 46, 138–145. [Google Scholar] [CrossRef]
  155. Nabben, M.; Hoeks, J.; Briede, J.J.; Glatz, J.F.; Moonen-Kornips, E.; Hesselink, M.K.; Schrauwen, P. The effect of UCP3 overexpression on mitochondrial ROS production in skeletal muscle of young versus aged mice. FEBS Lett. 2008, 582, 4147–4152. [Google Scholar] [CrossRef] [Green Version]
  156. Goglia, F.; Skulachev, V.P. A function for novel uncoupling proteins: Antioxidant defense of mitochondrial matrix by translocating fatty acid peroxides from the inner to the outer membrane leaflet. FASEB J. 2003, 17, 1585–1591. [Google Scholar] [CrossRef] [Green Version]
  157. Carrageta, D.F.; Oliveira, P.F.; Alves, M.G.; Monteiro, M.P. Obesity and male hypogonadism: Tales of a vicious cycle. Obes. Rev. 2019, 20, 1148–1158. [Google Scholar] [CrossRef] [PubMed]
  158. Betz, M.J.; Enerback, S. Targeting thermogenesis in brown fat and muscle to treat obesity and metabolic disease. Nat. Rev. Endocrinol. 2018, 14, 77–87. [Google Scholar] [CrossRef] [PubMed]
  159. Hosseini, B.; Mirzaei, K.; Maghbooli, Z.; Keshavarz, S.A.; Hossein-Nezhad, A. Compare the resting metabolic rate status in the healthy metabolically obese with the unhealthy metabolically obese participants. J. Nutr. Intermed. Metab. 2016, 6, 48–53. [Google Scholar] [CrossRef] [Green Version]
  160. Rolfe, D.F.; Brand, M.D. Contribution of mitochondrial proton leak to skeletal muscle respiration and to standard metabolic rate. Am. J. Physiol. 1996, 271, C1380–C1389. [Google Scholar] [CrossRef]
  161. Harper, M.E.; Green, K.; Brand, M.D. The efficiency of cellular energy transduction and its implications for obesity. Annu. Rev. Nutr. 2008, 28, 13–33. [Google Scholar] [CrossRef]
  162. Abumrad, N.A. The Liver as a Hub in Thermogenesis. Cell Metab. 2017, 26, 454–455. [Google Scholar] [CrossRef] [Green Version]
  163. Enerback, S.; Jacobsson, A.; Simpson, E.M.; Guerra, C.; Yamashita, H.; Harper, M.E.; Kozak, L.P. Mice lacking mitochondrial uncoupling protein are cold-sensitive but not obese. Nature 1997, 387, 90–94. [Google Scholar] [CrossRef]
  164. Liu, X.; Rossmeisl, M.; McClaine, J.; Riachi, M.; Harper, M.E.; Kozak, L.P. Paradoxical resistance to diet-induced obesity in UCP1-deficient mice. J. Clin. Investig. 2003, 111, 399–407. [Google Scholar] [CrossRef] [PubMed]
  165. Kontani, Y.; Wang, Y.; Kimura, K.; Inokuma, K.I.; Saito, M.; Suzuki-Miura, T.; Wang, Z.; Sato, Y.; Mori, N.; Yamashita, H. UCP1 deficiency increases susceptibility to diet-induced obesity with age. Aging Cell 2005, 4, 147–155. [Google Scholar] [CrossRef]
  166. Feldmann, H.M.; Golozoubova, V.; Cannon, B.; Nedergaard, J. UCP1 ablation induces obesity and abolishes diet-induced thermogenesis in mice exempt from thermal stress by living at thermoneutrality. Cell Metab. 2009, 9, 203–209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Rowland, L.A.; Maurya, S.K.; Bal, N.C.; Kozak, L.; Periasamy, M. Sarcolipin and uncoupling protein 1 play distinct roles in diet-induced thermogenesis and do not compensate for one another. Obesity 2016, 24, 1430–1433. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Von Essen, G.; Lindsund, E.; Cannon, B.; Nedergaard, J. Adaptive facultative diet-induced thermogenesis in wild-type but not in UCP1-ablated mice. Am. J. Physiol. Endocrinol. Metab. 2017, 313, E515–E527. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Luijten, I.H.N.; Feldmann, H.M.; von Essen, G.; Cannon, B.; Nedergaard, J. In the absence of UCP1-mediated diet-induced thermogenesis, obesity is augmented even in the obesity-resistant 129S mouse strain. Am. J. Physiol. Endocrinol. Metab. 2019, 316, E729–E740. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Almind, K.; Manieri, M.; Sivitz, W.I.; Cinti, S.; Kahn, C.R. Ectopic brown adipose tissue in muscle provides a mechanism for differences in risk of metabolic syndrome in mice. Proc. Natl. Acad. Sci. USA 2007, 104, 2366–2371. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Winn, N.C.; Vieira-Potter, V.J.; Gastecki, M.L.; Welly, R.J.; Scroggins, R.J.; Zidon, T.M.; Gaines, T.L.; Woodford, M.L.; Karasseva, N.G.; Kanaley, J.A.; et al. Loss of UCP1 exacerbates Western diet-induced glycemic dysregulation independent of changes in body weight in female mice. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2017, 312, R74–R84. [Google Scholar] [CrossRef]
  172. Chondronikola, M.; Volpi, E.; Borsheim, E.; Porter, C.; Annamalai, P.; Enerback, S.; Lidell, M.E.; Saraf, M.K.; Labbe, S.M.; Hurren, N.M.; et al. Brown adipose tissue improves whole-body glucose homeostasis and insulin sensitivity in humans. Diabetes 2014, 63, 4089–4099. [Google Scholar] [CrossRef] [Green Version]
  173. Cypess, A.M.; Lehman, S.; Williams, G.; Tal, I.; Rodman, D.; Goldfine, A.B.; Kuo, F.C.; Palmer, E.L.; Tseng, Y.H.; Doria, A.; et al. Identification and importance of brown adipose tissue in adult humans. N. Engl. J. Med. 2009, 360, 1509–1517. [Google Scholar] [CrossRef] [Green Version]
  174. Inokuma, K.; Ogura-Okamatsu, Y.; Toda, C.; Kimura, K.; Yamashita, H.; Saito, M. Uncoupling protein 1 is necessary for norepinephrine-induced glucose utilization in brown adipose tissue. Diabetes 2005, 54, 1385–1391. [Google Scholar] [CrossRef] [Green Version]
  175. Thoonen, R.; Ernande, L.; Cheng, J.; Nagasaka, Y.; Yao, V.; Miranda-Bezerra, A.; Chen, C.; Chao, W.; Panagia, M.; Sosnovik, D.E.; et al. Functional brown adipose tissue limits cardiomyocyte injury and adverse remodeling in catecholamine-induced cardiomyopathy. J. Mol. Cell. Cardiol. 2015, 84, 202–211. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Saito, M.; Matsushita, M.; Yoneshiro, T.; Okamatsu-Ogura, Y. Brown Adipose Tissue, Diet-Induced Thermogenesis, and Thermogenic Food Ingredients: From Mice to Men. Front. Endocrinol. 2020, 11, 222. [Google Scholar] [CrossRef] [Green Version]
  177. Fernandez Vazquez, G.; Reiter, R.J.; Agil, A. Melatonin increases brown adipose tissue mass and function in Zucker diabetic fatty rats: Implications for obesity control. J. Pineal Res. 2018, 64, e12472. [Google Scholar] [CrossRef] [PubMed]
  178. Kopecky, J.; Hodny, Z.; Rossmeisl, M.; Syrovy, I.; Kozak, L.P. Reduction of dietary obesity in aP2-Ucp transgenic mice: Physiology and adipose tissue distribution. Am. J. Physiol. 1996, 270, E768–E775. [Google Scholar] [CrossRef] [PubMed]
  179. Klaus, S.; Keipert, S.; Rossmeisl, M.; Kopecky, J. Augmenting energy expenditure by mitochondrial uncoupling: A role of AMP-activated protein kinase. Genes Nutr. 2012, 7, 369–386. [Google Scholar] [CrossRef] [Green Version]
  180. Li, B.; Nolte, L.A.; Ju, J.S.; Han, D.H.; Coleman, T.; Holloszy, J.O.; Semenkovich, C.F. Skeletal muscle respiratory uncoupling prevents diet-induced obesity and insulin resistance in mice. Nat. Med. 2000, 6, 1115–1120. [Google Scholar] [CrossRef] [PubMed]
  181. Adjeitey, C.N.; Mailloux, R.J.; Dekemp, R.A.; Harper, M.E. Mitochondrial uncoupling in skeletal muscle by UCP1 augments energy expenditure and glutathione content while mitigating ROS production. Am. J. Physiol. Endocrinol. Metab. 2013, 305, E405–E415. [Google Scholar] [CrossRef] [Green Version]
  182. Keipert, S.; Ost, M.; Chadt, A.; Voigt, A.; Ayala, V.; Portero-Otin, M.; Pamplona, R.; Al-Hasani, H.; Klaus, S. Skeletal muscle uncoupling-induced longevity in mice is linked to increased substrate metabolism and induction of the endogenous antioxidant defense system. Am. J. Physiol. Endocrinol. Metab. 2013, 304, E495–E506. [Google Scholar] [CrossRef] [Green Version]
  183. Neschen, S.; Katterle, Y.; Richter, J.; Augustin, R.; Scherneck, S.; Mirhashemi, F.; Schurmann, A.; Joost, H.G.; Klaus, S. Uncoupling protein 1 expression in murine skeletal muscle increases AMPK activation, glucose turnover, and insulin sensitivity in vivo. Physiol. Genom. 2008, 33, 333–340. [Google Scholar] [CrossRef]
  184. Ost, M.; Werner, F.; Dokas, J.; Klaus, S.; Voigt, A. Activation of AMPKalpha2 is not crucial for mitochondrial uncoupling-induced metabolic effects but required to maintain skeletal muscle integrity. PLoS ONE 2014, 9, e94689. [Google Scholar] [CrossRef] [PubMed]
  185. Hong, Y.; Fink, B.D.; Dillon, J.S.; Sivitz, W.I. Effects of adenoviral overexpression of uncoupling protein-2 and -3 on mitochondrial respiration in insulinoma cells. Endocrinology 2001, 142, 249–256. [Google Scholar] [CrossRef]
  186. Zhang, C.Y.; Parton, L.E.; Ye, C.P.; Krauss, S.; Shen, R.; Lin, C.T.; Porco, J.A., Jr.; Lowell, B.B. Genipin inhibits UCP2-mediated proton leak and acutely reverses obesity- and high glucose-induced beta cell dysfunction in isolated pancreatic islets. Cell Metab. 2006, 3, 417–427. [Google Scholar] [CrossRef] [Green Version]
  187. De Souza, C.T.; Araujo, E.P.; Stoppiglia, L.F.; Pauli, J.R.; Ropelle, E.; Rocco, S.A.; Marin, R.M.; Franchini, K.G.; Carvalheira, J.B.; Saad, M.J.; et al. Inhibition of UCP2 expression reverses diet-induced diabetes mellitus by effects on both insulin secretion and action. FASEB J. 2007, 21, 1153–1163. [Google Scholar] [CrossRef] [Green Version]
  188. Saleh, M.C.; Wheeler, M.B.; Chan, C.B. Endogenous islet uncoupling protein-2 expression and loss of glucose homeostasis in ob/ob mice. J. Endocrinol. 2006, 190, 659–667. [Google Scholar] [CrossRef]
  189. Pi, J.; Collins, S. Reactive oxygen species and uncoupling protein 2 in pancreatic beta-cell function. Diabetes Obes. Metab. 2010, 12 (Suppl. 2), 141–148. [Google Scholar] [CrossRef]
  190. Affourtit, C.; Brand, M.D. Uncoupling protein-2 contributes significantly to high mitochondrial proton leak in INS-1E insulinoma cells and attenuates glucose-stimulated insulin secretion. Biochem. J. 2008, 409, 199–204. [Google Scholar] [CrossRef] [PubMed]
  191. Pi, J.; Bai, Y.; Zhang, Q.; Wong, V.; Floering, L.M.; Daniel, K.; Reece, J.M.; Deeney, J.T.; Andersen, M.E.; Corkey, B.E.; et al. Reactive oxygen species as a signal in glucose-stimulated insulin secretion. Diabetes 2007, 56, 1783–1791. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Chen, J.; Stimpson, S.E.; Fernandez-Bueno, G.A.; Mathews, C.E. Mitochondrial Reactive Oxygen Species and Type 1 Diabetes. Antioxid. Redox Signal. 2018, 29, 1361–1372. [Google Scholar] [CrossRef] [PubMed]
  193. Diano, S.; Horvath, T.L. Mitochondrial uncoupling protein 2 (UCP2) in glucose and lipid metabolism. Trends Mol. Med. 2012, 18, 52–58. [Google Scholar] [CrossRef] [PubMed]
  194. Affourtit, C.; Jastroch, M.; Brand, M.D. Uncoupling protein-2 attenuates glucose-stimulated insulin secretion in INS-1E insulinoma cells by lowering mitochondrial reactive oxygen species. Free Radic. Biol. Med. 2011, 50, 609–616. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Clapham, J.C.; Arch, J.R.; Chapman, H.; Haynes, A.; Lister, C.; Moore, G.B.; 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]
  196. Son, C.; Hosoda, K.; Ishihara, K.; Bevilacqua, L.; Masuzaki, H.; Fushiki, T.; Harper, M.E.; Nakao, K. Reduction of diet-induced obesity in transgenic mice overexpressing uncoupling protein 3 in skeletal muscle. Diabetologia 2004, 47, 47–54. [Google Scholar] [CrossRef] [PubMed]
  197. Costford, S.R.; Chaudhry, S.N.; Salkhordeh, M.; Harper, M.E. Effects of the presence, absence, and overexpression of uncoupling protein-3 on adiposity and fuel metabolism in congenic mice. Am. J. Physiol. Endocrinol. Metab. 2006, 290, E1304–E1312. [Google Scholar] [CrossRef] [Green Version]
  198. Aguer, C.; Fiehn, O.; Seifert, E.L.; Bezaire, V.; Meissen, J.K.; Daniels, A.; Scott, K.; Renaud, J.M.; Padilla, M.; Bickel, D.R.; et al. Muscle uncoupling protein 3 overexpression mimics endurance training and reduces circulating biomarkers of incomplete beta-oxidation. FASEB J. 2013, 27, 4213–4225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Gong, D.W.; Monemdjou, S.; Gavrilova, O.; Leon, L.R.; Marcus-Samuels, B.; Chou, C.J.; Everett, C.; Kozak, L.P.; Li, C.; Deng, C.; et al. Lack of obesity and normal response to fasting and thyroid hormone in mice lacking uncoupling protein-3. J. Biol. Chem. 2000, 275, 16251–16257. [Google Scholar] [CrossRef] [Green Version]
  200. Lomax, T.M.; Ashraf, S.; Yilmaz, G.; Harmancey, R. Loss of Uncoupling Protein 3 Attenuates Western Diet-Induced Obesity, Systemic Inflammation, and Insulin Resistance in Rats. Obesity 2020, 28, 1687–1697. [Google Scholar] [CrossRef]
  201. Choi, C.S.; Fillmore, J.J.; Kim, J.K.; Liu, Z.X.; Kim, S.; Collier, E.F.; Kulkarni, A.; Distefano, A.; Hwang, Y.J.; Kahn, M.; et al. Overexpression of uncoupling protein 3 in skeletal muscle protects against fat-induced insulin resistance. J. Clin. Investig. 2007, 117, 1995–2003. [Google Scholar] [CrossRef] [Green Version]
  202. Costford, S.R.; Chaudhry, S.N.; Crawford, S.A.; Salkhordeh, M.; Harper, M.E. Long-term high-fat feeding induces greater fat storage in mice lacking UCP3. Am. J. Physiol. Endocrinol. Metab. 2008, 295, E1018–E1024. [Google Scholar] [CrossRef]
  203. Senese, R.; Valli, V.; Moreno, M.; Lombardi, A.; Busiello, R.A.; Cioffi, F.; Silvestri, E.; Goglia, F.; Lanni, A.; de Lange, P. Uncoupling protein 3 expression levels influence insulin sensitivity, fatty acid oxidation, and related signaling pathways. Pflugers Arch. 2011, 461, 153–164. [Google Scholar] [CrossRef] [PubMed]
  204. Schrauwen, P.; Hesselink, M.K.; Blaak, E.E.; Borghouts, L.B.; Schaart, G.; Saris, W.H.; Keizer, H.A. Uncoupling protein 3 content is decreased in skeletal muscle of patients with type 2 diabetes. Diabetes 2001, 50, 2870–2873. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Schrauwen, P.; Mensink, M.; Schaart, G.; Moonen-Kornips, E.; Sels, J.P.; Blaak, E.E.; Russell, A.P.; Hesselink, M.K. Reduced skeletal muscle uncoupling protein-3 content in prediabetic subjects and type 2 diabetic patients: Restoration by rosiglitazone treatment. J. Clin. Endocrinol. Metab. 2006, 91, 1520–1525. [Google Scholar] [CrossRef] [Green Version]
  206. Mensink, M.; Hesselink, M.K.; Borghouts, L.B.; Keizer, H.; Moonen-Kornips, E.; Schaart, G.; Blaak, E.E.; Schrauwen, P. Skeletal muscle uncoupling protein-3 restores upon intervention in the prediabetic and diabetic state: Implications for diabetes pathogenesis? Diabetes Obes. Metab. 2007, 9, 594–596. [Google Scholar] [CrossRef] [PubMed]
  207. MacLeod, J. The role of oxygen in the metabolism and motility of human spermatozoa. Am. J. Physiol. 1943, 138, 0512–0518. [Google Scholar] [CrossRef] [Green Version]
  208. Tremellen, K. Oxidative stress and male infertility—A clinical perspective. Hum. Reprod. Update 2008, 14, 243–258. [Google Scholar] [CrossRef] [PubMed]
  209. Shekarriz, M.; Thomas, A.J., Jr.; Agarwal, A. Incidence and level of seminal reactive oxygen species in normal men. Urology 1995, 45, 103–107. [Google Scholar] [CrossRef]
  210. Ochsendorf, F.R.; Thiele, J.; Fuchs, J.; Schuttau, H.; Freisleben, H.J.; Buslau, M.; Milbradt, R. Chemiluminescence in semen of infertile men. Andrologia 1994, 26, 289–293. [Google Scholar] [CrossRef]
  211. Zini, A.; de Lamirande, E.; Gagnon, C. Reactive oxygen species in semen of infertile patients: Levels of superoxide dismutase- and catalase-like activities in seminal plasma and spermatozoa. Int. J. Androl. 1993, 16, 183–188. [Google Scholar] [CrossRef]
  212. Iwasaki, A.; Gagnon, C. Formation of reactive oxygen species in spermatozoa of infertile patients. Fertil. Steril. 1992, 57, 409–416. [Google Scholar] [CrossRef]
  213. Agarwal, A.; Said, T.M.; Bedaiwy, M.A.; Banerjee, J.; Alvarez, J.G. Oxidative stress in an assisted reproductive techniques setting. Fertil. Steril. 2006, 86, 503–512. [Google Scholar] [CrossRef]
  214. Agarwal, A.; Sharma, R.K.; Sharma, R.; Assidi, M.; Abuzenadah, A.M.; Alshahrani, S.; Durairajanayagam, D.; Sabanegh, E. Characterizing semen parameters and their association with reactive oxygen species in infertile men. Reprod. Biol. Endocrinol. 2014, 12, 33. [Google Scholar] [CrossRef] [Green Version]
  215. Agarwal, A.; Saleh, R.A.; Bedaiwy, M.A. Role of reactive oxygen species in the pathophysiology of human reproduction. Fertil. Steril. 2003, 79, 829–843. [Google Scholar] [CrossRef] [Green Version]
  216. Said, T.M.; Agarwal, A.; Sharma, R.K.; Mascha, E.; Sikka, S.C.; Thomas, A.J., Jr. Human sperm superoxide anion generation and correlation with semen quality in patients with male infertility. Fertil. Steril. 2004, 82, 871–877. [Google Scholar] [CrossRef] [PubMed]
  217. Aziz, N.; Saleh, R.A.; Sharma, R.K.; Lewis-Jones, I.; Esfandiari, N.; Thomas, A.J., Jr.; Agarwal, A. Novel association between sperm reactive oxygen species production, sperm morphological defects, and the sperm deformity index. Fertil. Steril. 2004, 81, 349–354. [Google Scholar] [CrossRef]
  218. Moustafa, M.H.; Sharma, R.K.; Thornton, J.; Mascha, E.; Abdel-Hafez, M.A.; Thomas, A.J., Jr.; Agarwal, A. Relationship between ROS production, apoptosis and DNA denaturation in spermatozoa from patients examined for infertility. Hum. Reprod. 2004, 19, 129–138. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  219. Smith, R.; Kaune, H.; Parodi, D.; Madariaga, M.; Rios, R.; Morales, I.; Castro, A. Increased sperm DNA damage in patients with varicocele: Relationship with seminal oxidative stress. Hum. Reprod. 2006, 21, 986–993. [Google Scholar] [CrossRef] [Green Version]
  220. De Lamirande, E.; Jiang, H.; Zini, A.; Kodama, H.; Gagnon, C. Reactive oxygen species and sperm physiology. Rev. Reprod. 1997, 2, 48–54. [Google Scholar] [CrossRef]
  221. Du Plessis, S.S.; Agarwal, A.; Halabi, J.; Tvrda, E. Contemporary evidence on the physiological role of reactive oxygen species in human sperm function. J. Assist. Reprod. Genet. 2015, 32, 509–520. [Google Scholar] [CrossRef] [Green Version]
  222. De Lamirande, E.; Gagnon, C. Human sperm hyperactivation and capacitation as parts of an oxidative process. Free Radic. Biol. Med. 1993, 14, 157–166. [Google Scholar] [CrossRef]
  223. Aitken, R.J.; Irvine, D.S.; Wu, F.C. Prospective analysis of sperm-oocyte fusion and reactive oxygen species generation as criteria for the diagnosis of infertility. Am. J. Obstet. Gynecol. 1991, 164, 542–551. [Google Scholar] [CrossRef]
  224. Dias, T.R.; Martin-Hidalgo, D.; Silva, B.M.; Oliveira, P.F.; Alves, M.G. Endogenous and exogenous antioxidants as a tool to ameliorate male infertility induced by reactive oxygen species. Antioxid. Redox Signal. 2020, 33, 767–785. [Google Scholar] [CrossRef] [PubMed]
  225. Lettieri, G.; d’Agostino, G.; Mele, E.; Cardito, C.; Esposito, R.; Cimmino, A.; Giarra, A.; Trifuoggi, M.; Raimondo, S.; Notari, T.; et al. Discovery of the Involvement in DNA Oxidative Damage of Human Sperm Nuclear Basic Proteins of Healthy Young Men Living in Polluted Areas. Int. J. Mol. Sci. 2020, 21, 4198. [Google Scholar] [CrossRef]
  226. Lettieri, G.; Marra, F.; Moriello, C.; Prisco, M.; Notari, T.; Trifuoggi, M.; Giarra, A.; Bosco, L.; Montano, L.; Piscopo, M. Molecular alterations in spermatozoa of a family case living in the land of fires. A first look at possible transgenerational effects of pollutants. Int. J. Mol. Sci. 2020, 21, 6710. [Google Scholar] [CrossRef] [PubMed]
  227. Martin-Hidalgo, D.; Bragado, M.J.; Batista, A.R.; Oliveira, P.F.; Alves, M.G. Antioxidants and Male Fertility: From Molecular Studies to Clinical Evidence. Antioxidants 2019, 8, 89. [Google Scholar] [CrossRef] [Green Version]
  228. Athayde, K.S.; Cocuzza, M.; Agarwal, A.; Krajcir, N.; Lucon, A.M.; Srougi, M.; Hallak, J. Development of normal reference values for seminal reactive oxygen species and their correlation with leukocytes and semen parameters in a fertile population. J. Androl. 2007, 28, 613–620. [Google Scholar] [CrossRef]
  229. Hosseinzadeh Colagar, A.; Karimi, F.; Jorsaraei, S.G. Correlation of sperm parameters with semen lipid peroxidation and total antioxidants levels in astheno- and oligoasheno-teratospermic men. Iran. Red Crescent Med. J. 2013, 15, 780–785. [Google Scholar] [CrossRef] [Green Version]
  230. De Lamirande, E.; Gagnon, C. Reactive oxygen species and human spermatozoa. I. Effects on the motility of intact spermatozoa and on sperm axonemes. J. Androl. 1992, 13, 368–378. [Google Scholar]
  231. De Lamirande, E.; Gagnon, C. Reactive oxygen species and human spermatozoa. II. Depletion of adenosine triphosphate plays an important role in the inhibition of sperm motility. J. Androl. 1992, 13, 379–386. [Google Scholar] [PubMed]
  232. Kurkowska, W.; Bogacz, A.; Janiszewska, M.; Gabrys, E.; Tiszler, M.; Bellanti, F.; Kasperczyk, S.; Machon-Grecka, A.; Dobrakowski, M.; Kasperczyk, A. Oxidative Stress is Associated with Reduced Sperm Motility in Normal Semen. Am. J. Men’s Health 2020, 14, 1557988320939731. [Google Scholar] [CrossRef]
  233. Lewis, S.E.; Simon, L. Clinical implications of sperm DNA damage. Hum. Fertil. (Camb.) 2010, 13, 201–207. [Google Scholar] [CrossRef]
  234. Tarozzi, N.; Bizzaro, D.; Flamigni, C.; Borini, A. Clinical relevance of sperm DNA damage in assisted reproduction. Reprod. Biomed. Online 2007, 14, 746–757. [Google Scholar] [CrossRef]
  235. Cissen, M.; Wely, M.V.; Scholten, I.; Mansell, S.; Bruin, J.P.; Mol, B.W.; Braat, D.; Repping, S.; Hamer, G. Measuring Sperm DNA Fragmentation and Clinical Outcomes of Medically Assisted Reproduction: A Systematic Review and Meta-Analysis. PLoS ONE 2016, 11, e0165125. [Google Scholar] [CrossRef] [Green Version]
  236. La Vignera, S.; Condorelli, R.; Vicari, E.; D’Agata, R.; Calogero, A.E. Diabetes mellitus and sperm parameters. J. Androl. 2012, 33, 145–153. [Google Scholar] [CrossRef] [PubMed]
  237. Oliveira, P.F.; Sousa, M.; Silva, B.M.; Monteiro, M.P.; Alves, M.G. Obesity, energy balance and spermatogenesis. Reproduction 2017, 153, R173–R185. [Google Scholar] [CrossRef] [Green Version]
  238. Garolla, A.; Torino, M.; Miola, P.; Caretta, N.; Pizzol, D.; Menegazzo, M.; Bertoldo, A.; Foresta, C. Twenty-four-hour monitoring of scrotal temperature in obese men and men with a varicocele as a mirror of spermatogenic function. Hum. Reprod. 2015, 30, 1006–1013. [Google Scholar] [CrossRef]
  239. Mieusset, R.; Bujan, L.; Mondinat, C.; Mansat, A.; Pontonnier, F.; Grandjean, H. Association of scrotal hyperthermia with impaired spermatogenesis in infertile men. Fertil. Steril. 1987, 48, 1006–1011. [Google Scholar] [CrossRef]
  240. Håkonsen, L.B.; Thulstrup, A.M.; Aggerholm, A.S.; Olsen, J.; Bonde, J.P.; Andersen, C.Y.; Bungum, M.; Ernst, E.H.; Hansen, M.L.; Ernst, E.H. Does weight loss improve semen quality and reproductive hormones? Results from a cohort of severely obese men. Reprod. Health 2011, 8, 1–8. [Google Scholar] [CrossRef] [Green Version]
  241. Crisóstomo, L.; Videira, R.A.; Jarak, I.; Starčević, K.; Mašek, T.; Rato, L.; Raposo, J.F.; Batterham, R.L.; Oliveira, P.F.; Alves, M.G. Diet during early life defines testicular lipid content and sperm quality in adulthood. Am. J. Physiol.-Endocrinol. Metab. 2020, 319, E1061–E1073. [Google Scholar] [CrossRef] [PubMed]
  242. Crisóstomo, L.; Rato, L.; Jarak, I.; Silva, B.M.; Raposo, J.F.; Batterham, R.L.; Oliveira, P.F.; Alves, M.G. A switch from high-fat to normal diet does not restore sperm quality but prevents metabolic syndrome. Reproduction 2019, 158, 377–387. [Google Scholar] [CrossRef]
  243. Maresch, C.C.; Stute, D.C.; Alves, M.G.; Oliveira, P.F.; de Kretser, D.M.; Linn, T. Diabetes-induced hyperglycemia impairs male reproductive function: A systematic review. Hum. Reprod. Update 2018, 24, 86–105. [Google Scholar] [CrossRef] [Green Version]
  244. Alves, M.G.; Martins, A.D.; Rato, L.; Moreira, P.I.; Socorro, S.; Oliveira, P.F. Molecular mechanisms beyond glucose transport in diabetes-related male infertility. Biochim. Biophys. Acta 2013, 1832, 626–635. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Leisegang, K.; Sengupta, P.; Agarwal, A.; Henkel, R. Obesity and male infertility: Mechanisms and management. Andrologia 2021, 53, e13617. [Google Scholar] [CrossRef] [PubMed]
  246. Amaral, S.; Oliveira, P.J.; Ramalho-Santos, J. Diabetes and the impairment of reproductive function: Possible role of mitochondria and reactive oxygen species. Curr. Diabetes Rev. 2008, 4, 46–54. [Google Scholar] [CrossRef]
  247. Kreiter, J.; Rupprecht, A.; Zimmermann, L.; Moschinger, M.; Rokitskaya, T.I.; Antonenko, Y.N.; Gille, L.; Fedorova, M.; Pohl, E.E. Molecular Mechanisms Responsible for Pharmacological Effects of Genipin on Mitochondrial Proteins. Biophys. J. 2019, 117, 1845–1857. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Proton transport and formation of ROS during oxidative phosphorylation. (A) In complex I, the oxidation of NADH (generated through glycolysis, β-oxidation, and Krebs cycle) to NAD+ transfers two electrons to the complex. Complex I will then transfer the electrons to the quinone reservoir by reducing ubiquinone to ubiquinol. In complex II occurs the oxidation of succinate to fumarate, a reaction that reduces FAD to FADH2. FADH2 also gives two electrons to ubiquinone, originating FAD and ubiquinol. Ubiquinol is released into the quinone reservoir, joining those from NADH. Ubiquinol transports electrons trough the intermembrane space to complex III, where they are again oxidized to quinones. Complex III transfers electrons to cytochrome c, which moves to complex IV. Complex IV receives the electrons from cytochrome c and transfers them to molecular oxygen, reducing it to water. Same electrons escape to oxygen, which lead to the formation of superoxide anions. Superoxide leads to a cascade of redox reactions, which leads to the formation of other ROS, such as the hydroxyl radical (OH) and hydrogen peroxide (H2O2). During the transport of electrons through the respiratory chain, protons are pumped from the matrix into the intermembrane space by complexes I, II, and IV. This action creates an electrochemical gradient that is used to (B) convert ADP into ATP, by ATP synthase. The proton transport to the mitochondrial matrix can be also made through (C) basal proton leak or (D) inducible proton leak.
Figure 1. Proton transport and formation of ROS during oxidative phosphorylation. (A) In complex I, the oxidation of NADH (generated through glycolysis, β-oxidation, and Krebs cycle) to NAD+ transfers two electrons to the complex. Complex I will then transfer the electrons to the quinone reservoir by reducing ubiquinone to ubiquinol. In complex II occurs the oxidation of succinate to fumarate, a reaction that reduces FAD to FADH2. FADH2 also gives two electrons to ubiquinone, originating FAD and ubiquinol. Ubiquinol is released into the quinone reservoir, joining those from NADH. Ubiquinol transports electrons trough the intermembrane space to complex III, where they are again oxidized to quinones. Complex III transfers electrons to cytochrome c, which moves to complex IV. Complex IV receives the electrons from cytochrome c and transfers them to molecular oxygen, reducing it to water. Same electrons escape to oxygen, which lead to the formation of superoxide anions. Superoxide leads to a cascade of redox reactions, which leads to the formation of other ROS, such as the hydroxyl radical (OH) and hydrogen peroxide (H2O2). During the transport of electrons through the respiratory chain, protons are pumped from the matrix into the intermembrane space by complexes I, II, and IV. This action creates an electrochemical gradient that is used to (B) convert ADP into ATP, by ATP synthase. The proton transport to the mitochondrial matrix can be also made through (C) basal proton leak or (D) inducible proton leak.
Antioxidants 10 01746 g001
Figure 2. Molecular structure and hypothesized models for proton transport by UCPs. (A) UCPs are constituted by three repetition domains, each composed of two α-helix regions. (B) The proposed models for proton transport by UCPs are (a) flip-flop model; (b) cofactor model (transport linked to fatty acids); (c) cofactor model (independent transport).
Figure 2. Molecular structure and hypothesized models for proton transport by UCPs. (A) UCPs are constituted by three repetition domains, each composed of two α-helix regions. (B) The proposed models for proton transport by UCPs are (a) flip-flop model; (b) cofactor model (transport linked to fatty acids); (c) cofactor model (independent transport).
Antioxidants 10 01746 g002
Table 1. Expression and hypothesized functions of all mitochondrial uncoupling proteins (UCPs) homologs.
Table 1. Expression and hypothesized functions of all mitochondrial uncoupling proteins (UCPs) homologs.
IsoformLocalizationPutative FunctionReferences
UCP1Human brown adipose tissue, white adipose tissue, keratinocytes, sweat glands, sebum glands, hair follicles, and granular layer of the epidermisNon-shivering thermogenesis, metabolic and bioenergetic regulation, modulation of ROS production[21,22,23]
Mouse brown adipose tissue, white adipose tissue, and adrenal gland[24,25,26]
Rat brown adipose tissue and white adipose tissue[27,28]
UCP2Human proximal tubular cells, cytotrophoblasts, syncytiotrophoblast, lungs *, keratinocytes *, skin fibroblasts *, spermatozoa, white adipose tissue *, skeletal muscle *, brown adipose tissue *, and pancreatic β-cells *, brainModulation of ROS production, insulin sensitivity and secretion, lipid and glucose metabolism, control of mitochondrial Ca2+-uptake, inflammation, immunomodulation[23,29,30,31,32,33,34,35]
Mouse thymocytes, tubular epithelial cells, granulosa cells *, theca cells *, endometrium glandular epithelium cells *, uterine glands *, oviduct mucosa epithelial cells *, mammary gland, stomach, macrophages, splenocytes, B lymphocytes, T lymphocytes, dendritic cells, neutrophils, microglial cells *, neurons *, cardiomyocytes, lungs, testicular germ cells, testicular interstitial cells, white adipose tissue *, skeletal muscle *, brown adipose tissue *, pancreatic β-cells *, pancreatic α-cells, and hepatocytes *[13,34,35,36,37,38,39,40,41,42,43,44,45,46]
Rat proximal tubular cells, Kupffer cells *, microglial cells, neurons, cardiomyocytes, skeletal muscle *, testis, brown adipose tissue *, white adipose tissue *, pancreatic β-cells, lungs *, spleen *, and thymus *[33,47,48,49,50,51,52,53,54,55,56]
UCP3Human skeletal muscle, heart *, spleen *, thymus *, keratinocytes *, skin fibroblasts, sweat glands, hair follicles, stratum basale of the epidermis, pancreatic β-cells, thyroid *, and bone marrow *Modulation of ROS production, insulin secretion, lipid metabolism, protection against lipotoxicity, control of mitochondrial Ca2+-uptake, immunomodulation[23,33,57,58,59,60]
Mouse skeletal muscle, brown adipose tissue, brain *, kidney *, colon *, liver *, heart, white adipose tissue, thymocytes, splenic lymphocytes, and peripheral naive CD4+ T cells *[60,61,62,63]
Rat skeletal muscle, brown adipose tissue *, white adipose tissue *, kidney *, spleen reticulocytes, spleen monocytes, spleen lymphocytes, thymocytes, and heart[33,52,57,64]
UCP4Human Purkinje cells, most of the brain tissues *, and cartilage *Regulation of oxidative stress, thermoregulation, protection against mitochondrial Ca2+ overload[7,9,15,65]
Rat Merkel cells, modiolus ear (fibrocyte, satellite cells), organ of Corti (spiral ganglion neurons, supporting cells, hair cells), brain (pyramidal cells of hippocampus and cortex, Purkinje cells of cerebellum, neurons of hippocampus, substantia nigra, striatum, neocortex *), chondrocytes, vestibular ganglion *, heart *, lungs *, skeletal muscle *, kidney *, and liver *[7,55,65,66,67,68,69]
Mouse brain (neurons, astrocytes, cortex, brainstem, cerebellum), spinal cord, mast cells, spiral ganglion *, vestibular ganglion *, kidney *, heart *, white adipose tissue *, brown adipose tissue *, skeletal muscle *, and liver *[55,66,70,71,72]
UCP5Human kidney *, heart *, lungs *, stomach *, liver *, spleen *, skeletal muscle *, brain (cerebellum, cortex, medulla, occipital pole, frontal lobe, putamen, amygdala, caudate nucleus, hippocampus, substantia nigra, thalamus, corpus callosum) *, spinal cord *, pituitary *, uterus *, and testis *Regulation of oxidative stress, thermoregulation, transport of metabolites, regulation of mitochondrial metabolism[14,73,74]
Rat heart *, lungs *, adrenals *, kidney *, gonadal fat *, ovary *, brain (striatum, cortex, hippocampus, cerebellum) *, skeletal muscle *, and liver *[14,55,75]
Mouse heart, kidney, skeletal muscle, white adipose tissue, spinal cord, brain (neurons, astrocytes, cortex, thalamus, hippocampus, substantia nigra, cerebellum, basal ganglia, hypothalamus *, amygdala *, neocortex *, caudate putamen *), spiral ganglion *, vestibular ganglion *, hepatocytes*, brown adipose tissue *, spleen *, intestine *, lungs *, testis *, uterus *, and periovarian fat *[14,30,44,71,72,74,76,77,78]
UCP6Mouse kidney (proximal tubules, distal tubules, surrounding nephron segments, glomeruli, medullary part of the loop of Henle, collecting duct) *, white adipose tissue *, brown adipose tissue *, brain *, heart *, muscle *, liver *, lungs *, spleen *, and testis *Regulation of oxidative stress, transport of metabolites, regulation of mitochondrial metabolism[30,74]
* Only mRNA has been reported or predicted based on mRNA microarrays.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Monteiro, B.S.; Freire-Brito, L.; Carrageta, D.F.; Oliveira, P.F.; Alves, M.G. Mitochondrial Uncoupling Proteins (UCPs) as Key Modulators of ROS Homeostasis: A Crosstalk between Diabesity and Male Infertility? Antioxidants 2021, 10, 1746. https://doi.org/10.3390/antiox10111746

AMA Style

Monteiro BS, Freire-Brito L, Carrageta DF, Oliveira PF, Alves MG. Mitochondrial Uncoupling Proteins (UCPs) as Key Modulators of ROS Homeostasis: A Crosstalk between Diabesity and Male Infertility? Antioxidants. 2021; 10(11):1746. https://doi.org/10.3390/antiox10111746

Chicago/Turabian Style

Monteiro, Bruno S., Laís Freire-Brito, David F. Carrageta, Pedro F. Oliveira, and Marco G. Alves. 2021. "Mitochondrial Uncoupling Proteins (UCPs) as Key Modulators of ROS Homeostasis: A Crosstalk between Diabesity and Male Infertility?" Antioxidants 10, no. 11: 1746. https://doi.org/10.3390/antiox10111746

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