Hostname: page-component-8448b6f56d-c4f8m Total loading time: 0 Render date: 2024-04-16T09:41:08.852Z Has data issue: false hasContentIssue false

Spermatic mitochondria: role in oxidative homeostasis, sperm function and possible tools for their assessment

Published online by Cambridge University Press:  18 September 2018

João Diego de Agostini Losano*
Affiliation:
Department of Animal Reproduction, College of Veterinary Medicine and Animal Science, University of São Paulo, Brazil
Daniel de Souza Ramos Angrimani
Affiliation:
Department of Animal Reproduction, College of Veterinary Medicine and Animal Science, University of São Paulo, Brazil
Roberta Ferreira Leite
Affiliation:
Department of Animal Reproduction, College of Veterinary Medicine and Animal Science, University of São Paulo, Brazil
Bárbara do Carmo Simões da Silva
Affiliation:
Department of Animal Reproduction, College of Veterinary Medicine and Animal Science, University of São Paulo, Brazil
Valquíria Hyppolito Barnabe
Affiliation:
Department of Animal Reproduction, College of Veterinary Medicine and Animal Science, University of São Paulo, Brazil
Marcilio Nichi
Affiliation:
Department of Animal Reproduction, College of Veterinary Medicine and Animal Science, University of São Paulo, Brazil
*
Author for correspondence: Department of Animal Reproduction, College of Veterinary Medicine and Animal Science, University of São Paulo, Av. Prof. Orlando Marques de Paiva, 87-05508-270, São Paulo, Brazil. Tel: +55 11 30911423. Fax: +55 11 30911437. E-mail: jdalosano@usp.br
Rights & Permissions [Opens in a new window]

Summary

Despite sperm mitochondrial relevance to the fertilization capacity, the processes involved in the production of ATP and functional dynamics of sperm mitochondria are not fully understood. One of these processes is the paradox involved between function and formation of reactive oxygen species performed by the organelle. Therefore, this review aimed to provide data on the role of sperm mitochondria in oxidative homeostasis and functionality as well the tools to assess sperm mitochondrial function.

Type
Review Article
Copyright
© Cambridge University Press 2018 

Introduction

Nuclear power, including the production of low-carbon electricity, is responsible at this time for almost one-third of global total energy generation. According to the International Atomic Energy Agency (IAEA), the main reason for this use is the low investment cost relative to the amount of energy produced and the fact that nuclear power plants produce virtually no greenhouse gas emissions or air pollutants during their operation, and very low emissions over their entire life cycle. Therefore, due to these benefits, there are now 447 nuclear power reactors in operation in 30 countries and another 58 reactors are under construction, based on IAEA information. However, the use of nuclear energy remains a cause of concern around the world, due to the devastating effects of accidents caused by core damage in nuclear plants. This concern is based on the long-term consequences of accidents, such as the accidents at Chernobyl (1986) and Fukushima-Daiichi (2011).

As in nuclear power plants, mitochondria exhibit high energy production capacities. However, in situations in which the structure of this organelle is compromised, the potential to release extremely toxic products is also worrying. Such toxic substances may lead to damage of the surrounding cells and other tissues. In fact, several studies have linked mitochondrial dysfunction to some pathological conditions such as neurodegenerative diseases (Lin & Beal, Reference Lin and Beal2006), type 2 diabetes (Lowell & Shulman, Reference Lowell and Shulman2005) and neoplasia (Modica-Napolitano & Singh, Reference Modica-Napolitano and Singh2004).

In regards to the spermatozoa, several studies related mitochondria to be the main source of cell energy, playing an important role in cellular homeostasis and sperm motility (Travis et al., Reference Travis, Foster, Rosenbaum, Visconti, Gerton, Kopf and Moss1998; St John, Reference St John2002). However, for some species, evidence suggests that glycolysis may be also an important source of ATP production for sperm motility, superior to oxidative phosphorylation (Mukai and Okuno, Reference Mukai and Okuno2004; Ford, Reference Ford2006; Nascimento et al., Reference Nascimento, Shi, Tam, Chandsawangbhuwana, Durrant, Botvinick and Berns2008).

Despite the importance of mitochondria to sperm metabolism, during oxidative phosphorylation, metabolites called reactive oxygen species (ROS) are produced and are a trigger for several reproductive physiological mechanisms (de Lamirande et al., Reference de Lamirande, Jiang, Zini, Kodama and Gagnon1997). Nevertheless, an unbalance between ROS production and mechanisms aiming to avoid their powerful oxidative potential (i.e. antioxidants), may be extremely harmful to the spermatozoa (Halliwell, Reference Halliwell1999; Nichi et al., Reference Nichi, Goovaerts, Cortada, Barnabe, De Clercq and Bols2007a).

In this context, mitochondria are highlighted as source of pro-oxidative factors that are crucial in the disruption of oxidative homeostasis (Agarwal et al., Reference Agarwal, Virk, Ong and du Plessis2014). Several studies have demonstrated correlations between impaired mitochondrial activity, oxidative stress and sperm DNA fragmentation, indicating a close relationship between these variables on sperm damage (Barros, Reference Barros2007; Nichi et al., Reference Nichi, Goovaerts, Cortada, Barnabe, De Clercq and Bols2007b; Blumer et al., Reference Blumer, Restelli, Giudice, Soler, Fraietta, Nichi, Bertolla and Cedenho2012).

Since the accident at Chernobyl, several safety improvements have been adopted and, after the Fukushima accidents, new generations of more safe designs for nuclear power stations have been developed. The main concern of nuclear energy specialists and the community, in general, is the approaches made to prevent the destruction and long-term consequences caused by an eventual nuclear disaster. If possible, the deactivation of power plants prior to predictable stressful events would probably avoid most damage. Similarly, mitochondrial therapy is applied in situations in which organelle dysfunction occurs (i.e., sperm cryopreservation) (O’Connell et al., Reference O’Connell, McClure and Lewis2002; Sariozkan et al., Reference Sariozkan, Bucak, Tuncer, Ulutas and Bilgen2009; Thomson et al., Reference Thomson, Fleming, Aitken, De Iuliis, Zieschang and Clark2009), and aimed to improve sperm viability by prevention of pro-oxidative factors release. Actually, some studies have suggested that, for certain cell types, uncouplers of oxidative phosphorylation are capable of reducing oxidative stress (Vincent et al., Reference Vincent, Olzmann, Brownlee, Sivitz and Russell2004; Mailloux & Harper, Reference Mailloux and Harper2011).

This review aimed to compile available data on the role of mitochondria in oxidative homeostasis and sperm functionality as well as suggesting some tools to assess the sperm mitochondrial function.

The mitochondrial paradox: physiological and pathological role on spermatozoa

According to the endosymbiotic theory, millions of years ago the mitochondrion was a prokaryotic unicellular organism. Formerly a free-living bacterium, the mitochondrion was capable of metabolizing oxygen in an environment rich in carbon dioxide. After penetrating a host eukaryotic cell that was incapable of metabolizing oxygen, a symbiotic relationship was established, later evolving into a more complex organism capable of producing energy more efficiently than the previously available glycolysis pathways (Margulis, Reference Margulis1970; Cummins, Reference Cummins1998). In fact, aerobic metabolism is highly dependent on mitochondrial functionality. The aerobic respiration is then, a consequence of the mitochondrial demand for oxygen which, by means of oxidative phosphorylation, is capable of producing approximately 90% of cellular energy (Saraste, Reference Saraste1999; Copeland, Reference Copeland2002).

Role of mitochondria in ATP production and sperm physiology

Studies have demonstrated the main role of mitochondria on sperm functionality, referring to this organelle as the main source of ATP for cellular homeostasis and motility (Travis et al., Reference Travis, Foster, Rosenbaum, Visconti, Gerton, Kopf and Moss1998; St John, Reference St John2002). However, its role in sperm metabolism has been a matter of debate. Mukai & Okuno (Reference Mukai and Okuno2004), when inhibiting sperm mitochondrial activity in mice concomitantly to glycolytic pathway supplementation, observed that ATP production and flagella beat remained unaltered. However, when glycolysis was inhibited and oxidative phosphorylation was stimulated, a drastically reduction in flagella beat and ATP production occurred. This finding suggested that glycolysis is more relevant than oxidative phosphorylation in the energetic metabolism of murine sperm. In a recent study conducted by our group, we observed similar results in bovine epididymal spermatozoa subjected to mitochondrial uncoupling and glycolysis stimulation (Losano et al., Reference Losano, Padín, Méndez-López, Angrimani, García and Barnabe2017a; Fig. 1). In addition, Nascimento et al. (Reference Nascimento, Shi, Tam, Chandsawangbhuwana, Durrant, Botvinick and Berns2008) observed similar results in human sperm. These authors suggested that, despite the important contribution of oxidative phosphorylation for ATP production, glycolysis is the primary source of energy in human sperm. Conversely, other studies in humans have described the opposite effect when sperm samples are incubated with inhibitors of the enzymatic electron transport complexes, with a decrease in sperm motility (Ruiz-Pesini et al., Reference Ruiz-Pesini, Lapeña, Díez-Sánchez, Pérez-Martos, Montoya, Alvarez, Díaz, Urriés, Montoro, López-Pérez and Enríquez2000; John et al., Reference John, Jokhi and Barratt2005). Furthermore, we verified that ovine sperm undergoing mitochondrial depolarization that did not alter their total motility. Spermatic kinetic patterns were affected, suggesting that mitochondria are very important in maintaining the quality of ovine spermatozoa movement (Losano et al., Reference Losano, Angrimani, Dalmazzo, Rui, Brito and Mendes2017b; Fig. 2).

Figure 1 In this study, we verified that the stimulated glycolytic pathway (glucose 5 mM) is able to maintaining total (A) and progressive (B) motilities and ATP levels (C) of bovine epididymal spermatozoa subjected to mitochondrial uncoupling [carbonyl cyanide 4-trifluoromethoxy phenylhydrazone (FCCP); 0.1, 0.3, 1 and 3 µM] (Losano et al., Reference Losano, Padín, Méndez-López, Angrimani, García and Barnabe2017a). a,b,c,dDifferent letters on the bars indicate significant differences between treatments (P<0.05).

Figure 2 We verified that mitochondrial uncoupling [carbonyl cyanide 3 chlorophenylhydrazone (CCCP); 20, 40 and 80 µM] impairs ovine sperm kinetic patterns such as progressive motility (A), straight-line velocity (VSL; B) and linearity (LIN; C), indicating an essential role of mitochondria to sperm quality movement related to progressivity (Losano et al., Reference Losano, Angrimani, Dalmazzo, Rui, Brito and Mendes2017b). a,bDifferent letters on the bars indicate significant differences between treatments (P<0.05).

Mitochondria are essential to sperm functionality due to the relationship between their functional and fertilizing capacities (Marchetti et al., Reference Marchetti, Obert, Deffosez, Formstecher and Marchetti2002, Reference Marchetti, Jouy, Leroy-Martin, Defossez, Formstecher and Marchetti2004; Gallon et al., Reference Gallon, Marchetti, Jouy and Marchetti2006; St John et al., Reference St John, Bowles and Amaral2006). Nonetheless, it is still not clear how mitochondria can contribute to the energy capacity of sperm. The organelle has distinct contributions to sperm metabolism, dependent on experimental conditions and animal species (Storey, Reference Storey2008; Amaral et al., Reference Amaral, Lourenço, Marques and Ramalho-Santos2013).

The importance of the glycolytic pathway on ATP generation and on sperm function, has been described previously (Mukai & Okuno, Reference Mukai and Okuno2004). Lardy and colleagues (Reference Lardy, Winchester and Phillips1945) first showed that mitochondrial inhibition leads to asthenospermia. However, with glucose supplementation to the samples, sperm motility was re-acquired. In addition, White & Wales (Reference White and Wales1961) observed that ovine sperm maintain their motility through two parallel mechanisms of energy generation, i.e. glycolysis and oxidative phosphorylation. Moreover, Krzyzosiak and colleagues (Reference Krzyzosiak, Molan and Vishwanath1999) also observed that bovine sperm were capable of maintaining similar motility patterns in both aerobic and anaerobic conditions, assuming that glycol sable substrates are available. Furthermore, previous studies have suggested that ATP molecules supplied by oxidative phosphorylation in the sperm midpiece are not efficiently diffused to the more distal regions of the tail, indicating that glycolysis would probably play a key role in flagella beat in this region (Nevo & Rikmenspoel, Reference Nevo and Rikmenspoel1970; Turner, Reference Turner2003).

Role of calcium on mitochondrial function

A hypothesis on the main regulatory mechanisms of oxidative phosphorylation considers ADP and inorganic phosphate as feedback substrates for ATP synthesis through several cellular kinases. Therefore, an interesting analogy can be employed with the economic model of supply and demand, with ATP as the unit price for cellular energy. Evidence to support this theory showed that mitochondria isolated in suspension increased their ATP production when ADP and inorganic phosphate was supplemented in the presence of oxygen. Despite the well known ‘economic model of equilibrium’, recent studies have shown that ATP synthesis rate is not strictly controlled by such mechanisms (Gunter et al., Reference Gunter, Yule, Gunter, Eliseev and Salter2004).

Mitochondrial calcium ([Ca2+]m) has been referred to as the central regulator of oxidative phosphorylation, acting as the primary metabolic mediator for NADH production and activity controller of the enzymatic complexes pyruvate dehydrogenase, isocitrate dehydrogenase and α-ketoglutarate dehydrogenase (McCormack et al., Reference McCormack, Halestrap and Denton1990; McCormack & Denton, Reference McCormack and Denton1993). The [Ca2+]m is also directly involved in ATP production, playing an important role in ADP phosphorylation through the enzyme ATP synthase (Territo et al., Reference Territo, French, Dunleavy, Evans and Balaban2001). Moreover, mitochondrial calcium also participates in the apoptotic mechanism of somatic cells, triggering the release of pro-apoptotic agents by the mitochondria (Szalai et al., Reference Szalai, Krishnamurthy and Hajnóczky1999).

If [Ca2+]m action on physiological processes of somatic cells is established, the precise role of this ion in sperm mitochondria is still under debate (Amaral et al., Reference Amaral, Lourenço, Marques and Ramalho-Santos2013). In a proteomic approach, studies identified sperm mitochondrial calcium uniporter (MCU) proteins that are responsible for controlling mitochondria calcium signalling, metabolism and cellular survival. However, sperm mitochondrial calcium concentration is seemingly unaltered by mitochondrial uncoupling (Machado-Oliveira et al., Reference Machado-Oliveira, Lefièvre, Ford, Herrero, Barratt, Connolly, Nash, Morales-García, Kirkman-Brown and Publicover2008; Wang et al., Reference Wang, Guo, Zhou, Shi, Yu, Yang, Wu, Wang, Liu, Chen, Tu, Zeng, Jiang, Li, Zhang, Zhou, Zheng, Yu, Zhou, Guo and Sha2013). Additionally, in bulls, mitochondrial activity on hyperactivated sperm appears to be unregulated by calcium release. In this context, further studies are vital to establish the real function of calcium in mitochondrial physiology, the reference values for [Ca2+]m, and to correlate such values with sperm function (Irvine and Aitken, Reference Irvine and Aitken1986; Ramalho-Santos et al., Reference Ramalho-Santos, Varum, Amaral, Mota, Sousa and Amaral2009; Amaral et al., Reference Amaral, Lourenço, Marques and Ramalho-Santos2013).

Reactive oxygen species and the spermatozoa

During aerobic cell metabolism, ROS are formed. This event occurs firstly because the mitochondrial environment is rich in oxygen and electrons, and almost all of these electrons participate in the reduction of oxygen directly to water, the final product of oxidative phosphorylation. Physiologically, some of these electrons escape from the oxidative phosphorylation enzymatic complex and bind to molecular oxygen, leading to first ROS, the superoxide anion, generation. From this primary product, a redox reaction cascade occurs leading to the formation of other reactive oxygen species, such as hydrogen peroxide (H2O2) and the hydroxyl radical (OH) respectively.

ROS produced are involved in many physiological triggers such as sperm hyperactivation (de Lamirande & Cagnon, Reference de Lamirande and Cagnon1993), sperm capacitation (Aitken et al., Reference Aitken, Ryan, Baker and McLaughlin2004), acrosome reaction (de Lamirande et al., Reference de Lamirande, Tsai, Harakat and Gagnon1998), and interaction between spermatozoa and the zona pellucida (Aitken et al., Reference Aitken, Paterson, Fisher, Buckingham and van Duin1995). While ROS are formed by other mechanisms, such as glycolysis, mitochondria are the main ROS source with approximately 2% of consumed oxygen being converted to superoxide anions (Koppers et al., Reference Koppers, De Iuliis, Finnie, McLaughlin and Aitken2008).

Some enzymatic and non-enzymatic antioxidants act synergistically to prevent ROS accumulation, in which each of these metabolites is inactivated by specific antioxidants. Superoxide dismutase (SOD) is considered the primary line of antioxidant defence acting through dismutation of two molecules of superoxide anion (O2 ) forming an oxygen molecule and a hydrogen peroxide molecule (H2O2; Fig. 3) (Alvarez et al., Reference Alvarez, Touchstone, Blasco and Storey1987). Hydrogen peroxide can be destroyed by two antioxidants independent systems, the enzyme catalase and the glutathione peroxidase/reductase systems (Fig. 3; Nordberg & Arnér, Reference Nordberg and Arnér2001). If these two systems fail, the H2O2 will react with an Fe2+ or Cu+ molecule (called the Fenton reaction) and will produce the hydroxyl radical (OH, Fig. 3). This ROS is considered the most reactive in biological systems, and can be destroyed by non-enzymatic antioxidants such as ascorbic acid and α-tocopherol (Fig. 3; Halliwell & Gutteridge, Reference Halliwell and Gutteridge1985).

Figure 3 Reactive oxygen species formed by the oxy-reduction process from O2 to H2O and their respective inactivation antioxidant systems. The enzyme superoxide dismutase (SOD) acting through dismutation of two molecules of superoxide anion (O2 ) forming an oxygen molecule and a hydrogen peroxide molecule. Hydrogen peroxide (H2O2) can be destroyed by two antioxidants independent systems, the enzyme catalase and glutathione peroxidase (GPx)/glutathione reductase (GR) system, with the participation of oxidized (GSSG) and reduced (GSH) glutathione. If these two systems fail, H2O2 will react with an iron (Fe2+) or (Cu+) molecule (Fenton reaction) and will form the hydroxyl radical (OH). This ROS can be destroyed by non-enzymatic antioxidants such as ascorbic acid and α-tocopherol.

Mitochondrial dysfunctions and spermatozoa

Despite the physiological function, any imbalance in ROS production and antioxidant mechanisms can lead to oxidative stress, which may be lethal for sperm cells (Fig. 4; de Lamirande et al., Reference de Lamirande, Jiang, Zini, Kodama and Gagnon1997; Agarwal et al., Reference Agarwal, Nallella, Allamaneni and Said2004). Sperm is particularly susceptible to oxidative stress due to a limited amount of cytoplasm and consequently low antioxidant activity and also a high quantity of polyunsaturated fatty acids which is easily oxidized. Thus, oxidative stress may cause damage to different sperm structures such as in plasma and acrosomal membranes, mitochondria and sperm DNA. Spermatozoa cannot restore damages caused by oxidative stress due to deficiency of cytoplasmic repair enzymes (Vernet et al., Reference Vernet, Aitken and Drevet2004; Nichi et al., Reference Nichi, Goovaerts, Cortada, Barnabe, De Clercq and Bols2007b; Agarwal et al., Reference Agarwal, Virk, Ong and du Plessis2014).

Figure 4 Physiological and pathological role of sperm mitochondria. ROS play a important role in sperm physiology acting as triggers of fertilization processes such as hyperactivation, acrosome reaction and spermatozoa–oocyte binding. However, in cases of mitochondrial dysfunctions, there is an imbalance between ROS production and antioxidant capacity, the oxidative stress. In this case, ROS cause damage to sperm structures including lipid peroxidation of the plasma membrane and DNA damage leading to loss of biological function of spermatozoa.

As mitochondria are the major source of pro-oxidative agents, it is suggested therefore that dysfunction in this organelle would have a fundamental role in the oxidative imbalance affecting sperm function (Agarwal et al., Reference Agarwal, Virk, Ong and du Plessis2014). Wang and colleagues (Reference Wang, Sharma, Gupta, George, Thomas, Falcone and Agarwal2003) identified low mitochondrial membrane potential and high ROS production in sperm from infertile patients, probably as a consequence of such mitochondrial injury, suggesting that mitochondrial function can be a marker of male fertility. In fact, other researchers have observed changes in mitochondrial function in sperm derived from infertile men (Troiano et al., Reference Troiano, Granata, Cossarizza, Kalashnikova, Bianchi, Pini, Tropea, Carani and Franceschi1998; Gallon et al., Reference Gallon, Marchetti, Jouy and Marchetti2006). However, sperm samples with high mitochondrial membrane potential have been identified in fertile patients (Kasai et al., Reference Kasai, Ogawa, Mizuno, Nagai, Uchida, Ohta, Fujie, Suzuki, Hirata and Hoshi2002; Marchetti et al., Reference Marchetti, Obert, Deffosez, Formstecher and Marchetti2002).

Studies performed in different species have shown a negative correlation between both oxidative stress and high mitochondrial activity. The occurrence of this stress and the sperm DNA integrity indicated that these variables are linked and lead a single pathogenic mechanism (Barros, Reference Barros2007; Nichi et al., Reference Nichi, Goovaerts, Cortada, Barnabe, De Clercq and Bols2007b; Blumer et al., Reference Blumer, Restelli, Giudice, Soler, Fraietta, Nichi, Bertolla and Cedenho2012). Correlation was also found between variables in spermatic oxidative stress and lower blastocyst rates, as rates of blastomeres increased with DNA damage, confirming the negative effect of seminal oxidative stress in in vitro embryonic development (Simões et al., Reference Simões, Feitosa, Siqueira, Nichi, Paula-Lopes, Marques, Peres, Barnabe, Visintin and Assumpção2013).

Mitochondrial disorders have multifactorial origins, and some mechanisms have not been totally elucidated (Amaral et al., Reference Amaral, Lourenço, Marques and Ramalho-Santos2013). These changes can be triggered even in the testis during spermatogenesis, for example if the testicular thermoregulatory mechanism is inefficient. Only 50% of the blood supply reaches the testes through the testicular artery, therefore male gonads are subjected to near hypoxic environments (Meijer & Fentener Van Vlissingen, Reference Meijer and Fentener, Van Vlissingen1993). Testis metabolism increased as consequence of pathological conditions that raised testicular temperature and were not compensated by increase in blood flow, causing a testicular hypoxic condition (Paul et al., Reference Paul, Teng and Saunders2009). Beyond these conditions and at the beginning of oxygenation, there is an increase in ROS production that leads to oxidative stress. This mechanism is known as ischaemia-reperfusion injury (Nichi et al., Reference Nichi, Bols, Züge, Barnabe, Goovaerts, Barnabe and Cortada2006; Reyes et al., Reference Reyes, Farias, Henríquez-Olavarrieta, Madrid, Parraga, Zepeda and Moreno2012). Increase in ROS production in this condition is related to mitochondrial dysfunction and the subsequent activation of enzymes that play a role in a ROS generated systems, such as xanthine oxidase (XO). These changes in the mitochondria are related to lack of O2 during ischaemia, which leads to ATP depletion and consequently mitochondrial injury. Moreover, the increase in testicular temperature promotes an influx of calcium and is also related to changes in this organelle (Dorweiler et al., Reference Dorweiler, Pruefer, Andrasi, Maksan, Schmiedt, Neufang and Vahl2007; Reyes et al., Reference Reyes, Farias, Henríquez-Olavarrieta, Madrid, Parraga, Zepeda and Moreno2012).

Sperm cryopreservation is a key process in assisted reproduction techniques (Hammerstedt et al., Reference Hammerstedt, Graham and Nolan1990; Zapzalka et al., Reference Zapzalka, Redmon and Pryor1999; Holt, Reference Holt2000). However this technique promotes a decrease in sperm quality, and also in mitochondrial damage during cryopreservation due to excessive production of pro-oxidative factors that, ultimately, cause post-thaw sperm damage and decrease in motility (O’Connell et al., Reference O’Connell, McClure and Lewis2002; Sariozkan et al., Reference Sariozkan, Bucak, Tuncer, Ulutas and Bilgen2009; Thomson et al., Reference Thomson, Fleming, Aitken, De Iuliis, Zieschang and Clark2009). Additionally, the process promotes a reduction in antioxidant capacity after sperm cryopreservation, a further factor that also predisposes these cells to oxidative stress (Bilodeau et al., Reference Bilodeau, Chatterjee, Sirard and Gagnon2000).

Consequently, several studies have used antioxidant treatment in sperm samples submitted to cryopreservation, aimed at preventing oxidative stress caused by mitochondrial injuries (Askari et al., Reference Askari, Check, Peymer and Bollendorf1994; Bilodeau et al., Reference Bilodeau, Blanchette, Gagnon and Sirard2001; Fernández-Santos et al., Reference Fernández-Santos, Martínez-Pastor, García-Macías, Esteso, Soler, Paz, Anel and Garde2007; Taylor et al., Reference Taylor, Roberts, Sanders and Burton2009). However, the use of a specific mitochondrial shield during cryopreservation also appeared as an option, aimed at improving post-thaw sperm quality (Schober et al., Reference Schober, Aurich, Nohl and Gille2007). A possible alternative would be to reduce mitochondrial activity, which can be induced by uncouplers of oxidative phosphorylation during the cryopreservation process, thus preventing ROS accumulation caused by any mitochondrial dysfunction that can occur during this process. Some uncoupler activities have been identified in the physiological processes of somatic cells, acting even in oxidative stress reduction (Vincent et al., Reference Vincent, Olzmann, Brownlee, Sivitz and Russell2004; Brand & Esteves, Reference Brand and Esteves2005).

Inhibitors and uncouplers of oxidative phosphorylation: action mechanisms and their possible applications

Inhibitors and uncouplers of oxidative phosphorylation are important in the study of mitochondrial physiology, and have been widely used in pharmacology as many chemical compounds can inhibit the specific processes of oxidative phosphorylation. Therefore, it is possible to observe their role by preventing a single process without inhibiting other mechanisms (Nelson & Cox, Reference Nelson and Cox2008).

Inhibitors can act towards complex electron carriers and also in mitochondrial channels. Rotenone (e.g. insecticide class), can block the transfer of electrons from complex I to ubiquinone, inhibiting the overall process of oxidative phosphorylation (Sherer et al., Reference Sherer, Betarbet, Testa, Seo, Richardson, Kim, Miller, Yagi, Matsuno-Yagi and Greenamyre2003). Conversely, antimycin A, an antibiotic produced by the Streptomyces fungus, blocks the transport of electrons from complex III to complex IV (Slater, Reference Slater1973). Cyanide inhibits the electron transport complex IV to oxygen. Furthermore, it is possible to directly inhibit ATP synthesis with oligomycin, which is widely used in this process. This compound acts on the enzyme ATP synthase by blocking the flow of protons through the F0 subunit of this enzyme to the mitochondrial matrix and, consequently, preventing ATP synthesis (Penefsky, Reference Penefsky1985). As well as enzyme complex inhibitors, there are also calcium channel blockers such as RU360, Na+/Ca2+pump inhibitors or CGP 37157 (García-Rivas et al., Reference García-Rivas, de, Carvajal, Correa and Zazueta2006; Thu et al., Reference Thu, Ahn and Woo2006).

In addition to these inhibitors, uncouplers of oxidative phosphorylation were widely used not only as a tool to study cell physiology, but also as a possible therapeutic application (Kasianowicz et al., Reference Kasianowicz, Benz and McLaughlin1984). ATP synthesis occurs through coupling of two reactions, electron transport and phosphorylation, as a result of a proton gradient this class of substances uncouples these two reactions, preventing or decreasing ATP synthesis. However, electron flow activity across the mitochondrial complexes is not inhibited, and even could be increased (Terada, Reference Terada1990). Most of these molecules are hydrophobic and have protonophore activity, depolarization of mitochondrial membranes allows protons to return to the mitochondrial matrix and dissipate the mitochondrial membrane potential and pH difference, inhibiting the driving proton force, essential for ATP synthesis (Chen, Reference Chen1988; Terada, Reference Terada1990). Uncoupling proteins have been identified in some cells and are related to some physiological roles such as in adaptive thermogenesis in adipose tissue.

Moreover, these proteins have been identified in researches related to obesity, diabetes, neurodegenerative disease and ageing in humans (Brand and Esteves, Reference Brand and Esteves2005). These studies emerged as previous researches found out that mitochondrial uncouplers can control mitochondria ROS production and, therefore, prevent oxidative stress, which is related to these diseases. Therefore, the use of these proteins in cell therapy for the treatment of these pathologies is suggested (Brand & Esteves, Reference Brand and Esteves2005; Lowell & Shulman, Reference Lowell and Shulman2005; Lin & Beal, Reference Lin and Beal2006; Mailloux & Harper, Reference Mailloux and Harper2011). Decrease in ROS production promoted by uncouplers is due to an increase in the respiratory rate followed by a decrease in mitochondria intermediate reduced states, capable of donating single electrons to oxygen, thereby preventing the generation of superoxide anions.

The uncoupled process has been applied in an energy study of spermatozoa (Mukai & Okuno, Reference Mukai and Okuno2004), however there is still no evidence that these compounds can control ROS production by sperm mitochondria. However, the use of these substances may bring interesting results for the prevention of oxidative stress in seminal samples in front of possible mitochondrial dysfunction. Thus, the application of this treatment can be attractive, especially for use in reproductive biotechnologies due to the highest susceptibility to oxidative stress.

Tools for assessing sperm mitochondrial function

Sperm mitochondria can be involved in both physiological as pathological processes, therefore the importance of assessing the functionality of this organelle is evident. The use of tools to evaluate sperm mitochondrial function associated with other sperm analysis can be applied for the prediction of fertilizing capacity (Troiano et al., Reference Troiano, Granata, Cossarizza, Kalashnikova, Bianchi, Pini, Tropea, Carani and Franceschi1998; Kasai et al., Reference Kasai, Ogawa, Mizuno, Nagai, Uchida, Ohta, Fujie, Suzuki, Hirata and Hoshi2002; Aitken, Reference Aitken2006). In this context, sperm mitochondria have been studied for some decades (Christen et al., Reference Christen, Schackmann and Shapiro1983; Hrudka, Reference Hrudka1987; Graham et al., Reference Graham, Kunze and Hammerstedt1990) Thus, several tools have been developed for assessment of mitochondrial function (Table 1).

Table 1 Available tools for assessing sperm mitochondrial functionality (mitochondrial activity, mitochondrial membrane potential and calcium levels assessments)

Mitochondria activity evaluation aims to infer the efficiency of electron transport between enzymatic complexes and also in the redox processes involved in oxidative phosphorylation. In classic research, Hrudka (Reference Hrudka1987) developed a cytochemical technique to evaluate mitochondrial activity. This cytochemical assay is based on the oxidation of 3′3-diaminobenzidine (DAB) by cytochromec, an enzyme involved in electron transport between the enzymatic complexes. Subsequently, some fluorescent probes such as H2-CMXros and CMXros, were developed with the same purpose and commercially sold as Mito Tracker Red® (Poot et al., Reference Poot, Zhang, Krämer, Wells, Jones, Hanzel, Lugade, Singer and Haugland1996; Wojcik et al., Reference Wojcik, Sawicki, Marianowski, Benchaib, Czyba and Guerin2000; Celeghini et al., Reference Celeghini, De Arruda, De Andrade, Nascimento and Raphael2007).

Fluorescent probes, such as JC-1 (5,5′,6,6′-tetrachloro-1,1′,3,3′-tetraethylbenzimidazolyl-carbocyanine iodide; Garner et al., Reference Garner, Thomas, Joerg, DeJarnette and Marshall1997), Mito Tracker Green FM® (Gillan et al., Reference Gillan, Evans and Maxwell2005) and Rhodamine 123® (Graham et al., Reference Graham, Kunze and Hammerstedt1990), were also developed to assess mitochondrial membrane potential. These probes diffuse freely through the plasma membrane to the cell cytosol and accumulate electrophoretically in the mitochondrial matrix, determined by proton motive force and acting in accordance with mitochondria ability to pump protons from the matrix to the intermembrane area (Chen, Reference Chen1988; Garner et al., Reference Garner, Thomas, Joerg, DeJarnette and Marshall1997; Piccoli et al., Reference Piccoli, Scrima, D’Aprile, Ripoli, Lecce, Boffoli and Capitanio2006). Membrane potential and mitochondrial activity are indicators of mitochondrial function and are related, however it is important to note that these parameters cannot be confused, as the mitochondria can maintain their redox processes by electron transport even with low membrane potential (Chen, Reference Chen1988; Terada, Reference Terada1990). Therefore, evaluation of these two parameters can be used in a complementary form.

Furthermore, it is possible to measure mitochondria calcium levels, as this mineral is considered to be the central regulator of oxidative phosphorylation (Irvine and Aitken, Reference Irvine and Aitken1986; McCormack & Denton, Reference McCormack and Denton1993). Calcium measurement in spermatozoa has been reported by the use of the fluorescent probes Quin-2 AM (Irvine & Aitken, Reference Irvine and Aitken1986), fluo-3/AM (Giojalas, Reference Giojalas1998; Harrison et al., Reference Harrison, Mairet and Miller1993) and indo-1AM (Brewis et al., Reference Brewis, Morton, Mohammad, Browes and Moore2000). However, it would ideal to measure intramitochondrial calcium, as well as create reference indices, considering that calcium has other functions in the cell such as its role in sperm capacitation (Breitbart, Reference Breitbart2002).

Although these assessments are indicative of mitochondria function, these techniques cannot be applied to quantify energy efficiency of sperm cells. Studies aimed at the evaluation of sperm energy metabolism using measurement of ATP levels are important to complement the mitochondria status assessment (Mukai and Okuno, Reference Mukai and Okuno2004). High-performance liquid chromatography (Samizo et al., Reference Samizo, Ishikawa, Nakamura and Kohama2001) or dosage by commercial kits are among the methods that can be used to measure the ATP and ADP levels (Perchec et al., Reference Perchec, Jeulin, Cosson, Andre and Billard1995). Measurement of ATP and ADP molecules was performed in several species such as mice (Mukai & Okuno, Reference Mukai and Okuno2004), birds (Rowe et al., Reference Rowe, Laskemoen, Johnsen and Lifjeld2013) and humans. However, more studies are necessary to develop indexes between production and ATP consumption, and relate these with sperm function.

Conclusion

In conclusion, there are still several questions covering the real contribution of the mitochondrial metabolism in sperm function in each species, although it is clear that this organelle can affect reproductive processes both positively and negatively (Amaral et al., Reference Amaral, Lourenço, Marques and Ramalho-Santos2013). Moreover, as mitochondria are the main ROS source and sperm are extremely susceptible to oxidative damage (Nichi et al., Reference Nichi, Goovaerts, Cortada, Barnabe, De Clercq and Bols2007b; Vernet et al., Reference Vernet, Aitken and Drevet2004), the development of studies aimed at the prevention of mitochondrial dysfunction in sperm cells is extremely important, such as the regarding improvement of mechanisms to reduce ROS release or inactivation of mechanisms for a better mitochondrial function.

Acknowledgements

The authors thank the Department of Animal Reproduction of University of São Paulo for the knowledge designed to compile the data in this review.

Financial support

The authors thank the Fundação de Amparo à Pesquisa do Estado de São Paulo (FAPESP, project: 2017/13090-9), Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq) and Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES) for the financial support for the research mentioned in this review.

Conflicts of interest

None of the authors has any conflict of interest to declare.

References

Agarwal, A, Nallella, KP, Allamaneni, SSR Said, TM (2004) Role of antioxidants in treatment of male infertility: an overview of the literature. Reprod BioMed Online 8, 616627.Google Scholar
Agarwal, A, Virk, G, Ong, C du Plessis, SS (2014) Effect of oxidative stress on male reproduction. World J Mens Health 32, 117.Google Scholar
Aitken, RJ (2006) Sperm function tests and fertility. Int J Androl 29, 6975.Google Scholar
Aitken, RJ, Paterson, M, Fisher, H, Buckingham, DW van Duin, M (1995) Redox regulation of tyrosine phosphorylation in human spermatozoa and its role in the control of human sperm function. J Cell Sci 108, 20172025.Google Scholar
Aitken, RJ, Ryan, AL, Baker, MA McLaughlin, EA (2004) Redox activity associated with the maturation and capacitation of mammalian spermatozoa. Free Rad Biol Med 36, 9941010.Google Scholar
Alvarez, JG, Touchstone, JC, Blasco, L Storey, BT (1987) Spontaneous lipid peroxidation and production of hydrogen peroxide and superoxide in human spermatozoa superoxide dismutase as major enzyme protectant against oxygen toxicity. J Androl 8, 338348.Google Scholar
Amaral, A, Lourenço, B, Marques, M Ramalho-Santos, J (2013) Mitochondria functionality and sperm quality. Reproduction 146, R163R174.Google Scholar
Askari, HA, Check, JH, Peymer, N Bollendorf, A (1994) Effect of natural antioxidants tocopherol and ascorbic acids in maintenance of sperm activity during freeze-thaw process. Syst Biol Reprod Med 33, 1115.Google Scholar
Bailey, S Macardle, PJ (2006) A flow cytometric comparison of indo-1 to fluo-3 and Fura Red excited with low power lasers for detecting Ca2+ flux. J Immunol Methods 311, 220225.Google Scholar
Barros, P (2007) Estresse oxidativo e integridade do DNA em sêmen resfriado de gato-do-mato-pequeno (Leopardus tigrinus SCHREBER 1775).Google Scholar
Bilodeau, J.-F, Chatterjee, S, Sirard, M.-A Gagnon, C (2000) Levels of antioxidant defenses are decreased in bovine spermatozoa after a cycle of freezing and thawing. Mol Reprod Dev 55, 282288.Google Scholar
Bilodeau, JF, Blanchette, S, Gagnon, C Sirard, MA (2001) Thiols prevent H2O2-mediated loss of sperm motility in cryopreserved bull semen. Theriogenology 56, 275286.Google Scholar
Blumer, CG, Restelli, AE, Giudice, P.T.D, Soler, TB, Fraietta, R, Nichi, M, Bertolla, RP Cedenho, AP (2012) Effect of varicocele on sperm function and semen oxidative stress. BJU Int 109, 259265.Google Scholar
Brand, MD Esteves, TC (2005) Physiological functions of the mitochondrial uncoupling proteins, UCP2 and, UCP3. Cell Metab 2, 8593.Google Scholar
Breitbart, H (2002) Intracellular calcium regulation in sperm capacitation and acrosomal reaction. Mol Cell Endocrinol 187, 139144.Google Scholar
Brewis, IA, Morton, IE, Mohammad, SN, Browes, CE Moore, H.D.M (2000) Measurement of intracellular calcium concentration and plasma membrane potential in human spermatozoa using flow cytometry. J Androl 21, 238249.Google Scholar
Celeghini, E.C.C, De Arruda, RP, De Andrade, A.F.C, Nascimento, J Raphael, CF (2007) Practical techniques for bovine sperm simultaneous fluorimetric assessment of plasma., acrosomal and mitochondrial membranes. Reprod Domest Anim 42, 479488.Google Scholar
Chazotte, B (2011) Labeling mitochondria with MitoTracker dyes. Cold Spring Harb Protoc 8, 990992.Google Scholar
Chen, LB (1988) Mitochondrial membrane potential in living cells. Ann Rev Cell Biol 4, 155181.Google Scholar
Christen, R, Schackmann, RW Shapiro, BM (1983) Metabolism of sea urchin sperm., Interrelationships between intracellular pH, ATPase activity and mitochondrial respiration. J Biol Chem 258, 53925399.Google Scholar
Collin, S, Sirard, MA, Dufour, M Bailey, JL (2000) Sperm calcium levels and chlortetracycline fluorescence patterns are related to the in vivo fertility of cryopreserved bovine semen. J Androl. 21, 938943.Google Scholar
Copeland, WC (2002) Mitochondrial DNA, Springer.Google Scholar
Cummins, J (1998) Mitochondrial, DNA in mammalian reproduction. Rev Reprod 3, 172182.Google Scholar
de Lamirande, E Cagnon, C (1993) Human sperm hyperactivation and capacitation as parts of an oxidative process. Free Rad Biol Med 14, 157166.Google Scholar
de Lamirande, E, Jiang, H, Zini, A, Kodama, H Gagnon, C (1997) Reactive oxygen species and sperm physiology. Rev Reprod 2, 4854.Google Scholar
de Lamirande, E.V.E, Tsai, C, Harakat, A Gagnon, C (1998) Involvement of reactive oxygen species in human sperm acrosome reaction induced by A23187, lysophosphatidylcholine, and biological fluid ultrafiltrates. J Androl 19, 585594.Google Scholar
Del Olmo, E, Bisbal, A, Maroto-Morales, A, García-Alvarez, O, Ramon, M, Jimenez-Rabadan, P, Martínez-Pastor, F, Soler, AJ, Garde, JJ Fernández-santos, MR (2013) Fertility of cryopreserved ovine semen is determined by sperm velocity. Anim Reprod Sci 138, 102109.Google Scholar
Dorweiler, B, Pruefer, D, Andrasi, T, Maksan, S, Schmiedt, W, Neufang, A Vahl, C (2007) Ischemia-reperfusion injury. Eur J Trauma Emerg Surg 33, 600612.Google Scholar
Fernández-Santos, MR, Martínez-Pastor, F, García-Macías, V, Esteso, MC, Soler, AJ, Paz, P, Anel, L Garde, JJ (2007) Sperm characteristics and, D.NA integrity of, Iberian red deer (Cervus elaphus hispanicus) epididymal spermatozoa frozen in the presence of enzymatic and nonenzymatic antioxidants. J Androl 28, 294305.Google Scholar
Ford, W.C.L (2006) Glycolysis and sperm motility: does a spoonful of sugar help the flagellum go round? Hum Reprod Update 12, 269274.Google Scholar
Forster, S, Thumser, AE, Hood, SR Plant, N (2012) Characterization of rhodamine-123 as a tracer dye for use in in vitro drug transport assays. PLoS One 7, e33253.Google Scholar
Gallon, F, Marchetti, C, Jouy, N Marchetti, P (2006) The functionality of mitochondria differentiates human spermatozoa with high and low fertilizing capability. Fertil Steril 86, 15261530.Google Scholar
García-Rivas, G, de, J, Carvajal, K, Correa, F Zazueta, C (2006) Ru360, a specific mitochondrial calcium uptake inhibitor., improves cardiac post-ischaemic functional recovery in rats in vivo . Brit J Pharmacol 149, 829837.Google Scholar
Garner, DL, Thomas, CA, Joerg, HW, DeJarnette, JM Marshall, CE (1997) Fluorometric assessments of mitochondrial function and viability in cryopreserved bovine spermatozoa. Biol Reprod 57, 14011406.Google Scholar
Gillan, L, Evans, G Maxwell, WMC (2005) Flow cytometric evaluation of sperm parameters in relation to fertility potential. Theriogenology 63, 445457.Google Scholar
Giojalas, LC (1998) Correlation between response to progesterone and other functional parameters in human spermatozoa. Fertil Steril 69, 107111.Google Scholar
Graham, JK, Kunze, E Hammerstedt, RH (1990) Analysis of sperm cell viability., acrosomal integrity, and mitochondrial function using flow cytometry. Biol Reprod 43, 5564.Google Scholar
Gunter, TE, Yule, DI, Gunter, KK, Eliseev, RA Salter, JD (2004) Calcium and mitochondria. FEBS Lett 567, 96102 2004.Google Scholar
Hallap, T, Nagy, S, Jaakma, U, Johannisson, A Rodriguez-Martinez, H (2005) Mitochondrial activity of frozen–thawed spermatozoa assessed by MitoTracker Deep Red 633. Theriogenology 63, 23112322.Google Scholar
Halliwell, B (1999) Free Radicals in Biology and Medicine, Oxford University Press.Google Scholar
Halliwell, B Gutteridge, J (1985) Free Radicals in Biology and Medicine, Pergamon.Google Scholar
Hammerstedt, RH, Graham, JK Nolan, JP (1990) cryopreservation of mammalian sperm: what we ask them to survive. J Androl 11, 7388.Google Scholar
Harrison, R.A.P, Mairet, B Miller, N.G.A (1993) Flow cytometric studies of bicarbonate-mediated, Ca2+ influx in boar sperm populations. Mol Reprod Dev 35, 197208.Google Scholar
Herzog, V Fahimi, HD (1973) A new sensitive colorimetric assay for peroxidase using 3,3′-diaminobenzidine as hydrogen donor. Anal Biochem 55, 554562.Google Scholar
Holt, WV (2000) Fundamental aspects of sperm cryobiology: the importance of species and individual differences. Theriogenology 53, 4758.Google Scholar
Hrudka, F (1987) Cytochemical and ultracytochemical demonstration of cytochrome c oxidase in spermatozoa and dynamics of its changes accompanying ageing or induced by stress. Int J Androl 10, 809828.Google Scholar
Hu, CH, Zhuang, XJ, Wei, YM, Zhang, M, Lu, SS, Lu, YG, Yang, XG Lu, KH (2017) comparison of mitochondrial function in boar and bull spermatozoa throughout cryopreservation based on JC-1 staining. Cryo Lett 38, 7579.Google Scholar
Irvine, DS Aitken, RJ (1986) Measurement of intracellular calcium in human spermatozoa. Gamete Res 15, 5771.Google Scholar
John, J.C.S, Jokhi, RP Barratt, C.L.R (2005) The effect of mitochondrial genetics on male infertility. Int J Androl 28, 6573.Google Scholar
Kasai, T, Ogawa, K, Mizuno, K, Nagai, S, Uchida, Y, Ohta, S, Fujie, M, Suzuki, K, Hirata, S Hoshi, K (2002) Relationship between sperm mitochondrial membrane potential, sperm motility, and fertility potential. Asian J Androl 4, 97104.Google Scholar
Kasianowicz, J, Benz, R McLaughlin, S (1984) The kinetic mechanism by which CCCP (carbonyl cyanidem-chlorophenylhydrazone) transports protons across membranes. J Membr Biol 82, 179190.Google Scholar
Koppers, AJ, De Iuliis, GN, Finnie, JM, McLaughlin, EA Aitken, RJ (2008) Significance of mitochondrial reactive oxygen species in the generation of oxidative stress in spermatozoa. J Clin Endocrinol Metab 93, 31993207.Google Scholar
Krzyzosiak, J, Molan, P Vishwanath, R (1999) Measurements of bovine sperm velocities under true anaerobic and aerobic conditions. Anim Reprod Sci 55, 163173.Google Scholar
Lardy, H, Winchester, B Phillips, P (1945) The respiratory metabolism of ram spermatozoa. Arch Biochem 6, 3340.Google Scholar
Lin, MT Beal, MF (2006) Mitochondrial dysfunction and oxidative stress in neurodegenerative diseases. Nature 443, 787795.Google Scholar
Losano, JD, Padín, JF, Méndez-López, I, Angrimani, DS, García, AG, Barnabe, VH et al. 2017a) The stimulated glycolytic pathway is able to maintain ATP levels and kinetic patterns of bovine epididymal sperm subjected to mitochondrial uncoupling. oxidative medicine and cellular longevity. Oxid Med Cell Longev 2017, 1682393.Google Scholar
Losano, J, Angrimani, D, Dalmazzo, A, Rui, B, Brito, M, Mendes, C et al. 2017b) Effect of mitochondrial uncoupling and glycolysis inhibition on ram sperm functionality. Reprod Domest Anim 52, 289297.Google Scholar
Lowell, BB Shulman, GI (2005) Mitochondrial dysfunction and type 2 diabetes. Science 307, 384397.Google Scholar
Machado-Oliveira, G, Lefièvre, L, Ford, C, Herrero, MB, Barratt, C, Connolly, TJ, Nash, K, Morales-García, A, Kirkman-Brown, J Publicover, S (2008) Mobilisation of Ca2+ stores and flagellar regulation in human sperm by S-nitrosylation: a role for NO synthesised in the female reproductive tract. Development 135, 36773686.Google Scholar
Mahanes, MS, Ochs, DL Eng, LA (1986) Cell calcium of ejaculated rabbit spermatozoa before and following in vitro capacitation. Biochem Biophys Res Commun 134, 664670.Google Scholar
Mailloux, RJ Harper, M.-E (2011) Uncoupling proteins and the control of mitochondrial reactive oxygen species production. Free Rad Biol Med 51, 11061115.Google Scholar
Marchetti, C, Jouy, N, Leroy-Martin, B, Defossez, A, Formstecher, P Marchetti, P (2004) Comparison of four fluorochromes for the detection of the inner mitochondrial membrane potential in human spermatozoa and their correlation with sperm motility. Hum Reprod 19, 22672276.Google Scholar
Marchetti, C, Obert, G, Deffosez, A, Formstecher, P Marchetti, P (2002) Study of mitochondrial membrane potential reactive oxygen species, DNA fragmentation and cell viability by flow cytometry in human sperm. Hum Reprod 17, 12571265.Google Scholar
Margulis, L (1970) Origin of Eukaryotic Cells. Evidence and Research Implications for a Theory of the Origin and Evolution of Microbial, Plant, and Animal Cells on the Precambrian Earth. Yale, University, Press: New Haven.Google Scholar
McCormack, JG Denton, RM (1993) Mitochondrial Ca2+ transport and the role of intramitochondrial Ca2+ in the regulation of energy metabolism. Dev Neurosci 15, 165173.Google Scholar
McCormack, JG, Halestrap, AP Denton, RM (1990) Role of calcium ions in regulation of mammalian intramitochondrial metabolism. Physiol Rev 70, 391425.Google Scholar
Meijer, J Fentener, Van Vlissingen, J (1993) Gross structure and development of reproductive organs. In Reproduction in Domesticated Animals, (ed. G.L. King), pp. 9–26. World Animal, Science-B9. Amsterdam: Elsevier Science Publishers.Google Scholar
Merritt, JE, McCarthy, SA, Davies, MP Moores, KE (1990) Use of fluo-3 to measure cytosolic Ca2+ in platelets and neutrophils, loading cells with the dye, calibration of traces., measurements in the presence of plasma, and buffering of cytosolic Ca2+ . Biochem J 269, 513519.Google Scholar
Modica-Napolitano, JS Singh, KK (2004) Mitochondrial dysfunction in cancer. Mitochondrion 4, 755762.Google Scholar
Mukai, C Okuno, M (2004) glycolysis plays a major role for adenosine triphosphate supplementation in mouse sperm flagellar movement. Biol Reprod 71, 540547.Google Scholar
Nascimento, JM, Shi, LZ, Tam, J, Chandsawangbhuwana, C, Durrant, B, Botvinick, EL Berns, MW (2008) Comparison of glycolysis and oxidative phosphorylation as energy sources for mammalian sperm motility, using the combination of fluorescence imaging., laser tweezers., and real-time automated tracking and trapping. J Cell Physiol 217, 745751.Google Scholar
Nelson, DL Cox, MM (2008) Principles of Biochemistry, 5th edn, New York, NY: W.H. Freeman and Company.Google Scholar
Nevo, AC Rikmenspoel, R (1970) Diffusion of ATP in sperm flagella. J Theoret Biol 26, 1118.Google Scholar
Nichi, M, Bols, P.E.J, Züge, RM, Barnabe, VH, Goovaerts, I.G.F, Barnabe, RC Cortada, C.N.M (2006) Seasonal variation in semen quality in Bos indicus and Bos taurus bulls raised under tropical conditions. Theriogenology 66, 822888.Google Scholar
Nichi, M, Goovaerts, IG, Cortada, CN, Barnabe, VH, De Clercq, JB Bols, PE (2007a) Roles of lipid peroxidation and cytoplasmic droplets on in vitro fertilization capacity of sperm collected from bovine epididymides stored at 4 and 34 degrees C. Theriogenology 67, 334340.Google Scholar
Nichi, M, Goovaerts, IGF, Cortada, C.N.M, Barnabe, VH, De Clercq, J.B.P Bols, P.E.J (2007b) Roles of lipid peroxidation and cytoplasmic droplets on in vitro fertilization capacity of sperm collected from bovine epididymides stored at 4 and 34°C. Theriogenology 67, 334340.Google Scholar
Nordberg, J Arnér, E.S.J (2001) Reactive oxygen species., antioxidants., and the mammalian thioredoxin system. Free Rad Biol Med 31, 12871312.Google Scholar
O’Connell, M, McClure, N Lewis, S.E.M (2002) The effects of cryopreservation on sperm morphology., motility and mitochondrial function. Hum Reprod 17, 704709.Google Scholar
Paul, C, Teng, S Saunders, P.T.K (2009) A single, mild, transient scrotal heat stress causes hypoxia and oxidative stress in mouse testes, which induces germ cell death. Biol Reprod 80, 913919.Google Scholar
Pariz, JR Hallak, J (2016) Effects of caffeine supplementation in post-thaw human semen over different incubation periods. Andrologia 48, 961966.Google Scholar
Penefsky, HS (1985) Mechanism of inhibition of mitochondrial adenosine triphosphatase by dicyclohexylcarbodiimide and oligomycin: relationship to ATP synthesis. Proc Natl Acad Sci USA 82, 15891593.Google Scholar
Perchec, G, Jeulin, C, Cosson, J, Andre, F Billard, R (1995) Relationship between sperm ATP content and motility of carp spermatozoa. J Cell Sci 108, 747753.Google Scholar
Piccoli, C, Scrima, R, D’Aprile, A, Ripoli, M, Lecce, L, Boffoli, D Capitanio, N (2006) Mitochondrial dysfunction in hepatitis C virus infection. Biochim Biophys Acta 1757, 14291437.Google Scholar
Poot, M, Zhang, YZ, Krämer, JA, Wells, KS, Jones, LJ, Hanzel, DK, Lugade, AG, Singer, VL Haugland, RP (1996) Analysis of mitochondrial morphology and function with novel fixable fluorescent stains. J Histochem Cytochem 44, 13631372.Google Scholar
Ramalho-Santos, J, Varum, S, Amaral, S, Mota, PC, Sousa, AP Amaral, A (2009) Mitochondrial functionality in reproduction: from gonads and gametes to embryos and embryonic stem cells. Hum Reprod Update 15, 553572.Google Scholar
Rasola, A Geuna, M (2001) A flow cytometry assay simultaneously detects independent apoptotic parameters. Cytometry 45, 151157.Google Scholar
Reyes, JG, Farias, JG, Henríquez-Olavarrieta, S, Madrid, E, Parraga, M, Zepeda, AB Moreno, RD (2012) The hypoxic testicle: physiology and pathophysiology. Oxid Med Cell Longev 2012, 929285.Google Scholar
Rowe, M, Laskemoen, T, Johnsen, A Lifjeld, JT (2013) Evolution of sperm structure and energetics in passerine birds. Proc Biol Sci 280 (1753):20122616.Google Scholar
Ruiz-Pesini, E, Lapeña, A.-C, Díez-Sánchez, C, Pérez-Martos, A, Montoya, J, Alvarez, E, Díaz, M, Urriés, A, Montoro, L, López-Pérez, MJ Enríquez, JA (2000) Human mtDNA haplogroups associated with high or reduced spermatozoa motility. Am J Hum Genet 67, 682696.Google Scholar
Samizo, K, Ishikawa, R, Nakamura, A Kohama, K (2001) A highly sensitive method for measurement of myosin atpase activity by reversed-phase high-performance liquid chromatography. Anal Biochem 293, 212215.Google Scholar
Saraste, M (1999) Oxidative phosphorylation at the fin de siècle. Science 283, 14881493.Google Scholar
Sariozkan, S, Bucak, MN, Tuncer, PB, Ulutas, PA Bilgen, A (2009) The influence of cysteine and taurine on microscopic-oxidative stress parameters and fertilizing ability of bull semen following cryopreservation. Cryobiology 58, 134138.Google Scholar
Schober, D, Aurich, C, Nohl, H Gille, L (2007) Influence of cryopreservation on mitochondrial functions in equine spermatozoa. Theriogenology 68, 745754.Google Scholar
Sherer, TB, Betarbet, R, Testa, CM, Seo, BB, Richardson, JR, Kim, JH, Miller, GW, Yagi, T, Matsuno-Yagi, A Greenamyre, JT (2003) Mechanism of toxicity in rotenone models of Parkinson’s disease. J Neurosci 23, 1075610764.Google Scholar
Simões, R, Feitosa, WB, Siqueira, A.F.P, Nichi, M, Paula-Lopes, FF, Marques, MG, Peres, MA, Barnabe, VH, Visintin, JA Assumpção, M.E.O (2013) Influence of bovine sperm, DNA fragmentation and oxidative stress on early embryo in vitro development outcome. Reproduction 146, 433441.Google Scholar
Slater, EC (1973) The mechanism of action of the respiratory inhibitor, antimycin. Biochim Biophys Acta 301, 129154.Google Scholar
Smiley, ST, Reers, M, Mottola-Hartshorn, C, Lin, M, Chen, A, Smith, TW, Steele, GD Chen, LB (1991) Intracellular heterogeneity in mitochondrial membrane potentials revealed by a, J.-aggregate-forming lipophilic cation, J.C-1. Proc Natl Acad Sci USA 88, 36713675.Google Scholar
St John, J, Bowles, EJ Amaral, A (2006) Sperm mitochondria and fertilisation. Soc Reprod Fertil Suppl 65, 399416.Google Scholar
St John, JC (2002) The transmission of mitochondrial, DNA following assisted reproductive techniques. Theriogenology 57, 109123.Google Scholar
Storey, BT (2008) Mammalian sperm metabolism: oxygen and sugar, friend and foe. Int J Dev Biol 52, 427.Google Scholar
Szalai, G, Krishnamurthy, R Hajnóczky, G (1999) Apoptosis driven by IP3-linked mitochondrial calcium signals. EMBO J 18, 63496361.Google Scholar
Taylor, K, Roberts, P, Sanders, K Burton, P (2009) Effect of antioxidant supplementation of cryopreservation medium on post-thaw integrity of human spermatozoa. Reprod BioMed Online 18, 184189.Google Scholar
Terada, H (1990) Uncouplers of oxidative phosphorylation. Environ Health Perspect 87, 213.Google Scholar
Territo, PR, French, SA, Dunleavy, MC, Evans, FJ Balaban, RS (2001) Calcium activation of heart mitochondrial oxidative phosphorylation: rapid kinetics of mv, O2, NADH, and light scattering. J Biol Chem 276, 25862599.Google Scholar
Thomson, LK, Fleming, SD, Aitken, RJ, De Iuliis, GN, Zieschang, JA Clark, AM (2009) Cryopreservation-induced human sperm, DNA damage is predominantly mediated by oxidative stress rather than apoptosis. Hum Reprod 24, 20612070.Google Scholar
Thu, LT, Ahn, JR Woo, S.-H (2006) Inhibition of L-type Ca2+ channel by mitochondrial, Na+–Ca2+ exchange inhibitor CGP-37157 in rat atrial myocytes. Eur J Pharmacol 552, 1519.Google Scholar
Travis, AJ, Foster, JA, Rosenbaum, NA, Visconti, PE, Gerton, GL, Kopf, GS Moss, SB (1998) Targeting of a germ cell-specific type 1 hexokinase lacking a porin-binding domain to the mitochondria as well as to the head and fibrous sheath of murine spermatozoa. Mol Biol Cell 9, 263276.Google Scholar
Troiano, L, Granata, ARM, Cossarizza, A, Kalashnikova, G, Bianchi, R, Pini, G, Tropea, F, Carani, C Franceschi, C (1998) mitochondrial membrane potential and DNA stainability in human sperm cells: a flow cytometry analysis with implications for male infertility. Exp Cell Res 241, 384393.Google Scholar
Turner, RM (2003) Tales from the tail: what do we really know about sperm motility? J Androl 24, 790803.Google Scholar
Vernet, P, Aitken, RJ Drevet, JR (2004) Antioxidant strategies in the epididymis. Mol Cell Endocrinol 216, 3139.Google Scholar
Vincent, AM, Olzmann, JA, Brownlee, M, Sivitz, WI Russell, JW (2004) Uncoupling proteins prevent glucose-induced neuronal oxidative stress and programmed cell death. Diabetes 53, 726734.Google Scholar
Wang, G, Guo, Y, Zhou, T, Shi, X, Yu, J, Yang, Y, Wu, Y, Wang, J, Liu, M, Chen, X, Tu, W, Zeng, Y, Jiang, M, Li, S, Zhang, P, Zhou, Q, Zheng, B, Yu, C, Zhou, Z, Guo, X Sha, J (2013) In-depth proteomic analysis of the human sperm reveals complex protein compositions. J Proteomics 79, 114122.Google Scholar
Wang, X, Sharma, RK, Gupta, A, George, V, Thomas, AJ Jr, Falcone, T Agarwal, A (2003) Alterations in mitochondria membrane potential and oxidative stress in infertile men: a prospective observational study. Fertil Steril 80 (Suppl 2), :844850.Google Scholar
White, IG Wales, RG (1961) Comparison of epididymal and ejaculated semen of the ram. J Reprod Fertil 2, 225237.Google Scholar
Wojcik, C, Sawicki, W, Marianowski, P, Benchaib, M, Czyba, JC Guerin, JF (2000) Cyclodextrin enhances spermicidal effects of magainin-2-amide. Contraception 62, 99103.Google Scholar
Zhang, CL Wu, BJ (1996) Development of calcium fluorescent probes and their application in life sciences. Progr Physiol 27, 3742.Google Scholar
Zapzalka, DM, Redmon, JB Pryor, JL (1999) A survey of oncologists regarding sperm cryopreservation and assisted reproductive techniques for male cancer patients. Cancer 86, 18121817.Google Scholar
Figure 0

Figure 1 In this study, we verified that the stimulated glycolytic pathway (glucose 5 mM) is able to maintaining total (A) and progressive (B) motilities and ATP levels (C) of bovine epididymal spermatozoa subjected to mitochondrial uncoupling [carbonyl cyanide 4-trifluoromethoxy phenylhydrazone (FCCP); 0.1, 0.3, 1 and 3 µM] (Losano et al., 2017a). a,b,c,dDifferent letters on the bars indicate significant differences between treatments (P<0.05).

Figure 1

Figure 2 We verified that mitochondrial uncoupling [carbonyl cyanide 3 chlorophenylhydrazone (CCCP); 20, 40 and 80 µM] impairs ovine sperm kinetic patterns such as progressive motility (A), straight-line velocity (VSL; B) and linearity (LIN; C), indicating an essential role of mitochondria to sperm quality movement related to progressivity (Losano et al., 2017b). a,bDifferent letters on the bars indicate significant differences between treatments (P<0.05).

Figure 2

Figure 3 Reactive oxygen species formed by the oxy-reduction process from O2 to H2O and their respective inactivation antioxidant systems. The enzyme superoxide dismutase (SOD) acting through dismutation of two molecules of superoxide anion (O2) forming an oxygen molecule and a hydrogen peroxide molecule. Hydrogen peroxide (H2O2) can be destroyed by two antioxidants independent systems, the enzyme catalase and glutathione peroxidase (GPx)/glutathione reductase (GR) system, with the participation of oxidized (GSSG) and reduced (GSH) glutathione. If these two systems fail, H2O2 will react with an iron (Fe2+) or (Cu+) molecule (Fenton reaction) and will form the hydroxyl radical (OH). This ROS can be destroyed by non-enzymatic antioxidants such as ascorbic acid and α-tocopherol.

Figure 3

Figure 4 Physiological and pathological role of sperm mitochondria. ROS play a important role in sperm physiology acting as triggers of fertilization processes such as hyperactivation, acrosome reaction and spermatozoa–oocyte binding. However, in cases of mitochondrial dysfunctions, there is an imbalance between ROS production and antioxidant capacity, the oxidative stress. In this case, ROS cause damage to sperm structures including lipid peroxidation of the plasma membrane and DNA damage leading to loss of biological function of spermatozoa.

Figure 4

Table 1 Available tools for assessing sperm mitochondrial functionality (mitochondrial activity, mitochondrial membrane potential and calcium levels assessments)