Next Article in Journal
Oxidative Stress and Its Role in Cd-Induced Epigenetic Modifications: Use of Antioxidants as a Possible Preventive Strategy
Next Article in Special Issue
Roles of Reactive Oxygen Species in Vascular Complications of Diabetes: Therapeutic Properties of Medicinal Plants and Food
Previous Article in Journal
Solid and Liquid Oxygen under Ultrahigh Magnetic Fields
Previous Article in Special Issue
Biological and Pharmacological Properties of Carbon Monoxide: A General Overview
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Biological Relevance of Free Radicals in the Process of Physiological Capacitation and Cryocapacitation

Institute of Applied Biology, Faculty of Biotechnology and Food Sciences, Slovak University of Agriculture, Trieda Andreja Hlinku 2, 94976 Nitra, Slovakia
*
Author to whom correspondence should be addressed.
Submission received: 4 May 2022 / Revised: 23 May 2022 / Accepted: 24 May 2022 / Published: 27 May 2022
(This article belongs to the Special Issue Review Papers in Oxygen)

Abstract

:
Before fertilization, spermatozoa must undergo a process called capacitation in order to fulfill their fertilization potential. This includes a series of structural, biochemical, and functional changes before a subsequent acrosome reaction and fusion with the oocyte. However, low temperatures during cryopreservation may induce a premature activation of capacitation-like changes, also known as cryocapacitation, immediately after thawing, which may lead to a decreased viability, motility, and fertilization ability of cryopreserved spermatozoa. Furthermore, cryopreservation is responsible for the overgeneration of reactive oxygen species (ROS) such as superoxide, hydrogen peroxide, and hydroxyl radicals, which may result in the development of oxidative stress, cell membrane damage, and lipid peroxidation. Despite that, both capacitation and cryocapacitation are considered to be oxidative events; however, potential beneficial or detrimental effects of ROS depend on a wide array of circumstances. This review summarizes the available information on the role of free radicals in the process of capacitation and cryocapacitation of spermatozoa.

1. Introduction

Free radicals (FRs) may be described as unstable molecules with an unpaired electron in its atomic orbital. As a natural by-product of oxygen metabolism, they present with both toxic as well as beneficial properties, which are involved in the regulation of a multitude of intracellular pathways responsible for sperm capacitation, hyperactivation, acrosome reaction (AR), and fusion with the ovum [1]. On the contrary, if the production of FRs accelerates, the process may lead to the development of oxidative stress (OS). OS is a condition primarily caused by an imbalance between the generation of FRs and the antioxidant ability of cells to detoxify these highly reactive molecules [2,3]. This disproportion negatively affects the structural integrity and functional activity of a wide range of cellular components including carbohydrates, nucleic acids, lipids, and proteins [4].
Based on their backbone, FR can be divided into reactive oxygen species (ROS), reactive nitrogen species (RNS), and other nonradical reactive species [5]. The most common FRs belong to the family of ROS and include the superoxide anion (O2•−), hydroxyl radical (OH), and hydrogen peroxide (H2O2). The group of RNS is presented by nitric oxide (NO), nitrogen dioxide (NO2), and peroxynitrite (NO3) [6,7]. In general, oxygen is necessary for aerobic metabolism of spermatozoa, but it also plays the main role in the generation of ROS, which could eventually cause a variety of anomalies such as head and acrosome defects, mid piece abnormalities, cytoplasmatic droplets, or tail dysfunction [8]. Nowadays, seminal OS is one of the main reasons for the development of male sub- or infertility [9,10].
Spermatozoa contain a high concentration of polyunsaturated fatty acids (PUFAs), which make them vulnerable to lipid peroxidation (LPO) as lipid structures of the cell membranes are one of the primary targets for ROS. Naturally, male reproductive cells have their own defense against ROS represented by the antioxidant enzymes present in the cytoplasm; however, most of them are lost during spermatogenesis [11].
As mentioned earlier, a specific amount of ROS is important for the regulation of biological events such as sperm capacitation. The capacitation process includes a cascade of biochemical and structural transformations, which are necessary for spermatozoa to reach their full fertilization potential [12]. Under physiological conditions, capacitation begins immediately following ejaculation, or it may be induced under laboratory conditions by using a specific capacitation medium enriched with heparin or L-arginine [13,14]. After completing capacitation, spermatozoa are able to undergo AR and fertilize the oocyte. The process includes changes in the membrane fluidity and composition, a higher concentration of intracellular calcium, alkalization of cytoplasm, triggering of ion channels, and generation of a small amount of ROS, which work as molecular messengers for the initiation of protein tyrosine phosphorylation modulated by the cAMP-dependent pathway [14,15] (Figure 1).
Over the years, cryopreservation has become a routine technique used for the long-term preservation of fertile male gametes collected not only from domestic animals [16,17] but also from humans [18,19]. Despite the benefits of cryopreservation, this technology may cause irreversible changes to cellular compounds and structures. What is more important is that cryodamage is associated with an increased generation of ROS such as O2•−, H2O2, and OH [16,17,19]. It has been reported that the cryopreservation process including dilution, cooling, and freezing/thawing protocols is responsible for the occurrence of capacitation-like changes in spermatozoa, also known as cryocapacitation [17] (Figure 1). The cryopreservation itself significantly reduces the quality of semen after thawing and is characterized by the destabilization of the cell plasma membrane accompanied by the loss of sperm movement and viability. Disruption of the membrane lipid architecture decreases the capability of spermatozoa to reach and penetrate the oocyte, which is a commonly observed complication of premature capacitation and AR [20].
As such, the objective of this review was: (1) to summarize and evaluate the involvement of free radicals in the process of capacitation and cryocapacitation, (2) to define physiological and pathological roles of ROS in the sperm function, and (3) to describe the negative effects of cryopreservation and suggest possible prevention strategies against cryodamage.

2. The Origin of ROS in Semen

Several sources of ROS in semen have been identified such as seminal leukocytes, abnormal spermatozoa, or mitochondria. The presence of leukocytes in ejaculate is often associated with infection or inflammation [7,11]. According to Li et al., high concentrations of ROS and inflammatory cytokines strongly correlate with the quantity of leukocytes in semen collected from sub-fertile men [21]. Activated peroxide-positive leukocytes including polymorphonuclear (PMN) leukocytes and macrophages are able to produce 100 times higher levels of ROS when compared to inactive leukocytes [22]. Activated leukocytes are also responsible for NADPH oxidase catalysis of free radicals by hexose for inflammatory protection. Leukocytospermia is a condition characterized by a high concentration of white blood cells in semen. According to the World Health Organization (WHO), it is defined as the incidence of 1 × 106 leukocytes/mL in semen, which may depend on the presence and/or range of a urogenital infection in men [23,24].
Based on previous studies, a positive correlation has been established between high levels of seminal ROS and poor sperm morphology. Out of several different sperm abnormalities, cytoplasmic droplets or residual cytoplasm were identified as the most involved in ROS generation. Usually, these cytoplasmatic residues appear attached to the elongated spermatid released from the germinal epithelium and migrate away from the sperm neck toward the end of the midpiece during epididymal transport [25]. The number of immature or morphologically retarded spermatozoa is expressed by the sperm deformity index (SDI), which is a quantitative representation of the sperm morphological quality in human andrology [26,27,28]. Cytoplasmic droplets mediate the production of ROS by the cytosolic enzyme called glucose-6-phosphate dehydrogenase (G6PD) via two pathways. The first one is catalyzed by the nicotinamide adenine dinucleotide phosphate (NADPH), which is localized in the plasmatic membrane, while the second one is driven by the NADPH-dependent oxido-reductase present in the middle piece of spermatozoa. Indeed, plasmatic membranes of spermatozoa and mitochondria have been identified as two main producers of ROS in human semen [29,30].
Despite most of the cytoplasm being reduced during spermatogenesis, mitochondria remain in the midpiece of the sperm flagellum. The main role of these organelles lies in the production of energy in the form of adenosine triphosphate (ATP) through oxidative phosphorylation, which is elementary for the sperm movement [31]. The key to the generation of ROS is associated with the activation of the mitochondrial electron transport chain via five protein complexes (complex I–V), which can transport electrons from NADPH to oxygen (O2). Positively charged protons and protein complex V (ATP synthase, 19 protein subunits) create the mitochondrial membrane potential, allowing the production of ATP [31,32]. However, a potential leakage of electrons from the mitochondrial electron transport chain caused by stress conditions leads to a partial reduction of O2 into O2•−. Following that, superoxide dismutases (SOD enzymes) will dismutate O2•− to H2O2 in the intermembrane space of mitochondria. Moreover, O2•− has the ability to react with nitric oxide (NO) and generate peroxynitrite (ONOO2•−) [33,34].
On the other hand, high levels of PUFAs present in the membranes stimulate the production of lipid aldehydes such as acrolein, malondialdehyde (MDA), or 4-hydroxynonenal (4HNE), as well as lipid metabolites including lipid peroxyl or alkoxyl radicals [35]. These substances are strong electrophiles that may disrupt the functionality of the mitochondrial electron transport chain by their covalent binding into mitochondrial proteins (mainly to succinic acid dehydrogenase) and thus dysregulate the electron flow. The result of this complex chemical event is a self-repetitive cycle where the amount of PUFAs positively corelates with the generation of mitochondrial ROS and lipid aldehydes. Moreover, 4HNE may bind to mitochondrial proteins, and subsequently induce electron leakage and ROS overgeneration [35,36].

3. Relationship between Capacitation and ROS

The process of mammalian sperm capacitation was first mentioned in 1951 by Chang [37] and Austin [38]. This molecular event combines a series of morphological and biochemical changes, which allow spermatozoa to achieve their full fertilization potential. According to Visconti et al., capacitation may be divided into fast and slow events. The fast or early activation is characterized by a vigorous and asymmetric movement of the flagellum right after ejaculation. Slow and late events include changes in the sperm motility patterns and acquisition of the ability to fertilize the ovum [39]. The initiation of capacitation is accompanied with an efflux of cholesterol from the plasmatic membrane by albumin, increasing the membrane permeability. Subsequently, membrane translocation of bicarbonate (HCO3−) increases the intracellular pH and influx of calcium (Ca2+). After that, activated Ca2+ channels and secondary messenger systems including adenylyl cyclase (AC) lead into the escalation of cAMP production and induction of protein phosphorylation, specifically tyrosine phosphorylation [39,40]. This stimulates the complexes of protein kinases and the generation of a specific amount of ROS necessary for this oxidative redox-regulated process [41,42].
Under physiological conditions, a low concentration of ROS produced by spermatozoa is required for the activation of signal transduction processes associated with capacitation. More than 20 years ago, de Lamirande and Gagnon [43] suggested that capacitation is a ROS-dependent process. According to the authors, the addition of superoxide dismutase (SOD) had a capacitation-preventive effect on human spermatozoa, while the supplementation by exogenous O2•− induced capacitation. SOD with other enzymes (catalase—CAT, glutathione peroxidase/reductase—GPX/GSR) is a part of a complex scavenger system provided by the mammalian seminal plasma for the protection against oxidative stress or premature capacitation [44]. ROS, and particularly their dismutated products O2•−, H2O2, and hydroxyl radicals (OH), are capable of trespassing the cell membranes and disrupting the structure of various intracellular molecules such as proteins, lipids, and nucleic acids. Nevertheless, a specific amount of ROS plays a pivotal role in the regulation of cholesterol efflux, cAMP production, and tyrosine phosphorylation [41,42,43]. According to de Lamirande [45], the concentration of O2•− reached a peak after 15–25 min of in vitro capacitation and constantly decreased for the next 1–2 h in human spermatozoa. This comes along with the theory that ROS are important mainly at the beginning of the capacitation process [40].
Spermatozoa themselves may produce ROS from sources such as NADPH oxidases (NOX) such as NADPH oxidase localized in the plasmatic membrane or NADPH-dependent oxidoreductase in the mitochondria, which plays a major role in high ATP generation. Increasing levels of O2•−, H2O2, NO, and peroxynitrite (ONOO) produced constantly during capacitation stimulate the activity of AC, cAMP, and protein kinase A (PKA), which is elementary for late tyrosine phosphorylation [46,47,48,49,50]. The activation of PKA phosphorylates and protein phosphatase inhibition stimulate tyrosine kinase, which accelerates actin polymerization that leads to hyperactivated sperm motility in humans [51]. Ghanbari et al. [52] demonstrated a key role of selected NADPH oxidase NOX5 in mammalian spermatozoa, which is involved in O2•− production. This isoform is different from other members of the NOX family. The activation of NOX5 is Ca2+-dependent because of special calcium-binding sites in the N-terminal region, which means that it does not need any NADPH oxidase subunit [53]. Numerous authors proposed that NOX5-driven O2•− production initiates the activation of AC, the cAMP/PKA pathway, and tyrosine phosphorylation associated with capacitation, as confirmed in human [43], ram [53], equine [54], and rat spermatozoa [55].
It was suggested that the sperm oxidase is responsible for the production of O2•−, right after incubation with selected inductors of capacitation. This hypothesis was supported by the fact that the generation of H2O2 under in vitro conditions is characteristic of the production of diphenyliodonium (DPI), which acts as an inhibitor of NADPH oxidase and a capacitation blocker. In general, O2•− is known for a short life-span and spontaneous dismutation into H2O2 by SOD. An increased consumption of H2O2 is associated with acrosome reaction in bovine spermatozoa, where H2O2 is responsible for the activation of protein kinases and phospholipase A2 [15,55]. Nevertheless, the exact role of the sperm oxidase is still elusive; however, the presence of NADPH is crucial for the generation of O2•− as well as NO. Nicotinamide adenine dinucleotide (NADH) is generated during the oxidation of lactate into pyruvate by a specific isoform of lactate dehydrogenase C4 (LDH-C4) localized in the cytosol, mitochondria, and the plasmatic membrane in spermatozoa of different species, such as humans [41], bulls [56], or mice [57]. Subsequently, pyruvate is translocated into the mitochondria and immediately converted to acetyl coenzyme A (acetyl-CoA) as a part of the Krebs cycle and ATP production. The generated NADH is then used to produce extracellular O2•− through cytosolic oxidases [41,42]. Previous studies reported a participation of LDH-C4 in the metabolic activity and capacitation process in bovine or mouse spermatozoa. On the other hand, low LDH-C4 activity was characterized by a limited or totally reduced sperm movement and concentration [57,58].
As mentioned before, another substantial ROS involved in the capacitation is NO. Lefiévre et al. [59] reported that NO is able to regulate capacitation by protein S-nitrosylation and activation of the cAMP/PKA pathway in humans, which leads to an increase in protein phosphorylation, especially of serine, threonine, and tyrosine [60]. Another possible way by which NO could regulate protein phosphorylation is via soluble isoforms of guanylate monophosphate (sGC). These isoforms increase the concentration of cyclic guanosine monophosphate (cGMP) in human spermatozoa; then, cGMP stimulates the activity of cGMP-dependent protein kinase (PKG), which supports the protein phosphorylation of serine/threonine. The intracellular increase in cGMP concentration possibly ceases the degradation of cAMP by the activity of cyclic nucleotide phosphodiesterase. In the end, higher amounts of cAMP may stimulate PKA and protein phosphorylation of tyrosine, which promotes the course of capacitation [15,61,62].
It was confirmed that several proteins associated with capacitation such as protein kinase C (PKC), Ras protein, and AC undergo a process called redox signaling. The highly reactive thiol groups of cysteine sulfhydryl (SH) residues interact with ROS or RNS and create oxidative post-translational modifications. A continuous increase in SH groups is associated with the reorganization of proteins in the sperm membranes at the beginning of capacitation. The most common reversible redox signaling modifications include disulfide bridges, S-nitrosylation, S-glutathionylation, and S-sulfenation, which may be reduced back to thiols or stable oxidized products. However, a covalent sulphinylation of cysteine is irreversible and leads to the formation of tyrosine, tryptophan, lysine, and histidine. These substances may cause cell damage by the loss of a proper protein function [63,64].

4. Cryodamage and ROS Overproduction

Nowadays, cryopreservation is a powerful technique used for the stabilization of spermatozoa at cryogenic temperatures for their long-term storage. It plays a primary role in the improvement of genetic resources in farm breeding programs, preservation of rare and endangered animals, as well as management of infertility in numerous species including humans [17,18,65]. Despite numerous benefits including the preservation of sperm motility, metabolic activity, and fertility, the cryopreservation process may present with negative effects on the sperm physiology and cause irreversible changes that may possibly lead to cell death and a reduced quality of frozen-thawed spermatozoa. Cryopreservation is essentially a three-step process consisting of cooling, the addition of cryoprotectant, and freezing/thawing. During freezing, spermatozoa are exposed to a wide spectrum of stress conditions such as cold shock, osmotic imbalance, and oxidative stress. Male gametes are protected by several membranes, which act as a natural barrier against external factors. However, for their proper function, these membranes must remain intact to secure the post-thaw viability of spermatozoa [66,67].
The generation of ROS during freezing/thawing procedures comes from mutual interactions between NADPH oxidase in the plasmatic membrane and the mitochondrial electron transport chain. Following cryopreservation, the antioxidant activity of the seminal plasma is dramatically reduced because of dilution and a decrease in enzymatic and nonenzymatic antioxidants such ascorbic acid, urate, vitamin E, and pyruvate [68].
In mammals, the plasma membrane of spermatozoa contains a high concentration of PUFAs, particularly docosahexaenoic acid (DHA), which are highly susceptible to oxidative damage. This process is also known as lipid peroxidation (LPO). Peroxidation of PUFAs is characterized as an autocatalytic self-repeating reaction, which may evolve into the loss of membrane integrity, fluidity, and functionality. The process is associated with electron removal from lipids and the production of reactive intermediates [69,70]. Bouwers and Gadella [71] observed that LPO occurs primarily in the midpiece and tail regions of the sperm flagellum, indicating endogenous ROS production by mitochondria. Peroxidation events significantly affect the mitochondria and cause an immediate decrease in the ATP production and sperm movement as reported in stallions [66] and bulls [71]. Dietrich et al. [72] suggested that specific forms of apolipoprotein A-I provide semen freezability by supporting the membranous structures of carp spermatozoa. Due to freezing, membrane proteins such as N-ethylmaleimide-sensitive fusion attachment protein α and anexin A4 were reduced in this study, which disintegrated the membrane integrity of spermatozoa in carps [72,73] and rats [74]. Other contraindications associated with membrane damage include the leakage of intracellular proteins needed for metabolism, cellular signaling, and the organization of the sperm cytoskeleton, which increase the concentration of intracellular enzymes in the extracellular space [73,74,75].
Due to oxidative stress, PUFAs are attacked by ROS and produce a wide spectrum of lipid metabolites, as mentioned above. Destabilization of the sperm plasma membrane is caused by the ability of lipid peroxyl radicals to take the hydrogen atoms from PUFAs. Subsequently, carbon-centered lipid radicals will react with oxygen and increase the generation of peroxyl radicals, leading to a self-propagating cycle of hydrogen removal from PUFAs and the promotion of LPO chain reactions [76,77]. These reactions will continue until one of the substrates is consumed or the case of a reaction between two radicals. This process is also responsible for the production of lysophospholipids that destabilize and disturb the microarchitecture of the plasma membrane. The changed structure and function of elemental membrane proteins has a negative effect on ATP-dependent ion pumps and voltage-regulated ion channels, which manage the motility of spermatozoa. Intensive peroxidative damage on the lipid membrane structures makes spermatozoa unable to participate in the membrane fusion processes during fertilization, including capacitation and acrosome reaction [70,76].

5. Capacitation-Like Changes of Spermatozoa during Cryopreservation

Over the years, various researchers have investigated specific changes in frozen–thawed spermatozoa that shared similar characteristics with physiological capacitation. Cryo-induced modifications or cryocapacitation destines spermatozoa to be more vulnerable to their environment because of the reorganization and redistribution of membrane phospholipids. The kinetics of cryocapacitation is not fully understood, because of numerous similarities with physiological capacitation. Following cryocapacitation, the leakage of membrane proteins from thawed spermatozoa is significantly higher in comparison to naturally capacitated spermatozoa [73]. Alterations to the sperm membrane proteins during the cryopreservation process include their segregation, inactivation of membrane-bound enzymes, and a decreased protein diffusion, all of which may eventually lead to the loss of structural proteins and receptors located on the sperm surface [78,79]. As such, a premature capacitation may reduce the fertilization potential and viability of cryopreserved spermatozoa.
The asymmetric phospholipid architecture from the outside of the plasmatic membrane consists of phosphatidylcholine and sphingomyelin, while the inner side is made from phosphatidylserine and phosphatidylethanolamine. During capacitation or cryopreservation, the asymmetric structure of phospholipids is disturbed by scrambling phosphatidylserine and phosphatidylethanolamine, which could lead to membrane lipid disorders [19], as observed in buffalo [79], equine [80], and boar spermatozoa [81]. According to Cross [82], this disarrangement and a subsequent efflux of cholesterol following in vitro capacitation increases the permeability of the phospholipid bilayer and disrupts phospholipid packing.
Cormier et al. [13] studied the differences of tyrosine phosphoprotein profiles in heparin (HEP)-capacitated and -cryocapacitated bovine spermatozoa. The obtained results revealed the presence of two phosphotyrosine proteins 56-PP and 114-PP in the HEP-capacitated group; however, only 56-PP was found in the case of cryocapacitated spermatozoa. Cryopreserved spermatozoa presented with a characteristic tyrosine-phosphorylated state, which resembled early/fast modifications of the membrane during physiological capacitation. This hypothesis was postulated earlier by Bailey and Bérubé [83], according to who cryopreserved bovine spermatozoa exhibited the evidence of phosphotyrosine protein activity immediately after thawing. In contrary, fresh bull spermatozoa needed a minimal 4 h incubation with HEP for the detection of phosphotyrosine proteins [84].
Hyperactivation is a necessary part of capacitation, when the sperm proteins are tyrosine-phosphorylated and regulated with cAMP by activation of PKA. Several tyrosine-phosphorylated proteins have been localized in the sperm flagellum, which suggests their involvement in sperm hyperactivation. After cryopreservation, male gametes collected from buffaloes [46,79] and bulls [13,84] presented with a hyperactivated motility and weakened Ca2+ mechanisms, which may lead to a significant accumulation of the Ca2+ ion. If the intracellular Ca2+ concentrations reach critical levels, the acrosome reaction is activated too early, even before any contact with the receptors located in the zona pellucida of oocyte has occurred [13,19,78].
One of the most common signs of capacitation is the degradation of cholesterol from the sperm plasma membrane. Previous studies have recorded a decline in cholesterol concentration parallel to a higher content of phospholipids and triglycerol after freezing/thawing as a consequence of a slow diffusion from cells. The membrane of cryocapacitated spermatozoa may be characterized by disintegration and swelling, changes in fluidity and permeability, phospholipid aggregation, and a decrease in the motility, viability, and enzymatic metabolism as observed in bulls [20], stallions [66], and boars [81]. These cryo-induced membrane modifications increase the amount of capacitated and acrosome-reacted spermatozoa after thawing. While these do not affect the post-thaw motility, they decrease the longevity of male gametes and their subsequent ability to fertilize the ovum [19,78,85].

6. Future Strategies against Cryodamage

Several strategies for the reduction in the negative impact of cryopreservation on spermatozoa have been developed through the years. Cryoinjury is a result of phase changes of water at low temperatures between the intra- or extra-cellular space. It was observed that spermatozoa are mainly sensitive to a fast temperature reduction between 5 and 25 °C, which leads to cold shock [19,85]. Cryo-induced damage is often associated with osmotic rupture of the cells due to the formation of intracellular ice crystals during cryopreservation, which leads to the loss of semipermeable properties of the sperm plasmatic membrane, release of intracellular enzymes, and ion redistribution. The cryotolerance and cold susceptibility of spermatozoa depend on the lipid content of the plasmatic membrane, which is composed mainly of omega-3/omega-6 fatty acids, especially docosahexaenoic (ω-3 PUFA) and docosapentaenoic (ω-6 PUFA), which belong to the family of PUFAs [77,82]. However, a difference in the lipid profiles between species has been observed. As a supplement for the protection of spermatozoa against cryogenic temperatures, cryoprotective additives such as antioxidants, fatty acids, antifreeze proteins, and essential oils or bioactive substances from plants may be used [86,87,88].
In general, cryoprotectants should have certain properties in order to secure the protection against cryoinjury such as biological accessibility, cell penetration ability, and low toxicity. Based on their ability to penetrate the cells, the cryoprotectants are divided into membrane-permeable (dimethyl sulfoxide—DMSO, glycerol)—able to replace water in the cell, and nonpermeable (egg yolk, raffinose, albumin)—unable to pass through the membranes [89].
A possible antioxidant supplementation could provide protection against the generation of ROS and subsequent OS development. The antioxidants used in freezing extenders may exhibit enzymatic or nonenzymatic activity. Glutathione peroxidase (GPx), superoxide dismutase (SOD), and catalase (CAT) represent the major enzymatic antioxidants. SOD has the ability to catalyze the dismutation of O2•− into H2O2 [2,90,91]. The primary role of GPx and CAT is the conversion of H2O2 to water and molecular oxygen. Nonenzymatic antioxidants include vitamin C (ascorbic acid), vitamin E (α-tocopherol), vitamin B12 (cobalamin), and glutathione (GSH) [86,90,91]. This complex of free radical scavengers works as an effective defensive system, which protects the membranes of spermatozoa against ROS (vitamin E and C) [92,93] or cryodamage during the freezing/thawing process (vitamin B12) [94].
To inhibit ice crystals formation and decrease the freezing point, antifreeze proteins (AFPs) or glycoproteins are often used. The principle lies in the stabilization of phospholipids and unsaturated fatty acids in the plasma membrane, which improves the membrane integrity and osmotic resistance during the cryopreservation process [73]. According to previous findings, AFPs have significant positive effects on the motility and viability of frozen–thaw spermatozoa in a variety of species such as rams [95], rabbits [96], buffaloes [97], and chimpanzees [98]. Nevertheless, the use of AFPs depends on the type and concentration of selected proteins. Zilli et al. [99] reported that the application of AFPIII (antifreeze protein type III) preserved the sea bream sperm protein profile during cryopreservation. Interestingly, supplementation with AFPIII and AFPI (antifreeze protein type I) significantly increased the membrane integrity but had no effect on the motility in fish [75,99].
Another possibility by which to protect spermatozoa against cryoinjury lies in the application of bioactive molecules as supplements to the freezing extender. Substances such as carotenoids, flavonoids, polyphenols, phytosterols, and phytoestrogens have a natural origin and present with the ability to penetrate the cell membranes and modulate numerous intracellular and metabolic processes [100]. Tvrdá et al. suggested a possible application of various bioactive compounds such as curcumin (CUR), lycopene (LYC), or epicatechin (EPI) to prevent cryo-induced oxidative stress in bovine spermatozoa. CUR is a bright yellow phenolic compound that comes from Curcuma longa and is a member of the ginger family. CUR acted as an effective ROS scavenger and inhibitor of LPO. Cryopreservation medium enriched with 50 µmol/L of CUR preserved the structural and functional characteristics, as well as the oxidative profile of cryopreserved bovine spermatozoa [101]. The flavonoid compound EPI belongs to the family of catechins naturally found in green tea, cocoa, and grapes. As with CUR, EPI was able to inhibit LPO and exhibited ROS scavenging properties particularly against O2•− and H2O2. The presence of 100 µmol/L significantly decreased the generation of ROS, the amount of protein carbonyls, and the level of LPO [102]. LYC was also considered as a potential cryosupplement because of its antioxidant effects. As a natural bright red carotenoid, LYC may be found in a wide spectrum of plants such as tomatoes, carrots, or grapefruits. Following the addition of 1.5 mmol/L of LYC into a commercial semen extender, the production of ROS and intracellular O2•− was decreased in comparison to the control [103]. These findings confirmed the beneficial antioxidant and ROS scavenging properties of selected natural bioactive compounds (CUR, EPI, LYC), which could at least partially eliminate oxidative damage during cryopreservation and support the vitality of male gametes after thawing.
For the inhibition of capacitation-like changes, melatonin (MEL) was also considered as a primary mitochondria-targeted antioxidant, which can detoxify cells at the mitochondrial level [104]. According to Carvajal-Serna et al. [105], ram semen samples incubated with 100 pmol/L or 1 µmol/L of MEL presented with a significantly higher amount of noncapacitated spermatozoa and a lower rate of male gametes undergoing capacitation. MEL also worked as an inhibitor of phosphatidylserine translocation, which is indicative of the initiation of apoptosis. In addition, MEL significantly decreased the level of protein tyrosine phosphorylation and cAMP in capacitated ram spermatozoa by the presence of cAMP-elevating agents in the medium [106].

7. Conclusions

In summary, we may conclude that capacitation as well as cryocapacitation are both oxidative events. The physiological concentration of ROS may promote the capacitation process. ROS are important primarily at the beginning of capacitation as regulators of cholesterol efflux, activation of the cAMP/PKA pathway, tyrosine phosphorylation, or redox signaling. However, the cryoinduced generation of ROS can accelerate OS by attacking PUFAs, which may lead to a serious membrane damage and LPO. It is clear that capacitation takes place mainly on the level of cell membranes. Meanwhile, cryo-induced capacitation in combination with high concentrations of ROS destabilizes the plasmatic membrane and reduces the fertility rate of thawed spermatozoa used for artificial insemination.
Nowadays, the application of supplements with antioxidant properties for the prevention of potential detrimental effects of ROS is very popular. These may work as effective ROS scavengers, which help to preserve the quality of cryopreserved spermatozoa in a multitude of species. Despite all of this, there are still numerous questions in the field of cryobiology, particularly regarding the process of cryocapacitation and changes associated with it. The exact course of cryocapacitation is still unclear, which is the reason why it is important to study this complex process of sperm metamorphosis from a molecular point of view.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/oxygen2020014/s1, Confirmation of Publication and Licensing Rights.

Author Contributions

Conceptualization, F.B. and E.T.; writing—original draft preparation, F.B., E.T. and M.Ď.; writing—review and editing, F.B., Š.B. and E.T.; supervision, N.L.; project administration, N.L. and E.T.; funding acquisition, N.L. and E.T. All authors have read and agreed to the published version of the manuscript.

Funding

This publication was supported by the Operational program Integrated Infrastructure within the project: Creation of nuclear herds of dairy cattle with a requirement for high health status through the use of genomic selection, innovative biotechnological methods, and optimal management of breeding, NUKLEUS 313011V387, cofinanced by the European Regional Development fund and by the Slovak Research and Development Agency grant no. APVV-21-0095.

Acknowledgments

We would like to extend our gratitude to the CeRA Team of excellence.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Thompson, A.; Agarwal, A.; Du Plessis, S.S. Physiological role of reactive oxygen species in sperm function: A review. In Antioxidants in Male Infertility: A Guide for Clinicians and Researchers; Parekattil, S.J., Agarwal, A., Eds.; Springer Science + Business Media: New York, NY, USA, 2013; pp. 69–89. [Google Scholar]
  2. Gosalvez, J.; Tvrda, E.; Agarwal, A. Free radical and superoxide reactivity detection in semen quality assessment: Past, present, and future. J. Assist. Reprod. Genet. 2017, 34, 697–707. [Google Scholar] [CrossRef] [PubMed]
  3. Pizzino, G.; Irrera, N.; Cucinotta, M.; Pallio, G.; Mannino, F.; Arcoraci, V.; Squadrito, F.; Altavilla, D.; Bitto, A. Oxidative stress: Harms and benefits for human health. Oxid. Med. Cell. Longev. 2017, 2017, 8416763. [Google Scholar] [CrossRef] [PubMed]
  4. Dutta, S.; Majzoub, A.; Agarwal, A. Oxidative stress and sperm function: A systematic review on evaluation and management. Arab J. Urol. 2019, 17, 87–97. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Phaniendra, A.; Jestadi, D.B.; Periyasamy, L. Free radicals: Properties, sources, targets, and their implication in various diseases. Indian J. Clin. Biochem. 2015, 30, 11–26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Kothari, S.; Thompson, A.; Agarwal, A.; du Plessis, S.S. Free radicals: Their beneficial and detrimental effects on sperm function. Indian J. Exp. Biol. 2010, 48, 425–435. [Google Scholar]
  7. Tvrdá, E.; Massanyi, P.; Lukáč, N. Physiological and pathological roles of free radicals in male reproduction. In Spermatozoa-Facts and Perspectives; Meccariello, R., Chianese, R., Eds.; IntechOpen: London, UK, 2017. [Google Scholar]
  8. Kobayashi, C.I.; Suda, T. Regulation of reactive oxygen species in stem cells and cancer stem cells. J. Cell. Physiol. 2012, 227, 421–430. [Google Scholar] [CrossRef]
  9. Chen, H.; Zhao, H.X.; Huang, X.F.; Chen, G.W.; Yang, Z.X.; Sun, W.J.; Tao, M.H.; Yuan, Y.; Wu, J.Q.; Sun, F.; et al. Does high load of oxidants in human semen contribute to male factor infertility? Antioxid. Redox. Signal. 2012, 16, 754–759. [Google Scholar] [CrossRef]
  10. Aitken, R.J.; De Iuliis, G.N.; Finnie, J.M.; Hedges, A.; McLachlan, R.I. Analysis of the relationships between oxidative stress, DNA damage and sperm vitality in a patient population: Development of diagnostic criteria. Hum. Reprod. 2010, 25, 2415–2426. [Google Scholar] [CrossRef] [Green Version]
  11. Sabeti, P.; Pourmasumi, S.; Rahiminia, T.; Akyash, F.; Talebi, A.R. Etiologies of sperm oxidative stress. Int. J. Reprod. Biomed. 2016, 14, 231. [Google Scholar] [CrossRef]
  12. Lee, D.; Moawad, A.R.; Morielli, T.; Fernandez, M.C.; O’Flaherty, C. Peroxiredoxins prevent oxidative stress during human sperm capacitation. MHR: Basic Sci. Reprod. Med. 2017, 23, 106–115. [Google Scholar] [CrossRef] [Green Version]
  13. Cormier, N.; Bailey, J.L. A differential mechanism is involved during heparin-and cryopreservation-induced capacitation of bovine spermatozoa. Biol. Reprod. 2003, 69, 177–185. [Google Scholar] [CrossRef] [PubMed]
  14. Aguiar, G.B.; Caldas-Bussiere, M.C.; Maciel, V.L.; Carvalho, C.S.P.D.; Souza, C.L.M.D. Association of L-arginine with heparin on the sperm capacitation improves in vitro embryo production in bovine. Anim. Reprod. 2019, 16, 938–944. [Google Scholar] [CrossRef] [PubMed]
  15. O’Flaherty, C.; de Lamirande, E.; Gagnon, C. Positive role of reactive oxygen species in mammalian sperm capacitation: Triggering and modulation of phosphorylation events. Free Radic. Biol. Med. 2006, 41, 528–540. [Google Scholar] [CrossRef] [PubMed]
  16. Bollwein, H.; Fuchs, I.; Koess, C. Interrelationship between plasma membrane integrity, mitochondrial membrane potential and DNA fragmentation in cryopreserved bovine spermatozoa. Reprod. Domest. Anim. 2008, 43, 189–195. [Google Scholar] [CrossRef] [PubMed]
  17. Bailey, J.L.; Bilodeau, J.F.; Cormier, N. Semen cryopreservation in domestic animals: A damaging and capacitating phenomenon. J. Androl. 2000, 21, 1–7. [Google Scholar]
  18. Sharma, V. Sperm storage for cancer patients in the UK: A review of current practice. Hum. Reprod. 2011, 26, 2935–2943. [Google Scholar] [CrossRef]
  19. O’Connell, M.; Mcclure, N.; Lewis, S.E.M. The effects of cryopreservation on sperm morphology, motility and mitochondrial function. Hum. Reprod. 2002, 17, 704–709. [Google Scholar] [CrossRef]
  20. Srivastava, N.; Srivastava, S.K.; Ghosh, S.K.; Kumar, A.; Perumal, P.; Jerome, A. Acrosome membrane integrity and cryocapacitation are related to cholesterol content of bull spermatozoa. Asian Pac. J. Reprod. 2013, 2, 126–131. [Google Scholar] [CrossRef]
  21. Li, X.; Ni, M.; Xing, S.; Yu, Y.; Zhou, Y.; Yang, S.; Li, H.; Zhu, R.; Han, M. Reactive Oxygen Species Secreted by Leukocytes in Semen Induce Self-Expression of Interleukin-6 and Affect Sperm Quality. Am. J. Men’s Health 2020, 14, 1557988320970053. [Google Scholar] [CrossRef]
  22. Agarwal, A.; Makker, K.; Sharma, R. Clinical relevance of oxidative stress in male factor infertility: An update. Am. J. Reprod. Immunol. 2008, 59, 2–11. [Google Scholar] [CrossRef]
  23. Jung, J.H.; Kim, M.H.; Kim, J.; Baik, S.K.; Koh, S.B.; Park, H.J.; Seo, J.T. Treatment of leukocytospermia in male infertility: A systematic review. World J. Men’s Health 2016, 34, 165–172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. 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] [PubMed] [Green Version]
  25. Cooper, T.G. The epididymis, cytoplasmic droplets and male fertility. Asian J. Androl. 2011, 13, 130–138. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. 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] [PubMed]
  27. Zini, A.; Defreitas, G.; Freeman, M.; Hechter, S.; Jarvi, K. Varicocele is associated with abnormal retention of cytoplasmic droplets by human spermatozoa. Fertil. Steril. 2000, 74, 461–464. [Google Scholar] [CrossRef]
  28. Moein, M.R.; Soleimani, M.; Tabibnejad, M.D. Reactive oxygen species (ROS) production in seminal fluid correlate with the severity of varicocele in infertile men. Iran. J. Reprod. Med. 2008, 6, 65–69. [Google Scholar]
  29. Park, J.; Rho, H.K.; Kim, K.H.; Choe, S.S.; Lee, Y.S.; Kim, J.B. Overexpression of glucose-6-phosphate dehydrogenase is associated with lipid dysregulation and insulin resistance in obesity. Mol. Cell. Biol. 2005, 25, 5146–5157. [Google Scholar] [CrossRef] [Green Version]
  30. Marchetti, C.; Obert, G.; Deffosez, A.; Formstecher, P.; Marchetti, P. Study of mitochondrial membrane potential, reactive oxygen species, DNA fragmentation and cell viability by flow cytometry in human sperm. Hum. Reprod. 2002, 17, 1257–1265. [Google Scholar] [CrossRef] [Green Version]
  31. Amaral, A.; Lourenço, B.; Marques, M.; Ramalho-Santos, J. Mitochondria functionality and sperm quality. Reproduction 2013, 146, R163–R174. [Google Scholar] [CrossRef] [Green Version]
  32. Li, X.; Fang, P.; Mai, J.; Choi, E.T.; Wang, H.; Yang, X.F. Targeting mitochondrial reactive oxygen species as novel therapy for inflammatory diseases and cancers. J. Hematol. Oncol. 2013, 6, 19. [Google Scholar] [CrossRef] [Green Version]
  33. Cadenas, E.; Davies, K.J. Mitochondrial free radical generation, oxidative stress, and aging. Free Radic. Biol. Med. 2000, 29, 222–230. [Google Scholar] [CrossRef]
  34. Chianese, R.; Pierantoni, R. Mitochondrial Reactive Oxygen Species (ROS) Production Alters Sperm Quality. Antioxidants 2021, 10, 92. [Google Scholar] [CrossRef] [PubMed]
  35. Aitken, R.J.; Whiting, S.; De Iuliis, G.N.; McClymont, S.; Mitchell, L.A.; Baker, M.A. Electrophilic aldehydes generated by sperm metabolism activate mitochondrial reactive oxygen species generation and apoptosis by targeting succinate dehydrogenase. J. Biol. Chem. 2012, 287, 33048–33060. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Koppers, A.J.; Garg, M.L.; Aitken, R.J. Stimulation of mitochondrial reactive oxygen species production by unesterified, unsaturated fatty acids in defective human spermatozoa. Free Radic. Biol. Med. 2010, 48, 112–119. [Google Scholar] [CrossRef]
  37. Chang, M.C. Fertilizing capacity of spermatozoa deposited into the fallopian tubes. Nature 1951, 168, 697–698. [Google Scholar] [CrossRef]
  38. Austin, C.R. Observations on the penetration of the sperm into the mammalian egg. Aust. J. Biol. Sci. 1951, 4, 581–596. [Google Scholar] [CrossRef] [Green Version]
  39. Visconti, P.E. Understanding the molecular basis of sperm capacitation through kinase design. Proc. Natl. Acad. Sci. USA 2009, 106, 667–668. [Google Scholar] [CrossRef] [Green Version]
  40. Jin, S.K.; Yang, W.X. Factors and pathways involved in capacitation: How are they regulated? Oncotarget 2017, 8, 3600. [Google Scholar] [CrossRef] [Green Version]
  41. Aitken, R.J. Reactive oxygen species as mediators of sperm capacitation and pathological damage. Mol. Reprod. Dev. 2017, 84, 1039–1052. [Google Scholar] [CrossRef]
  42. O’Flaherty, C. Redox regulation of mammalian sperm capacitation. Asian J. Androl. 2015, 17, 583–590. [Google Scholar] [CrossRef]
  43. de Lamirande, E.; Gagnon, C. Capacitation-associated production of superoxide anion by human spermatozoa. Free Radic. Biol. Med. 1995, 18, 487–495. [Google Scholar] [CrossRef]
  44. Papas, M.; Catalan, J.; Barranco, I.; Arroyo, L.; Bassols, A.; Yeste, M.; Miró, J. Total and specific activities of superoxide dismutase (SOD) in seminal plasma are related with the cryotolerance of jackass spermatozoa. Cryobiology 2020, 92, 109–116. [Google Scholar] [CrossRef] [PubMed]
  45. de Lamirande, E.; Harakat, A.; Gagnon, C. Human sperm capacitation induced by biological fluids and progesterone, but not by NADH or NADPH, is associated with the production of superoxide anion. J. Androl. 1998, 19, 215–225. [Google Scholar]
  46. Roy, S.C.; Atreja, S.K. Effect of reactive oxygen species on capacitation and associated protein tyrosine phosphorylation in buffalo (Bubalus bubalis) spermatozoa. Anim. Reprod. Sci. 2008, 107, 68–84. [Google Scholar] [CrossRef] [PubMed]
  47. Herrero, M.B.; Chatterjee, S.; Lefièvre, L.; de Lamirande, E.; Gagnon, C. Nitric oxide interacts with the cAMP pathway to modulate capacitation of human spermatozoa. Free Radic. Biol. Med. 2000, 29, 522–536. [Google Scholar] [CrossRef]
  48. O’Flaherty, C.; Rodriguez, P.; Srivastava, S. L-arginine promotes capacitation and acrosome reaction in cryopreserved bovine spermatozoa. Biochim. Biophys. Acta—Gen. Subj. 2004, 1674, 215–221. [Google Scholar] [CrossRef]
  49. Rodriguez, P.C.; O’flaherty, C.M.; Beconi, M.T.; Beorlegui, N.B. Nitric oxide-induced capacitation of cryopreserved bull spermatozoa and assessment of participating regulatory pathways. Anim. Reprod. Sci. 2005, 85, 231–242. [Google Scholar] [CrossRef]
  50. Dutta, S.; Henkel, R.; Sengupta, P.; Agarwal, A. Physiological role of ROS in sperm function. In Male Infertility; Parekattil, S.J., Ed.; Springer Nature: Cham, Switzerland, 2020; pp. 337–345. [Google Scholar]
  51. Battistone, M.A.; Da Ros, V.G.; Salicioni, A.M.; Navarrete, F.A.; Krapf, D.; Visconti, P.E.; Cuasnicu, P.S. Functional human sperm capacitation requires both bicarbonate-dependent PKA activation and down-regulation of Ser/Thr phosphatases by Src family kinases. MHR: Basic Sci. Reprod. Med. 2013, 19, 570–580. [Google Scholar] [CrossRef] [Green Version]
  52. Ghanbari, H.; Keshtgar, S.; Kazeroni, M. Inhibition of the CatSper channel and NOX5 enzyme activity affects the functions of the progesterone-stimulated human sperm. Iran. J. Med. Sci. 2018, 43, 18. [Google Scholar]
  53. Miguel-Jiménez, S.; Pina-Beltrán, B.; Gimeno-Martos, S.; Carvajal-Serna, M.; Casao, A.; Pérez-Pe, R. NADPH Oxidase 5 and Melatonin: Involvement in Ram Sperm Capacitation. Front. Cell Dev. Biol. 2021, 9, 655794. [Google Scholar] [CrossRef]
  54. Moreno-Irusta, A.; Dominguez, E.M.; Marín-Briggiler, C.I.; Matamoros-Volante, A.; Lucchesi, O.; Tomes, C.N.; Trevino, C.L.; Buffone, M.G.; Lascano, R.; Losinno, O.; et al. Reactive oxygen species are involved in the signaling of equine sperm chemotaxis. Reproduction 2020, 159, 423–436. [Google Scholar] [CrossRef] [PubMed]
  55. Lewis, B.; Aitken, R.J. A redox-regulated tyrosine phosphorylation cascade in rat spermatozoa. J. Androl. 2001, 22, 611–622. [Google Scholar] [PubMed]
  56. O’Flaherty, C.; Beorlegui, N.; Beconi, M.T. Participation of superoxide anion in the capacitation of cryopreserved bovine sperm. Int. J. Androl. 2003, 26, 109–114. [Google Scholar] [CrossRef] [PubMed]
  57. O’Flaherty, C.M.; Beorlegui, N.B.; Beconi, M.T. Lactate dehydrogenase-C4 is involved in heparin-and NADH-dependent bovine sperm capacitation. Andrologia 2002, 34, 91–97. [Google Scholar] [CrossRef] [PubMed]
  58. Duan, C.; Goldberg, E. Inhibition of lactate dehydrogenase C4 (LDH-C4) blocks capacitation of mouse sperm in vitro. Cytogenet. Genome Res. 2003, 103, 352–359. [Google Scholar] [CrossRef] [PubMed]
  59. Lefièvre, L.; Chen, Y.; Conner, S.J.; Scott, J.L.; Publicover, S.J.; Ford, W.C.L.; Barratt, C.L. Human spermatozoa contain multiple targets for protein S-nitrosylation: An alternative mechanism of the modulation of sperm function by nitric oxide? Proteomics 2007, 7, 3066–3084. [Google Scholar] [CrossRef] [Green Version]
  60. Visconti, P.E.; Westbrook, V.A.; Chertihin, O.; Demarco, I.; Sleight, S.; Diekman, A.B. Novel signaling pathways involved in sperm acquisition of fertilizing capacity. J. Reprod. Immunol. 2002, 53, 133–150. [Google Scholar] [CrossRef]
  61. Staicu, F.D.; Martínez-Soto, J.C.; Canovas, S.; Matás, C. Nitric oxide-targeted protein phosphorylation during human sperm capacitation. Sci. Rep. 2021, 11, 20979. [Google Scholar] [CrossRef]
  62. Miraglia, E.; De Angelis, F.; Gazzano, E.; Hassanpour, H.; Bertagna, A.; Aldieri, E.; Revelli, A.; Ghigo, D. Nitric oxide stimulates human sperm motility via activation of the cyclic GMP/protein kinase G signaling pathway. Reproduction 2011, 141, 47–54. [Google Scholar] [CrossRef] [Green Version]
  63. Mostek, A.; Janta, A.; Majewska, A.; Ciereszko, A. Bull Sperm Capacitation Is Accompanied by Redox Modifications of Proteins. Int. J. Mol. Sci. 2021, 22, 7903. [Google Scholar] [CrossRef]
  64. Duan, J.; Gaffrey, M.J.; Qian, W.J. Quantitative proteomic characterization of redox-dependent post-translational modifications on protein cysteines. Mol. Biosyst. 2017, 13, 816–829. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Yoon, S.J.; Kwon, W.S.; Rahman, M.S.; Lee, J.S.; Pang, M.G. A novel approach to identifying physical markers of cryo-damage in bull spermatozoa. PLoS ONE 2015, 10, e0126232. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Prien, S.; Iacovides, S. Cryoprotectants & cryopreservation of equine semen: A review of industry cryoprotectants and the effects of cryopreservation on equine semen membranes. J. Dairy Vet. Anim. Res. 2016, 3, 1–8. [Google Scholar]
  67. Benson, J.D.; Woods, E.J.; Walters, E.M.; Critser, J.K. The cryobiology of spermatozoa. Theriogenology 2012, 78, 1682–1699. [Google Scholar] [CrossRef] [PubMed]
  68. Tatone, C.; Di Emidio, G.; Vento, M.; Ciriminna, R.; Artini, P.G. Cryopreservation and oxidative stress in reproductive cells. Gynecol. Endocrinol. 2010, 26, 563–567. [Google Scholar] [CrossRef]
  69. Su, L.J.; Zhang, J.H.; Gomez, H.; Murugan, R.; Hong, X.; Xu, D.; Jiang, F.; Peng, Z.Y. Reactive oxygen species-induced lipid peroxidation in apoptosis, autophagy, and ferroptosis. Oxid. Med. Cell. Longev. 2019, 2019, 5080843. [Google Scholar] [CrossRef] [Green Version]
  70. Sanocka, D.; Kurpisz, M. Reactive oxygen species and sperm cells. Reprod. Biol. Endocrinol. 2004, 2, 12. [Google Scholar] [CrossRef] [Green Version]
  71. Brouwers, J.F.; Gadella, B.M. In situ detection and localization of lipid peroxidation in individual bovine sperm cells. Free Radic. Biol. Med. 2003, 35, 1382–1391. [Google Scholar] [CrossRef]
  72. Dietrich, M.A.; Irnazarow, I.; Ciereszko, A. Proteomic identification of seminal plasma proteins related to the freezability of carp semen. J. Proteom. 2017, 162, 52–61. [Google Scholar] [CrossRef]
  73. Li, P.; Hulak, M.; Koubek, P.; Sulc, M.; Dzyuba, B.; Boryshpolets, S.; Rodina, M.; Gela, D.; Manaskova-Postlerova, P.; Peknicova, J.; et al. Ice-age endurance: The effects of cryopreservation on proteins of sperm of common carp, Cyprinus carpio L. Theriogenology 2010, 74, 413–423. [Google Scholar] [CrossRef]
  74. Hill, W.G.; Kaetzel, M.A.; Kishore, B.K.; Dedman, J.R.; Zeidel, M.L. Annexin A4 reduces water and proton permeability of model membranes but does not alter aquaporin 2–mediated water transport in isolated endosomes. J. Gen. Physiol. 2003, 121, 413–425. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Xin, M.; Niksirat, H.; Shaliutina-Kolešová, A.; Siddique, M.A.M.; Sterba, J.; Boryshpolets, S.; Linhart, O. Molecular and subcellular cryoinjury of fish spermatozoa and approaches to improve cryopreservation. Rev. Aquac. 2020, 12, 909–924. [Google Scholar] [CrossRef]
  76. Aitken, R.J.; Gibb, Z.; Baker, M.A.; Drevet, J.; Gharagozloo, P. Causes and consequences of oxidative stress in spermatozoa. Reprod. Fertil. Dev. 2016, 28, 1–10. [Google Scholar] [CrossRef]
  77. Aitken, R.J. Free radicals, lipid peroxidation and sperm function. Reprod. Fertil. Dev. 1995, 7, 659–668. [Google Scholar] [CrossRef] [PubMed]
  78. Talukdar, D.J.; Ahmed, K.; Talukdar, P. Cryocapacitation and fertility of cryopreserved semen. Int. J. Livest. Res. 2015, 5, 11. [Google Scholar] [CrossRef] [Green Version]
  79. Kadirvel, G.; Kathiravan, P.; Kumar, S. Protein tyrosine phosphorylation and zona binding ability of in vitro capacitated and cryopreserved buffalo spermatozoa. Theriogenology 2011, 75, 1630–1639. [Google Scholar] [CrossRef]
  80. Thomas, A.D.; Meyers, S.A.; Ball, B.A. Capacitation-like changes in equine spermatozoa following cryopreservation. Theriogenology 2006, 65, 1531–1550. [Google Scholar] [CrossRef]
  81. Gadella, B.M.; Harrison, R.A.P. Capacitation induces cyclic adenosine 3′, 5′-monophosphate-dependent, but apoptosis-unrelated, exposure of aminophospholipids at the apical head plasma membrane of boar sperm cells. Biol. Reprod. 2002, 67, 340–350. [Google Scholar] [CrossRef] [Green Version]
  82. Cross, N.L. Decrease in order of human sperm lipids during capacitation. Biol. Reprod. 2003, 69, 529–534. [Google Scholar] [CrossRef] [Green Version]
  83. Bailey, J.L.; Berube, B. Mechanisms of cryopreservation-induced capacitation of bovine sperm. J. Androl. 1998, 20, 13. [Google Scholar]
  84. Galantino-Homer, H.L.; Visconti, P.E.; Kopf, G.S. Regulation of protein tyrosine phosphorylation during bovine sperm capacitation by a cyclic adenosine 3’, 5’-monophosphate-dependent pathway. Biol. Reprod. 1997, 56, 707–719. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Watson, P.F. The causes of reduced fertility with cryopreserved semen. Anim. Reprod. Sci. 2000, 60–61, 481–492. [Google Scholar] [CrossRef]
  86. Hezavehei, M.; Sharafi, M.; Kouchesfahani, H.M.; Henkel, R.; Agarwal, A.; Esmaeili, V.; Shahverdi, A. Sperm cryopreservation: A review on current molecular cryobiology and advanced approaches. Reprod. Biomed. Online 2018, 37, 327–339. [Google Scholar] [CrossRef] [PubMed]
  87. Gautier, C.; Aurich, C. “Fine feathers make fine birds”–The mammalian sperm plasma membrane lipid composition and effects on assisted reproduction. Anim. Reprod. Sci. 2021, 106884. [Google Scholar] [CrossRef]
  88. Aboagla, E.M.E.; Terada, T. Effects of egg yolk during the freezing step of cryopreservation on the viability of goat spermatozoa. Theriogenology 2004, 62, 1160–1172. [Google Scholar] [CrossRef]
  89. Jang, T.H.; Park, S.C.; Yang, J.H.; Kim, J.Y.; Seok, J.H.; Park, U.S.; Choi, C.W.; Lee, S.R.; Han, J. Cryopreservation and its clinical applications. Integr. Med. Res. 2017, 6, 12–18. [Google Scholar] [CrossRef]
  90. Cilio, S.; Rienzo, M.; Villano, G.; Mirto, B.F.; Giampaglia, G.; Capone, F.; Ferretti, G.; Di Zazzo, E.; Crocetto, F. Beneficial Effects of Antioxidants in Male Infertility Management: A Narrative Review. Oxygen 2022, 2, 1–11. [Google Scholar] [CrossRef]
  91. Papas, M.; Catalán, J.; Fernandez-Fuertes, B.; Arroyo, L.; Bassols, A.; Miró, J.; Yeste, M. Specific activity of superoxide dismutase in stallion seminal plasma is related to sperm cryotolerance. Antioxidants 2019, 8, 539. [Google Scholar] [CrossRef] [Green Version]
  92. Ener, K.; Aldemir, M.; Işık, E.; Okulu, E.M.R.A.H.; Özcan, M.F.; Uğurlu, M.; Tangal, A.; Özayar, A. The impact of vitamin E supplementation on semen parameters and pregnancy rates after varicocelectomy: A randomised controlled study. Andrologia 2016, 48, 829–834. [Google Scholar] [CrossRef]
  93. Cyrus, A.; Kabir, A.; Goodarzi, D.; Moghimi, M. The effect of adjuvant vitamin C after varicocele surgery on sperm quality and quantity in infertile men: A double blind placebo controlled clinical trial. Int. Braz. J. Urol. 2015, 41, 230–238. [Google Scholar] [CrossRef] [Green Version]
  94. Hosseinabadi, F.; Jenabi, M.; Ghafarizadeh, A.A.; Yazdanikhah, S. The effect of vitamin B12 supplement on post-thaw motility, viability and DNA damage of human sperm. Andrologia 2020, 52, e13877. [Google Scholar] [CrossRef] [PubMed]
  95. Payne, S.R.; Oliver, J.E.; Upreti, G.C. Effect of antifreeze proteins on the motility of ram spermatozoa. Cryobiology 1994, 31, 180–184. [Google Scholar] [CrossRef] [PubMed]
  96. Nishijima, K.; Tanaka, M.; Sakai, Y.; Koshimoto, C.; Morimoto, M.; Watanabe, T.; Fan, J.; Kitajima, S. Effects of type III antifreeze protein on sperm and embryo cryopreservation in rabbit. Cryobiology 2014, 69, 22–25. [Google Scholar] [CrossRef]
  97. Qadeer, S.; Khan, M.A.; Ansari, M.S.; Rakha, B.A.; Ejaz, R.; Iqbal, R.; Younis, M.; Ullah, N.; DeVries, A.L.; Akhter, S. Efficiency of antifreeze glycoproteins for cryopreservation of Nili-Ravi (Bubalus bubalis) buffalo bull sperm. Anim. Reprod. Sci. 2015, 157, 56–62. [Google Scholar] [CrossRef]
  98. Younis, A.I.; Rooks, B.; Khan, S.; Gould, K.G. The effects of antifreeze peptide III (AFP) and insulin transferrin selenium (ITS) on cryopreservation of chimpanzee (Pan troglodytes) spermatozoa. J. Androl. 1998, 19, 207–214. [Google Scholar] [PubMed]
  99. Zilli, L.; Beirão, J.; Schiavone, R.; Herraez, M.P.; Gnoni, A.; Vilella, S. Comparative proteome analysis of cryopreserved flagella and head plasma membrane proteins from sea bream spermatozoa: Effect of antifreeze proteins. PLoS ONE 2014, 9, e99992. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Hamzalıoğlu, A.; Gökmen, V. Interaction between bioactive carbonyl compounds and asparagine and impact on acrylamide. In Acrylamide in Food: Analysis, Content and Potential Health Effects; Gökmen, V., Ed.; Academic Press: Cambridge, MA, USA, 2016; pp. 355–376. [Google Scholar]
  101. Tvrdá, E.; Greifová, H.; Mackovich, A.; Hashim, F.; Lukáč, N. Curcumin offers antioxidant protection to cryopreserved bovine semen. Czech J. Anim. Sci. 2018, 63, 247–255. [Google Scholar] [CrossRef] [Green Version]
  102. Tvrda, E.; Straka, P.; Galbavy, D.; Ivanic, P. Epicatechin provides antioxidant protection to bovine spermatozoa subjected to induced oxidative stress. Molecules 2019, 24, 3226. [Google Scholar] [CrossRef] [Green Version]
  103. Tvrda, E.; Mackovich, A.; Greifova, H.; Hashim, F.; Lukac, N. Antioxidant effects of lycopene on bovine sperm survival and oxidative profile following cryopreservation. Vet. Med. 2017, 62, 429–436. [Google Scholar] [CrossRef] [Green Version]
  104. Balao da Silva, C.M.; Macías-García, B.; Miró-Morán, A.; González-Fernández, L.; Morillo-Rodriguez, A.; Ortega-Ferrusola, C.; Gallardo-Bolaños, J.M.; Stilwell, G.; Tapia, J.A.; Peña, F.J. Melatonin reduces lipid peroxidation and apoptotic-like changes in stallion spermatozoa. J. Pineal Res. 2011, 51, 172–179. [Google Scholar] [CrossRef]
  105. Carvajal-Serna, M.; Cardozo-Cerquera, J.A.; Grajales-Lombana, H.A.; Casao, A.; Pérez-Pe, R. Sperm Behavior and Response to Melatonin under Capacitating Conditions in Three Sheep Breeds Subject to the Equatorial Photoperiod. Animals 2021, 11, 1828. [Google Scholar] [CrossRef] [PubMed]
  106. Gimeno-Martos, S.; Casao, A.; Yeste, M.; Cebrián-Pérez, J.A.; Muiño-Blanco, T.; Pérez-Pé, R. Melatonin reduces cAMP-stimulated capacitation of ram spermatozoa. Reprod. Fertil. Dev. 2019, 31, 420–431. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Differences and similarities between capacitation and cryocapacitation of spermatozoa. Created with (Supplementary: Confirmation of Publication and Licensing Rights) BioRender.com (accessed on 23 May 2022).
Figure 1. Differences and similarities between capacitation and cryocapacitation of spermatozoa. Created with (Supplementary: Confirmation of Publication and Licensing Rights) BioRender.com (accessed on 23 May 2022).
Oxygen 02 00014 g001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Benko, F.; Ďuračka, M.; Baňas, Š.; Lukáč, N.; Tvrdá, E. Biological Relevance of Free Radicals in the Process of Physiological Capacitation and Cryocapacitation. Oxygen 2022, 2, 164-176. https://0-doi-org.brum.beds.ac.uk/10.3390/oxygen2020014

AMA Style

Benko F, Ďuračka M, Baňas Š, Lukáč N, Tvrdá E. Biological Relevance of Free Radicals in the Process of Physiological Capacitation and Cryocapacitation. Oxygen. 2022; 2(2):164-176. https://0-doi-org.brum.beds.ac.uk/10.3390/oxygen2020014

Chicago/Turabian Style

Benko, Filip, Michal Ďuračka, Štefan Baňas, Norbert Lukáč, and Eva Tvrdá. 2022. "Biological Relevance of Free Radicals in the Process of Physiological Capacitation and Cryocapacitation" Oxygen 2, no. 2: 164-176. https://0-doi-org.brum.beds.ac.uk/10.3390/oxygen2020014

Article Metrics

Back to TopTop