Next Article in Journal
How Do Young and Old Spontaneously Hypertensive Rats Respond to Antihypertensive Therapy? Comparative Studies on the Effects of Combined Captopril and Nifedipine Treatment
Next Article in Special Issue
Potential Antioxidant Multitherapy against Complications Occurring in Sepsis
Previous Article in Journal
Investigation of Cerebral Autoregulation Using Time-Frequency Transformations
Previous Article in Special Issue
The Identification of Prohibitin in the Rat Heart Mitochondria in Heart Failure
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Antioxidant Intervention against Male Infertility: Time to Design Novel Strategies

1
Molecular and Clinical Pharmacology Program, Institute of Biomedical Sciences, Faculty of Medicine, University of Chile, Santiago 8380000, Chile
2
Urology Department, University of Chile Clinical Hospital, Santiago 8380000, Chile
3
Andrology Unit, Shady Grove Fertility, Santiago 7650672, Chile
4
Department of Biochemical Sciences “A. Rossi-Fanelli”, Sapienza University of Rome, 00185 Rome, Italy
5
Department of Physiology and Pharmacology “Vittorio Erspamer”, Faculty of Pharmacy and Medicine, Sapienza University, 00185 Rome, Italy
*
Author to whom correspondence should be addressed.
Biomedicines 2022, 10(12), 3058; https://doi.org/10.3390/biomedicines10123058
Submission received: 30 September 2022 / Revised: 21 November 2022 / Accepted: 24 November 2022 / Published: 28 November 2022

Abstract

:
Infertility is a highly prevalent condition, affecting 9–20% of couples worldwide. Among the identifiable causes, the male factor stands out in about half of infertile couples, representing a growing problem. Accordingly, there has been a decline in both global fertility rates and sperm counts in recent years. Remarkably, nearly 80% of cases of male infertility (MI) have no clinically identifiable aetiology. Among the mechanisms likely plausible to account for idiopathic cases, oxidative stress (OS) has currently been increasingly recognized as a key factor in MI, through phenomena such as mitochondrial dysfunction, lipid peroxidation, DNA damage and fragmentation and finally, sperm apoptosis. In addition, elevated reactive oxygen species (ROS) levels in semen are associated with worse reproductive outcomes. However, despite an increasing understanding on the role of OS in the pathophysiology of MI, therapeutic interventions based on antioxidants have not yet provided a consistent benefit for MI, and there is currently no clear consensus on the optimal antioxidant constituents or regimen. Therefore, there is currently no applicable antioxidant treatment against this problem. This review presents an approach aimed at designing an antioxidant strategy based on the particular biological properties of sperm and their relationships with OS.

1. Introduction

Infertility is defined as the failure to achieve a spontaneous pregnancy despite 1 year of practicing regular sexual intercourse without a contraceptive [1]. It affects 9 to 20% of all couples [2] and a male factor is an identifiable cause in 50% of infertile couples [2,3]. Furthermore, current evidence suggests a decline in global fertility rates worldwide from 4.7 to 2.4 live births between 1950 and 2017 [4]. In addition, there has been a parallel decline in sperm counts (0.70 million/mL/year) observed between 1981 and 2013 [5].
Identifiable causes of male infertility comprise as few as 10% of the total cases, with around 5% being sperm transport disorders [6] and the other 5% being endocrine and systemic disorders with hypogonadotropic hypogonadism [6,7,8,9]. Most infertile men fall into two categories: primary testicular defects in spermatogenesis (65–80%, with the majority of these cases being idiopathic dysspermatogenesis, an isolated defect in spermatogenesis without an identifiable cause) and idiopathic or unexplained male infertility (15–30%) [10], in which an infertile man presents with normal seminal fluid analysis and no apparent cause for infertility [11].
The redox state of a biological system can be defined as the balance between its oxidative elements and the capacity of its antioxidant elements. The main oxidant elements are reactive oxygen species (ROS), a group of highly reactive oxidant agents, including radical anion superoxide, hydrogen peroxide, hydroxyl radical and singlet oxygen. At low concentrations, ROS are actually essential for normal cell homeostasis and are involved in various cell signalling pathways [12]. Conversely, when antioxidant systems of the cell are overwhelmed by ROS concentrations (a state known as oxidative stress [OS]), oxidation of biomolecules can occur. Consequently, resulting pathophysiological events can range from cell dysfunction to cell death [13]. Thus, ROS have been identified as a key factor in a variety of pathologies such as atherosclerosis [14], arterial hypertension [15], atrial fibrillation [16], heart failure [17], stroke [18], metabolic syndrome [19], NAFLD [20], Alzheimer’s disease [21] and other neurodegenerative diseases [22], among others.
Despite its epidemiological relevance and its widely accepted association to OS, no antioxidant strategy has shown consistent benefit on male infertility. This review focuses on the role of OS in idiopathic dysspermatogenesis and unexplained male infertility, and new approaches on the development of therapeutic strategies against male infertility.

2. Relationship between Male Infertility and Reactive Oxygen Species

In 1994, Aitken proposed the loss of sperm function subsequent to the induction of peroxidative damage by ROS originating from the spermatozoa as a theory for male infertility [23]. Since then, there has been a growing body of evidence pointing to OS as a key phenomenon in the pathophysiology of male infertility. A recent update on male infertility by the European Association of Urology discusses the increasing evidence suggesting that in the setting of unexplained infertility (men with normal sperm parameters but unable to conceive with a healthy female partner) can present abnormalities at the molecular level, such as increased sperm DNA fragmentation (SDF) [24,25,26] and elevated seminal reactive oxygen species (ROS) [26], which appear to result in worse reproductive outcomes in couples [24,26,27,28,29,30,31].
Before further exploring the pathophysiological mechanisms behind the relationship between OS and male infertility, it is imperative to properly characterise the oxidant and antioxidant elements playing in the particular setting of human spermatozoa.

3. ROS and Sperm Physiology

The cellular oxidative homeostasis in spermatozoa relies on two groups of components that interact in maintaining this balance: ROS and antioxidants.
Among reactive oxygen species that standout within the biology of spermatozoa are (1) superoxide radical anion (O2), primarily originating in mitochondria via the slippage of an electron from the electron transport chain (ETC) to molecular oxygen during oxidative phosphorylation, (2) hydrogen peroxide (H2O2) and the (3) hydroxyl radical (·OH) [32].
Main endogenous ROS sources include: (1) mitochondrial ETC and (2) membrane-associated reduced nicotinamide adenine dinucleotide phosphate (NADPH) oxidase. Of the aforementioned, the main endogenous source of ROS is the latter [33]; the electron flux through the ETC, located in the inner mitochondrial membrane, is instrumental to oxidative phosphorylation and ATP synthesis. Despite being a highly efficient process, 1–2% of oxygen is reduced to O2. by single electron transfer mediated by complexes I and III of the ETC [34]. Hydrogen peroxide levels are also a product of this phenomenon, given the high activity of superoxide dismutase 2 (SOD2) in mitochondria, which dismutates O2 to H2O2 [35]. Meanwhile, hydroxyl radical is a product of the reaction of transition metals both with superoxide anion (Haber-Weiss reaction) [36] and hydrogen peroxide (Fenton reaction) [37].
Although NADPH oxidase is expressed in human spermatozoa in the form of NOX5 [38], the real impact of its activity both in physiological and pathophysiological settings is unknown [39]. Moreover, there is evidence showing that it lacks significant activity [40] and that the ROS producing activity of NOX5 is significantly lower than that of the NADPH-oxidase found in leukocytes [41]. Alternatively, it could be argued that NADPH oxidase found in leukocytes is a relevant source of ROS in sperm. Aitken & West conducted a study analysing the relationship between ROS production and leukocyte infiltration in 109 samples of human ejaculate from non-selected volunteers. Interestingly, 6 cases showed an association of elevated ROS production and leukocyte concentrations. However, in the majority of cases exhibiting high ROS production, leukocyte numbers were low or absent, suggesting that the relative contribution of leukocytes to ROS production in sperm is low [42]. Furthermore, the author propounds that, under in vivo conditions, the induction of OS on spermatozoa by infiltrating leukocytes is unlikely given the protective properties of seminal plasma [23].
Additionally, there are several environmental risk factors for male infertility whose deleterious effect over spermatozoa is attributable to their role as exogenous ROS sources, namely: obesity [43,44,45], smoking [46,47,48,49], alcohol use [50], ionising and non-ionising radiations [51], air pollution [52], among others.
As previously outlined, ROS are essential to proper cell functioning under physiological conditions, and male germ cells are not the exception. Low ROS levels are crucial for the acquisition of sperm functions during the final step of development of the mature spermatozoa. Each of the following processes take place as the spermatozoa transport through the female reproductive tract:
  • Capacitation: Spermatozoa undergo molecular modifications such as basification of intracellular pH, activation of c-AMP dependent pathways, removal of cholesterol from sperm membrane, and phosphorylation of serine, threonine and tyrosine residues in proteins [33]. Capacitation has been shown to be suppressed when spermatozoa are incubated with catalase, which is consistent with the finding that low levels of H2O2 result in a higher rate of capacitated spermatozoa [53]. Low levels of ROS have been shown to positively modulate several pathways involved in the molecular modifications in the capacitation process [54].
  • Hyperactivation: Hyperactivation is considered a subprocess of capacitation. When reaching the oocyte, spermatozoa exhibit a high amplitude, nonlinear flagellar movement which is instrumental to the impulse of spermatozoa through cumulus cells surrounding the oocyte. O2 concentrations seem to trigger this phenomenon [55]. Higher degrees of SOD activity were shown to block hyperactivation induced by O2 in semen samples [56].
  • Acrosome reaction: Consists of exocytosis of the acrosomal matrix rich in digestive enzymes such as acrosin and hyaluronidase, essential for sperm penetration across the cumulus cells and zona pellucida. Capacitation is a necessary condition for this process to take place. Low concentrations of O2 and H2O2 have been shown to possess a role in acrosome reaction [57,58].
  • Sperm-oocyte fusion during fertilisation: Membrane fluidity of spermatozoa is crucial for sperm-oocyte fusion. In turn, membrane fluidity is influenced by polyunsaturated fatty acid (PUFA) content and phospholipase A2, whose activity is upregulated by O2/H2O2-dependent kinase activation [59,60].

4. Antioxidants and Sperm Physiology

Conversely, antioxidants can be broadly categorised into two types: enzymatic and non-enzymatic. The three most relevant enzymes of the antioxidants system in semen are superoxide dismutase (SOD), catalase (CAT) and glutathione peroxidase (GPX) [47].
SOD activity is significant both in spermatozoa and seminal fluid, being considerably higher in the latter when compared to other extracellular fluids. Three isoforms of this enzyme have been identified in mammals: SOD1/CuZn-SOD (cytosolic), SOD2/Mn-SOD (mitochondrial) and SOD3/EC-SOD (extracellular) [61]. SOD dismutates O2 into molecular oxygen and H2O2. Due to its location, SOD2 seems particularly critical in the context of human spermatozoa since increased SOD2 expression is associated with a reduction of mitochondrial O2 levels, attenuated age-dependent increase of oxidative damage, and better preservation of mitochondrial function [62,63,64].
GPX comprises a group of enzymes in charge of catalysing the reduction of H2O2 to oxygen and water. The most relevant isoforms in spermatozoa seem to be GPX1, GPX4, and phospholipid hydroperoxide glutathione peroxidase (PHGPx, which uses selenium as cofactor), playing a role in the protection of structural integrity, motility and viability of sperm [65,66].
CAT catalyses the conversion of H2O2 to molecular oxygen and water. A constant, low level of activity of this enzyme has been described in the testicles of rats [67]. Furthermore, lower levels of seminal CAT have been found in asthenozoospermic men when compared with normozoospermic men, suggesting the relevance of this enzyme in male fertility [68].
Non-enzymatic antioxidants can also be endogenous or exogenous, being the most relevant in spermatozoa: glutathione, selenium, carotenoids such a lycopene, ascorbic acid and α-tocopherol, which exert their antioxidant effects through direct neutralisation, promoting expression of antioxidant enzymes, or acting as cofactors for antioxidant enzymes.
In a 2011 study by Moretti et al., the total antioxidant power of seminal plasma was estimated to be ten times higher than that of blood and comprising a combination of small-molecular-mass free radical scavengers, including vitamin C, uric acid, tyrosine, unspecified polyphenols, reduced glutathione, and hypotaurine. Additional small-molecular-mass scavengers, such as vitamin E, are also present in seminal plasma [69].
This schematic division of elements in the antioxidant system is for the sole purpose of simplifying their characterization: indeed, these elements are intertwined and work synergically.
It is interesting to note that, since low levels of ROS are necessary during processes such as capacitation and acrosome reaction, the overuse of antioxidants could be deleterious due to reduced ROS generation and intracellular concentration [70]. This excessive administration of antioxidants could shift the redox balance of the cell in favour of a predominantly reductive state, known as reductive stress (RS) [71], which could even result in impairment of spermiogram variables such as concentration, morphology and motility [70]. Moreover, vitamin C and vitamin E, both widely used in clinical trials for male infertility, have shown detrimental effects in the context of fertility [72], such as increased sperm decondensation in men treated with vitamins C and E, β-carotene, selenium and zinc [73]. Further research is needed in order to describe the optimal antioxidant doses, but adverse effects could be related to excessive dosing and lack of understanding of the synergistic effects of antioxidant compounds [74].

5. Role of ROS in the Pathophysiology of Male Infertility

In a 2021 review paper, Ritchie et al. described mitochondrial dysfunction, lipid peroxidation, DNA damage and finally, apoptosis as the main phenomena by which OS produces its deleterious effect over spermatozoa morphology and function [32].
Alternatively, Aitken and colleagues assign a key role to mitochondrial dysfunction as the initiating event, from which the rest follow: cellular, environmental and lifestyle related-signalling trigger the generation of mitochondrial ROS. As a consequence of this process, electrophilic aldehyde levels increase and bind to mitochondrial ETC elements, contributing to the persistence of the phenomenon. It is proposed that the self-perpetuating nature of this process reaches a threshold, leading to the loss of mitochondrial membrane potential, oxidative DNA damage, caspase activation, motility loss, phosphatidylserine externalisation, vitality loss and DNA fragmentation (see Figure 1) [75]. This perspective is consistent with the critical role of mitochondria as main ROS generator in male germ cells, which has been discussed in a previous section of this work.

5.1. Mitochondrial Dysfunction

Excessive mitochondrial ROS production is a milestone in the creation of the OS state that underlies male infertility. A crucial trait of this phenomenon is that it establishes a self-perpetuating, vicious cycle in which OS further promotes the generation of excessive ROS generation at a mitochondrial level. This self-perpetuating feature is enabled by the production of aldehydes such as acrolein, malondialdehyde and 4-hydroxynonenal generated by lipid peroxidation, which, in the first instance, carry out an oxidative attack on histidine, lysine and valine residues in mitochondrial proteins, particularly succinate dehydrogenase, an enzyme located in the inner mitochondrial membrane that participates in the ETC [76]. This leads to dysregulation of electron flow along the mitochondrial ETC and further production of ROS, perpetuating the cycle via a positive feedback circuit [54]. In this setting, any factor that breaks the redox balance of the male germ cell inducing OS is prone to trigger mitochondrial ROS generation and begin the previously mentioned cycle. These inducing factors include both exogenous ROS sources (smoking, obesity, radiation, pollution, etc.) and dietary patterns leading to a deficit in exogenous antioxidants (non-enzymatic antioxidants and antioxidant enzyme cofactors).

5.2. Lipid Peroxidation

Lipid peroxidation is a result of the oxidative attack of ROS over PUFAs, which are abundant molecules in spermatozoa membranes, particularly docosahexaenoic and arachidonic acids [77], since they are a factor that confers high fluidity to sperm plasma membranes, a crucial trait for fertilisation. Unfortunately, the chemical characteristics responsible for this trait, also make them a target for free radical attack and the induction of lipid peroxidation [39]. Once lipid peroxidation is initiated, lipid radicals react with adjacent carbons, leading to a peroxidation chain reaction. Quantification of lipid peroxidation is usually performed by measuring its end products. Levels of aldehydic products (malondialdehyde [MDA] and 4-hydroxyalkenals) are inversely correlated with sperm motility, viability, capacity for prolonged survival in vitro and competence for sperm-oocyte fusion [75]. This is supported by a study by Hosseinzadeh Colagar and colleagues [78], in which malondialdehyde levels (an end product of lipid peroxidation) were significantly lower in normospermic men than in men with asthenoteratozoospermia and oligoasthenoteratozoospermia and had negative correlation with sperm count, motility and morphology.

5.3. Apoptosis

As previously discussed, the self-perpetuating cycle of ROS generation at a mitochondrial level results in high levels of peroxidation, loss of mitochondrial membrane potential, membrane damage in the midpiece of the spermatozoa and decreased sperm motility [76,79]. Subsequently, caspase activation and phosphatidylserine exteriorization ensue, as shown by the activation of caspase 3 and annexin V binding following exposure to H2O2 [80]. This is consistent with a proposed ROS-mediated activation of MAPK pathways resulting in apoptosis in mice [81]. Further down this pathway, apoptosis in sperm starts to exhibit features different from those seen in somatic cell apoptosis. Although there is evidence of endonuclease activity (e.g., mitochondrial endonuclease G or cytosolic caspase-activated DNAse), these are unable to translocate to the nucleus due to the physical architecture of spermatozoa [39], since mitochondria and most of cytoplasm are concentrated in the midpiece, while nuclear DNA is located in the spermatozoon’s head.
Moreover, nuclear DNA sperm is found in a quasicrystalline state, making penetration by endonucleases highly unlikely [75], possibly taking part in perimortem DNA fragmentation after ROS-dependent membrane breakdown [82]. Furthermore, an endogenous nuclease has been identified in human spermatozoa, which shares with somatic cell nucleases a dependence on calcium and magnesium, thus making likely its involvement in the final stages of apoptosis [83].
The apoptotic process of sperm is critical in male infertility because of its close association to DNA damage, under the premise that apoptosis is initiated during spermatogenesis but the process is arrested due to the profound conformational changes of male germ cells following the meiotic process, which removes the apoptotic intracellular machinery. Consequently, DNA damage could also be attributed to an abortive apoptotic process [39,84].

5.4. DNA Damage and Sperm DNA Fragmentation

Spermatozoa are proposed to be particularly susceptible to DNA damage, given both a lesser degree of condensation for mitochondrial DNA and the absence of DNA repair systems for nuclear DNA [85]. Although most of the available evidence focuses on nuclear DNA damage in spermatozoa, it is worth noting that mitochondrial DNA damage has been associated with poor motility and reduced fertility [86].
Two main mechanisms for DNA damage in the context of male infertility have been described: endonuclease-mediated cleavage and ROS [75]. Interestingly, 8-hydroxy-2′-deoxyguanosine levels (8OHdG), a biomarker of oxidative DNA damage, has been found to be consistently elevated in the spermatozoa of infertile men. Simultaneously, DNA damage measured by terminal deoxynucleotidyl transferase dUTP nick end labelling (TUNEL) assay is strongly correlated to 8OHdG levels [87]. This correlation is highly indicative that DNA damage related to male infertility is primarily oxidative in origin [87,88].
Consistent with Aitkens’ hypothesis, an early stage in this process and probable triggering event seems to be the induction of O2 generation in sperm mitochondria [79,89].
Types of DNA damage identified on spermatozoa include base mismatch, abasic sites, base modifications, DNA adducts and crosslink, pyrimidine dimers, single strand breaks and double strand breaks. All of the above can lead to SDF, and except for base mismatch and pyrimidine dimers, all are induced by ROS overproduction [90].
Indeed, ROS overproduction induces SDF among other deleterious effects [39]. Several mechanisms explain the induction of SDF by ROS: direct oxidative damage, lipid peroxidation by-products, and induction of abortive apoptosis. Direct oxidative damage to DNA bases is thought to be inflicted mainly by mitochondria-originated H2O2, which is able to penetrate the nucleus due to its low molecular weight and lack of electric charge [75,91]. Furthermore, by-products of lipid peroxidation such as MDA and 4-hydroxynonenal (4HNE) have been shown cause DNA damage through the introduction of DNA adducts such as 8OHdG, 1,N6-ethenoadenosine, and 1,N6-ethenoguanosine [92,93,94]. The relationship between ROS and apoptosis as a source of DNA damage has been described in the previous section.
Single strand breaks (SSB) and double strand breaks (DSB) are both main elements of SDF [90]. SSB are mainly a consequence of ROS-induced 8OHdG production, secondary to lipid peroxidation, mitochondrial and nuclear DNA base modifications [95,96]. 8OHdG destabilises DNA structure and produces DNA breaks [97]. SSB can present extensively at multiple points/regions of the genome and have been mainly associated with lower rates of clinical pregnancy or an increase in conception time [98]. Conversely, DSB can be a consequence of SSB during the replication process, ROS, and exogenous elements such as ionising radiation [99,100]. DSB have been associated with an increased risk of miscarriage, decreased embryo quality and lower successful implantation rates following the use of assisted reproduction techniques (ART) such as intracytoplasmic sperm injection (ICSI) [98].
There are several methods available for the evaluation of sperm DNA fragmentation/integrity including: TUNEL assay [101], sperm chromatin structure assay (SCSA) [102], and sperm chromatin dispersion (SCD) [103]. Nevertheless, the above assays are unable to discriminate SSB from DSB [104], unlike the Comet assay (particularly the alkaline and two-tailed methodologies), which is capable of discriminating between SSB and DSB [105,106].
Regarding the effects of DNA damage and sperm DNA fragmentation on human reproduction, SDF has been inversely associated with probability of natural pregnancy in a Danish cohort [107]. Furthermore, SDF becomes particularly relevant in the clinical setting of ART. Despite contradictory evidence [108], possibly owing to differences between SDF detection assays and other confounding factors, a high SDF value is associated with decreased pregnancy rate with IUI and IVF and with increased miscarriage rate following both IVF and ICSI [90]. All of the above is consistent with the correlation between high SDF levels and overall worse reproductive outcomes [24,25,26].
The current management of SDF and OS in unexplained infertility focuses on the use of several interventions, with a variable degree of supporting evidence: antioxidant supplementation [109,110], lifestyle modifications (such as weight loss and smoking cessation) [111,112,113], recurrent ejaculation [114], and ultimately, the use of sperm processing and preparation preceding ICSI [90].

6. Proposal for a Novel Antioxidant Therapeutic Strategy

Taking into account the specific intracellular phenomena and sensitive sperm-development timeframes, we hypothesise that positive achievements will result from interventions based on the relationship between the time course of stages of spermatogenesis and OS. Thus, these processes could be favoured throughout the spermatogenesis cycle by the pertinent effect of appropriate time-related pharmacological agents in order to improve the clinical outcomes of male infertility.
Aiming to properly contextualise the reasoning underlying our proposal, key aspects of sperm development are described below:
Spermatogenesis, the process by which spermatogonial cells become spermatozoa, is a complex process occurring within seminiferous tubules of the testis whose exhaustive description is beyond the aim of this work but can be broadly divided into four phases: (1) mitosis; (2) meiosis; (3) spermiogenesis and (4) spermiation [115]. In humans, a full spermatogenesis cycle lasts approximately 74 days [116], varying from 42 to 76 days [117], with a new cycle initiating every 16 days. At any given moment, 6 spermatogenesis cycles are taking place in each seminiferous tubule, each one in a particular phase. These phases are described below:
(1)
Mitosis: Progenitor Adark- and Apale-spermatogonia undergo continuous mitosis for renewal of the germ cell line, while Apale-spermatogonia commit to differentiation and divide to form B-spermatogonia, subsequently dividing to form preleptotene spermatocytes [118].
(2)
Meiosis: Diploid spermatocytes (2n2c) undergo 2 meiotic divisions to form haploid spermatids (nc). Primary spermatocytes are the first cells to undergo meiosis [119]. During meiotic division, chromosomal recombination and DNA exchange through crossing-over ensure genetic diversity of these cells from their adult precursors.
(3)
Spermiogenesis: Consists of the transformation of round spermatids into elongated, mature spermatozoa. It includes the loss of cytoplasm, migration of cytoplasmic organelles, formation of the acrosome from the Golgi apparatus, formation of the flagellum from the centriole, nuclear compaction to about 10% of former size, and reorganisation of mitochondria around the sperm midpiece [119]. Moreover, the spermatid nucleus undergoes remodelling and condensation, associated with the displacement of histones by transition proteins and then by protamines [120]. Consequently, the vast majority of sperm DNA is coiled into toroids by protamines, with a significantly lower fraction remaining bound to histones, and the DNA is attached to the sperm nuclear matrix at MARs (matrix attachment regions) at medium intervals of roughly 50 kb throughout the genome [121]. It is during this process that germ cells are most sensitive to OS [39]. This could be explained by a number of factors, such as the limited glutathione replenishment and DNA repairing capacities of spermatids [122]. This inherent susceptibility is compensated by the protective antioxidant role of Sertoli cells, mainly exerted through the activity of SOD, as well as reductase, transferase and peroxidase [67].
(4)
Spermiation: The mature spermatozoon is freed of its anchorage to the Sertoli cell, being released into the tubule lumen.
Our proposal includes 3 antioxidants (see Figure 2), whose respective properties are described below:
Omega-3 (ω-3) polyunsaturated fatty acids. PUFAs comprise five substances: eicosapentaenoic acid (EPA), alpha-linolenic-acid, stearidonic acid, docosahexaenoic acid (DHA), and docosapentaenoic acid. Of particular relevance are long-chain ω-3 PUFAs such as EPA and DHA, which are abundant in the body lipids of fatty fish, the liver of white lean fish, and the blubber of marine mammals [123].
EPA and DHA are highly susceptible to free-radical oxidation [124], yet they have been shown to reduce urinary F2-isoprostane levels (a lipid peroxidation biomarker) and enhance the antioxidant systems of the cell [125]. Gao and colleagues published an in vitro study which showed that oxidised EPA and DHA react directly with Keap1, a negative regulator of Nrf2, inducing Nrf2-directed gene expression [125], thus acting as indirect antioxidants.
The consumption of long chain ω-3 PUFAs has shown improvements in sperm SOD and CAT activities [126], sperm cell concentration [126], sperm motility [127] and DNA fragmentation [128]. These compounds have been safely used in doses of up to 1.8 g/day in a RCT [126].
Resveratrol. Resveratrol is a phenolic compound derived from stilbene naturally found in foods such as grapes, nuts, cranberries and red wine [129]. It possesses several beneficial effects on human health that are derived from its antioxidant and anti-inflammatory properties [130].
The antioxidant properties that resveratrol exhibits are partly due to being an excellent scavenger of ·OH, O2 and metal-induced ROS, thereby protecting cell membranes from lipid peroxidation and DNA from ROS-mediated damage [131].
Moreover, this compound has been shown to act through the transcription factor named nuclear factor erythroid 2–related factor 2 (Nrf2) pathway, as it downregulates Keap1 [132], a protein involved in the ubiquitination and proteasomal degradation of Nrf2 through its binding and secondary retention in the cytoplasm [133]. Consequently, it increases the expression and translocation to the nucleus of Nrf2, which binds to antioxidant response elements (AREs) in the promoter regions of genes encoding cytoprotective proteins and antioxidant enzymes such as catalase, superoxide dismutase, glutathione peroxidase, NADPH quinone oxidoreductase and glutathione-S-transferase [134]. This mechanism has been identified in hepatocytes [135], lung epithelial cells [136] and endothelial cells [137].
Furthermore, the upregulation of the Nrf2 pathway by resveratrol could be important in the maintenance of mitochondrial homeostasis and structural integrity [138]. The above becomes particularly relevant given the key role that mitochondrial ROS production appears to have in idiopathic male infertility.
Interestingly, resveratrol was also found to have a significant inhibitory effect on the NF-κB signalling pathway after cellular exposure to metal-induced radicals [131].
In 2020, Illiano et al. treated 20 patients with idiopathic infertility with a resveratrol-based multivitamin supplement containing a dose of 150 mg of resveratrol daily, for 6 months, after which they showed improved sperm motility and concentration compared to baseline measurements [139]. Limitations to the study include small sample size, lack of control group and the absence of measurements related to OS or antioxidant activity. Nevertheless, it sets a precedent regarding a beneficial effect of the compound in this particular clinical setting.
Melatonin. Melatonin is an ancient molecule with presence throughout Eukarya and Bacteria domains, and first appeared as a way to mitigate ROS production secondary to aerobic metabolism [140]. Since its discovery in plants in 1995, it has been identified in a variety of foods and medicinal plants [140,141].
It exhibits antioxidant activity through various direct and indirect mechanisms: ROS scavenging [142,143,144], transition metal binding [145,146], stimulating the activity of antioxidant enzymes such as SOD (particularly isoforms SOD1 and SOD2), catalase and glutathione peroxidase [147], which could be explained by its role in promoting the expression of Nrf2 [148]. Interestingly, its metabolites can also exert similar functions [149,150], suggesting that the antioxidant properties of melatonin continue beyond its metabolism [151].
Melatonin is an amphiphilic molecule, with the capacity to cross biological membranes and reach any cellular and subcellular compartment, owing to this property its high rate of distribution and the ability to exert its antioxidant activity [142,152]. The above translates into a protective effect against the attack of ROS over lipids [153,154], proteins [155,156] and DNA [157,158]. Furthermore, its transport could involve more than passive diffusion, as transport systems located in mitochondria, such as facilitative glucose transporters GLUT/SLC2A and proton-driven oligopeptide transporter PEPT1/2, have been shown to have an active role in facilitating melatonin transport across membranes [159].
Moreover, studies in mice showed melatonin’s ability to reverse the effects of cyanide, an ETC Complex IV inhibitor [160]. Furthermore, it showed a protective role against the effect of several neurotoxins over ETC Complex I. These actions could not be replicated through the use of vitamin C or E [161].
Furthermore, in recent work Malmir and colleagues studied the in vitro effect of melatonin over spermatozoa collected from men with idiopathic asthenoteratozoospermia. DNA fragmentation measured by TUNEL, sperm chromatin dispersion and malondialdehyde levels were significantly reduced while mitochondrial membrane potential and viability were significantly increased in the melatonin group compared to control [162].
Given its stability during digestion and fast absorption, melatonin possesses a high bioavailability [151]. Oral administration has been reported to be safe in humans at doses up to 1000 mg daily for one month [163,164,165].
These compounds share a few characteristics that might be relevant to the effectiveness of the regimen: they exert their antioxidant capacities through various mechanisms, many of which overlap. In addition, the three of them have been shown to promote the activity of intrinsic, enzymatic antioxidant systems through the Nrf2 pathway. In the case of PUFAs, the aim of their use during the mitotic phase of spermatogenesis is to generate non-hypoxic preconditioning of male germ cells in order to strengthen their enzymatic antioxidant defences. Moreover, we propose the use of resveratrol during meiosis because of its simultaneous role as direct and indirect antioxidant, aiming to neutralise the possible exogenous ROS before the most vulnerable moment of spermatogenesis: spermiogenesis. Furthermore, melatonin is the antioxidant proposed for the spermiogenesis phase due to the following features: (i) multiple antioxidant capabilities and mitochondrial stabilising properties; (ii) ample distribution throughout the cell given its amphiphilic nature and the proposed active transport at a mitochondrial level; and (iii) prolonged effect due to the analogous actions of its active metabolites. This proposal (graphically presented in Figure 3) intends to take into account the temporality of the spermatogenesis process, albeit in vivo several cycles of spermatogenesis are taking place simultaneously, as previously described.

7. Concluding Remarks and Future Perspectives on Combined Antioxidant Therapy against Male Infertility

As we have established, several reports support the association of high seminal ROS levels both with impaired sperm parameters [166,167] and seminal DNA fragmentation [168,169,170].
Therefore, the use of antioxidants as a treatment for male infertility presents a reasonable and biologically plausible intervention. Despite how promising it seems, few studies have been able to prove a consistent benefit of antioxidant-based therapy by themselves and evidence is largely controversial (see Appendix A, Table A1). Furthermore, available evidence is often low-quality: a 2019 Cochrane review of 61 studies showed that antioxidants may improve live birth and clinical pregnancy rates, although with concerns regarding quality evidence and high risk of bias. Moreover, antioxidant versus antioxidant comparisons were not performed, as insufficient studies compared the same interventions [171]. Indeed, there is no clear consensus regarding the optimal components and structure of an antioxidant treatment regimen [1].
It is worth taking into account that for most RCT and other trials using antioxidant treatment for male infertility, there is considerable variability regarding sample selection and general inclusion criteria. Despite the amount of evidence supporting OS as a core element in male infertility pathophysiology, idiopathic male infertility cohorts are highly heterogeneous in terms of the specific abnormalities in spermiogram and other clinical evidence that leads to their inclusion in this group. In other words, the concept of idiopathic male infertility becomes an umbrella term for a myriad of clinical conditions that lead to dysfunction of male reproductive function.
A necessary approach to improve the outcomes in future therapeutic interventions could be to effectively establish OS is in fact a main factor in the pathophysiology of their particular reproductive dysfunction by adding high OS biomarker levels as an inclusion criteria for the use of an antioxidant regimen on idiopathic male infertility. Regarding measurements and outcomes, the determination of said biomarkers during and after the intervention becomes mandatory along with sperm quality and in vivo fertility follow-ups in order to establish consistent correlations.
Concerning the components and structure of the antioxidant regimen itself, there is still need for further evidence on the role that commonly cited antioxidants have under in vivo conditions in human sperm. Nevertheless, given the fact that male fertility is a continuous, highly complex process with an equally nuanced pathophysiology, future antioxidant regimens should entail a dynamic, longitudinal approach to antioxidant compound selection and timing of interventions.

Author Contributions

C.Á.: conceptualization, investigation, writing—original draft preparation; J.I.V.: conceptualization, writing—review and editing, supervision; M.A.: conceptualization; L.S.: conceptualization; R.R.: conceptualization, writing—review and editing, supervision. All authors have read and agreed to the published version of the manuscript.

Funding

Not applicable.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Table A1. Studies regarding the use of antioxidants for male infertility. RCT: randomized controlled trial; TAC: total antioxidant capacity.
Table A1. Studies regarding the use of antioxidants for male infertility. RCT: randomized controlled trial; TAC: total antioxidant capacity.
AuthorsJournalPopulationNAntioxidantFollow-Up (Months)OutcomesStudy Design
Steiner et al., 2020 [172]Fertility and SterilityOligozoospermia and/or asthenozoospermia and/or teratozoospermia or DNA fragmentation >25%174500 mg vitamin C 400 mg vitamin E
0.20 mg selenium
1000 mg l-carnitine
20 mg zinc
1000 μg folic acid
10 mg of lycopene, daily for 3–6 months
6 Improvement in sperm concentration
No improvement in sperm morphology, motility or DNA fragmentation. No improvement over in vivo pregnancy or live-birth rates
RCT
Jannatifar et al., 2019 [173]Reproductive Biology and EndocrinologyOligoasthenoteratozoospermia50600 mg NAC daily for 3 months3Improvement in sperm count, morphology, motility, DNA fragmentation and protamine deficiency Clinical trial (no control group)
Alahmar & Sengupta, 2020 [174]Biological Trace Element ResearchOligoasthenoteratospermia70200 mg CoQ10 or 200 μg selenium daily for 3 months3Improvement in sperm count, progressive sperm motility, total sperm motility, TAC and SOD in both groupsRCT
Stenqvist et al., 2018 [175]AndrologyMen from infertile couples, with normal testosterone, LH and FSH levels and DNA fragmentation ≥25%7730 mg vitamin C
5 mg vitamin E
0.5 μg vitamin B12
750 mg L-carnitine
10 mg CoQ10
100 μg folic acid
5 mg zinc
25 μg selenium
6No improvement on DNA fragmentation or standard semen parametersRCT
Kessoupoulou et al., 1995 [176]Fertility and SterilityMen with increased ROS levels measured through chemiluminescent method30600 mg vitamin E daily for 6 months3No improvement in semen parameters, improvement in in vitro zona pellucida bindingRCT
Hawkes et al., 2009 [177]Journal of AndrologyNormozoospermia42300 μg selenium daily for 12 months24No improvement in spermiogram parameters. Se status of testes appears unresponsive to dietary Se intakeRCT
Moilanen et al., 1993 [178]International Journal of AndrologyAsthenozoospermia or oligoasthenozoospermia15300 mg vitamin E daily for 3 months3No improvement in sperm parameters or pregnancy ratesRCT
Greco et al., 2005 [110]Journal of AndrologyUnexplained infertility and ≥15% DNA fragmentation641 g vitamin C
1 g vitamin E daily for 2 months
2No improvement in spermiogram parametersRCT
Rolf et al., 1999 [179]Human ReproductionAsthenozoospermia with/without oligozoospermia311 g vitamin C
800 mg vitamin E daily for 2 months
2No improvement in spermiogram parameters, no pregnancyRCT
Sigman et al., 2006 [180]Fertility and SterilityAsthenozoospermia212 g L-carnitine
1 g L-acetyl-carnitine daily for 4 months
6No improvement in spermiogram parametersRCT

References

  1. Minhas, S.; Bettocchi, C.; Boeri, L.; Capogrosso, P.; Carvalho, J.; Cilesiz, N.C.; Cocci, A.; Corona, G.; Dimitropoulos, K.; Gül, M.; et al. European Association of Urology Guidelines on Male Sexual and Reproductive Health: 2021 Update on Male Infertility. Eur. Urol. 2021, 80, 603–620. [Google Scholar] [CrossRef] [PubMed]
  2. Boivin, J.; Bunting, L.; Collins, J.A.; Nygren, K.G. International estimates of infertility prevalence and treatment-seeking: Potential need and demand for infertility medical care. Hum. Reprod. 2007, 22, 1506–1512. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Thonneau, P.; Marchand, S.; Tallec, A.; Ferial, M.-L.; Ducot, B.; Lansac, J.; Lopes, P.; Tabaste, J.-M.; Spira, A. Incidence and main causes of infertility in a resident population (1,850,000) of three French regions (1988–1989) *. Hum. Reprod. 1991, 6, 811–816. [Google Scholar] [CrossRef] [PubMed]
  4. GBD 2017 Population and Fertility Collaborators. Population and fertility by age and sex for 195 countries and territories, 1950–2017: A systematic analysis for the Global Burden of Disease Study. Lancet 2018, 392, 1995–2051. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Levine, H.; Jørgensen, N.; Martino-Andrade, A.; Mendiola, J.; Weksler-Derri, D.; Mindlis, I.; Pinotti, R.; Swan, S.H. Temporal trends in sperm count: A systematic review and meta-regression analysis. Hum. Reprod. Update 2017, 23, 646–659. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Jungwirth, A.; Giwercman, A.; Tournaye, H.; Diemer, T.; Kopa, Z.; Dohle, G.; Krausz, C.; EAU Working Group on Male Infertility. European Association of Urology Guidelines on Male Infertility: The 2012 Update. Eur. Urol. 2012, 62, 324–332. [Google Scholar] [CrossRef]
  7. Barratt, C.L.R.; Björndahl, L.; De Jonge, C.J.; Lamb, D.J.; Osorio Martini, F.; McLachlan, R.; Oates, R.D.; Van Der Poel, S.; St John, B.; Sigman, M.; et al. The diagnosis of male infertility: An analysis of the evidence to support the development of global WHO guidance—Challenges and future research opportunities. Hum. Reprod. Update 2017, 23, 660–680. [Google Scholar] [CrossRef] [Green Version]
  8. Ventimiglia, E.; Capogrosso, P.; Boeri, L.; Ippolito, S.; Scano, R.; Moschini, M.; Gandaglia, G.; Papaleo, E.; Montorsi, F.; Salonia, A. Validation of the American Society for Reproductive Medicine guidelines/recommendations in white European men presenting for couple’s infertility. Fertil. Steril. 2016, 106, 1076–1082.e1. [Google Scholar] [CrossRef] [Green Version]
  9. Punab, M.; Poolamets, O.; Paju, P.; Vihljajev, V.; Pomm, K.; Ladva, R.; Korrovits, P.; Laan, M. Causes of male infertility: A 9-year prospective monocentre study on 1737 patients with reduced total sperm counts. Hum. Reprod. 2017, 32, 18–31. [Google Scholar] [CrossRef] [Green Version]
  10. Quaas, A.; Dokras, A. Diagnosis and treatment of unexplained infertility. Rev. Obstet. Gynecol. 2008, 1, 69–76. [Google Scholar]
  11. Winters, B.R.; Walsh, T.J. The Epidemiology of Male Infertility. Urol. Clin. N. Am. 2014, 41, 195–204. [Google Scholar] [CrossRef] [PubMed]
  12. Schieber, M.; Chandel, N.S. ROS Function in Redox Signaling and Oxidative Stress. Curr. Biol. 2014, 24, R453–R462. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Valko, M.; Leibfritz, D.; Moncol, J.; Cronin, M.T.D.; Mazur, M.; Telser, J. Free radicals and antioxidants in normal physiological functions and human disease. Int. J. Biochem. Cell Biol. 2007, 39, 44–84. [Google Scholar] [CrossRef] [PubMed]
  14. Kattoor, A.J.; Pothineni, N.V.K.; Palagiri, D.; Mehta, J.L. Oxidative Stress in Atherosclerosis. Curr. Atheroscler. Rep. 2017, 19, 42. [Google Scholar] [CrossRef] [PubMed]
  15. Baradaran, A.; Nasri, H.; Rafieian-Kopaei, M. Oxidative stress and hypertension: Possibility of hypertension therapy with antioxidants. J. Res. Med. Sci. 2014, 19, 358–367. [Google Scholar] [PubMed]
  16. Samman Tahhan, A.; Sandesara, P.B.; Hayek, S.S.; Alkhoder, A.; Chivukula, K.; Hammadah, M.; Mohamed-Kelli, H.; O’Neal, W.T.; Topel, M.; Ghasemzadeh, N.; et al. Association between oxidative stress and atrial fibrillation. Heart Rhythm 2017, 14, 1849–1855. [Google Scholar] [CrossRef]
  17. Tsutsui, H.; Kinugawa, S.; Matsushima, S. Oxidative stress and heart failure. Am. J. Physiol. Heart Circ. Physiol. 2011, 301, H2181–H2190. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Rodrigo, R.; Fernandez-Gajardo, R.; Gutierrez, R.; Matamala, J.M.; Carrasco, R.; Miranda-Merchak, A.; Feuerhake, W. Oxidative Stress and Pathophysiology of Ischemic Stroke: Novel Therapeutic Opportunities. CNS Neurol. Disord. Drug Targets 2013, 12, 698–714. [Google Scholar] [CrossRef]
  19. Vona, R.; Gambardella, L.; Cittadini, C.; Straface, E.; Pietraforte, D. Biomarkers of Oxidative Stress in Metabolic Syndrome and Associated Diseases. Oxidative Med. Cell. Longev. 2019, 2019, 8267234. [Google Scholar] [CrossRef] [Green Version]
  20. Videla, L.A.; Rodrigo, R.; Araya, J.; Poniachik, J. Oxidative stress and depletion of hepatic long-chain polyunsaturated fatty acids may contribute to nonalcoholic fatty liver disease. Free Radic. Biol. Med. 2004, 37, 1499–1507. [Google Scholar] [CrossRef]
  21. Chen, Z.; Zhong, C. Oxidative stress in Alzheimer’s disease. Neurosci. Bull. 2014, 30, 271–281. [Google Scholar] [CrossRef] [Green Version]
  22. Radi, E.; Formichi, P.; Battisti, C.; Federico, A. Apoptosis and Oxidative Stress in Neurodegenerative Diseases. J. Alzheimer’s Dis. 2014, 42 (Suppl. S3), S125–S152. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Aitken, R. A free radical theory of male infertility. Reprod. Fertil. Dev. 1994, 6, 19–23; discussion 23–24. [Google Scholar] [CrossRef] [PubMed]
  24. Faduola, P.; Kolade, C.O. Sperm chromatin structure assay results in Nigerian men with unexplained infertility. Clin. Exp. Reprod. Med. 2015, 42, 101–105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Santi, D.; Spaggiari, G.; Simoni, M. Sperm DNA fragmentation index as a promising predictive tool for male infertility diagnosis and treatment management—Meta-analyses. Reprod. Biomed. Online 2018, 37, 315–326. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Mayorga-Torres, B.J.M.; Camargo, M.; Cadavid, P.; du Plessis, S.S.; Maya, W.D.C. Are oxidative stress markers associated with unexplained male infertility? Andrologia 2017, 49, e12659. [Google Scholar] [CrossRef]
  27. Agarwal, A.; Allamaneni, S.S.R.; Nallella, K.P.; George, A.T.; Mascha, E. Correlation of reactive oxygen species levels with the fertilization rate after in vitro fertilization: A qualified meta-analysis. Fertil. Steril. 2005, 84, 228–231. [Google Scholar] [CrossRef]
  28. Saleh, R.A.; Agarwal, A.; Nada, E.A.; El-Tonsy, M.H.; Sharma, R.K.; Meyer, A.; Nelson, D.R.; Thomas, A.J. Negative effects of increased sperm DNA damage in relation to seminal oxidative stress in men with idiopathic and male factor infertility. Fertil. Steril. 2003, 79 (Suppl. S3), 1597–1605. [Google Scholar] [CrossRef]
  29. Oleszczuk, K.; Augustinsson, L.; Bayat, N.; Giwercman, A.; Bungum, M. Prevalence of high DNA fragmentation index in male partners of unexplained infertile couples. Andrology 2012, 1, 357–360. [Google Scholar] [CrossRef]
  30. Vandekerckhove, F.W.R.C.; De Croo, I.; Gerris, J.; Abbeel, E.V.; De Sutter, P. Sperm Chromatin Dispersion Test before Sperm Preparation Is Predictive of Clinical Pregnancy in Cases of Unexplained Infertility Treated with Intrauterine Insemination and Induction with Clomiphene Citrate. Front. Med. 2016, 3, 63. [Google Scholar] [CrossRef] [Green Version]
  31. Zandieh, Z.; Vatannejad, A.; Doosti, M.; Zabihzadeh, S.; Haddadi, M.; Bajelan, L.; Rashidi, B.; Amanpour, S. Comparing reactive oxygen species and DNA fragmentation in semen samples of unexplained infertile and healthy fertile men. Ir. J. Med. Sci. 2017, 187, 657–662. [Google Scholar] [CrossRef] [PubMed]
  32. Ritchie, C.; Ko, E.Y. Oxidative stress in the pathophysiology of male infertility. Andrologia 2021, 53, e13581. [Google Scholar] [CrossRef] [PubMed]
  33. Baskaran, S.; Finelli, R.; Agarwal, A.; Henkel, R. Reactive oxygen species in male reproduction: A boon or a bane? Andrologia 2021, 53, e13577. [Google Scholar] [CrossRef]
  34. Fukai, T.; Ushio-Fukai, M. Superoxide Dismutases: Role in Redox Signaling, Vascular Function, and Diseases. Antioxid. Redox Signal. 2011, 15, 1583–1606. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Murphy, M.P. How mitochondria produce reactive oxygen species. Biochem. J. 2009, 417, 1–13. [Google Scholar] [CrossRef] [Green Version]
  36. Azadi, L.; Tavalaee, M.; Deemeh, M.R.; Arbabian, M.; Nasr-Esfahani, M.H. Effects of Tempol and Quercetin on Human Sperm Function after Cryopreservation. Cryoletters 2017, 38, 29–36. [Google Scholar]
  37. Herman, S.; Lipiński, P.; Ogórek, M.; Starzyński, R.; Grzmil, P.; Bednarz, A.; Lenartowicz, M. Molecular Regulation of Copper Homeostasis in the Male Gonad during the Process of Spermatogenesis. Int. J. Mol. Sci. 2020, 21, 9053. [Google Scholar] [CrossRef]
  38. Bánfi, B.; Molnár, G.; Maturana, A.; Steger, K.; Hegedûs, B.; Demaurex, N.; Krause, K.-H. A Ca(2+)-activated NADPH oxidase in testis, spleen, and lymph nodes. J. Biol. Chem. 2001, 276, 37594–37601. [Google Scholar] [CrossRef] [Green Version]
  39. Aitken, R.J.; Curry, B.J. Redox Regulation of Human Sperm Function: From the Physiological Control of Sperm Capacitation to the Etiology of Infertility and DNA Damage in the Germ Line. Antioxid. Redox Signal. 2011, 14, 367–381. [Google Scholar] [CrossRef]
  40. Richer, S.C.; Ford, W.C. A critical investigation of NADPH oxidase activity in human spermatozoa. Mol. Hum. Reprod. 2001, 7, 237–244. [Google Scholar] [CrossRef]
  41. Armstrong, J.S.; Bivalacqua, T.J.; Chamulitrat, W.; Sikka, S.; Hellstrom, W.J.G. A comparison of the NADPH oxidase in human sperm and white blood cells. Int. J. Androl. 2002, 25, 223–229. [Google Scholar] [CrossRef]
  42. Aitken, R.J.; West, K.M. Analysis of the relationship between reactive oxygen species production and leucocyte infiltration in fractions of human semen separated on Percoll gradients. Int. J. Androl. 1990, 13, 433–451. [Google Scholar] [CrossRef]
  43. Liu, Y.; Ding, Z. Obesity, a serious etiologic factor for male subfertility in modern society. Reproduction 2017, 154, R123–R131. [Google Scholar] [CrossRef] [Green Version]
  44. Heydari, H.; Ghiasi, R.; Ghaderpour, S.; Keyhanmanesh, R. The Mechanisms Involved in Obesity-Induced Male Infertility. Curr. Diabetes Rev. 2021, 17, 259–267. [Google Scholar] [CrossRef]
  45. Leisegang, K.; Sengupta, P.; Agarwal, A.; Henkel, R. Obesity and male infertility: Mechanisms and management. Andrologia 2020, 53, e13617. [Google Scholar] [CrossRef]
  46. Harlev, A.; Agarwal, A.; Gunes, S.O.; Shetty, A.; du Plessis, S.S. Smoking and Male Infertility: An Evidence-Based Review. World J. Men’s Health 2015, 33, 143–160. [Google Scholar] [CrossRef] [Green Version]
  47. Adewoyin, M.; Ibrahim, M.; Roszaman, R.; Isa, M.L.M.; Alewi, N.A.M.; Rafa, A.A.A.; Anuar, M.N.N. Male Infertility: The Effect of Natural Antioxidants and Phytocompounds on Seminal Oxidative Stress. Diseases 2017, 5, 9. [Google Scholar] [CrossRef] [Green Version]
  48. Sharma, R.; Harlev, A.; Agarwal, A.; Esteves, S.C. Cigarette Smoking and Semen Quality: A New Meta-analysis Examining the Effect of the 2010 World Health Organization Laboratory Methods for the Examination of Human Semen. Eur. Urol. 2016, 70, 635–645. [Google Scholar] [CrossRef]
  49. Taha, E.A.; Ez-Aldin, A.M.; Sayed, S.K.; Ghandour, N.M.; Mostafa, T. Effect of Smoking on Sperm Vitality, DNA Integrity, Seminal Oxidative Stress, Zinc in Fertile Men. Urology 2012, 80, 822–825. [Google Scholar] [CrossRef]
  50. Ricci, E.; Al Beitawi, S.; Cipriani, S.; Candiani, M.; Chiaffarino, F.; Viganò, P.; Noli, S.; Parazzini, F. Semen quality and alcohol intake: A systematic review and meta-analysis. Reprod. Biomed. Online 2016, 34, 38–47. [Google Scholar] [CrossRef] [Green Version]
  51. Kesari, K.K.; Agarwal, A.; Henkel, R. Radiations and male fertility. Reprod. Biol. Endocrinol. 2018, 16, 118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Carré, J.; Gatimel, N.; Moreau, J.; Parinaud, J.; Léandri, R. Does air pollution play a role in infertility? A systematic review. Environ. Health 2017, 16, 82. [Google Scholar] [CrossRef] [PubMed]
  53. Griveau, J.E.; Renard, P.; Le Lannou, D. An in vitro promoting role for hydrogen peroxide in human sperm capacitation. Int. J. Androl. 1994, 17, 300–307. [Google Scholar] [CrossRef] [PubMed]
  54. Aitken, R.J. Reactive oxygen species as mediators of sperm capacitation and pathological damage. Mol. Reprod. Dev. 2017, 84, 1039–1052. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. de Lamirande, E.; Gagnon, C. A positive role for the superoxide anion in triggering hyperactivation and capacitation of human spermatozoa. Int. J. Androl. 1993, 16, 21–25. [Google Scholar] [CrossRef] [PubMed]
  56. de Lamirande, E.; Gagnon, C. Human sperm hyperactivation in whole semen and its association with low superoxide scavenging capacity in seminal plasma. Fertil. Steril. 1993, 59, 1291–1295. [Google Scholar] [CrossRef]
  57. de Lamirande, E.; Tsai, C.; Harakat, A.; Gagnon, C. Involvement of reactive oxygen species in human sperm arcosome reaction induced by A23187, lysophosphatidylcholine, and biological fluid ultrafiltrates. J. Androl. 1998, 19, 585–594. [Google Scholar]
  58. Du Plessis, S.S.; Agarwal, A.; Halabi, J.; Tvrda, E. Contemporary evidence on the physiological role of reactive oxygen species in human sperm function. J. Assist. Reprod. Genet. 2015, 32, 509–520. [Google Scholar] [CrossRef] [Green Version]
  59. Flesch, F.M.; Gadella, B.M. Dynamics of the mammalian sperm plasma membrane in the process of fertilization. Biochim. Biophys. Acta BBA Rev. Biomembr. 2000, 1469, 197–235. [Google Scholar] [CrossRef]
  60. Goldman, R.; Ferber, E.; Zort, U. Reactive oxygen species are involved in the activation of cellular phospholipase. AFEBS Lett. 1992, 309, 190–192. [Google Scholar] [CrossRef] [Green Version]
  61. Peeker, R.; Abramsson, L.; Marklund, S.L. Superoxide dismutase isoenzymes in human seminal plasma and spermatozoa. Mol. Hum. Reprod. 1997, 3, 1061–1066. [Google Scholar] [CrossRef] [PubMed]
  62. Lee, S.; Van Remmen, H.; Csete, M. Sod2 overexpression preserves myoblast mitochondrial mass and function, but not muscle mass with aging. Aging Cell 2009, 8, 296–310. [Google Scholar] [CrossRef] [PubMed]
  63. Hu, D.; Cao, P.; Thiels, E.; Chu, C.T.; Wu, G.-Y.; Oury, T.D.; Klann, E. Hippocampal long-term potentiation, memory, and longevity in mice that overexpress mitochondrial superoxide dismutase. Neurobiol. Learn. Mem. 2007, 87, 372–384. [Google Scholar] [CrossRef]
  64. Silva, J.P.; Shabalina, I.G.; Dufour, E.; Petrovic, N.; Backlund, E.C.; Hultenby, K.; Wibom, R.; Nedergaard, J.; Cannon, B.; Larsson, N.-G. SOD2 overexpression: Enhanced mitochondrial tolerance but absence of effect on UCP activity. EMBO J. 2005, 24, 4061–4070. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Foresta, C.; Flohé, L.; Garolla, A.; Roveri, A.; Ursini, F.; Maiorino, M. Male fertility is linked to the selenoprotein phospholipid hydroperoxide glutathione peroxidase. Biol. Reprod. 2002, 67, 967–971. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Meseguer, M.; Martínez-Conejero, J.A.; Muriel, L.; Pellicer, A.; Remohí, J.; Garrido, N. The Human Sperm Glutathione System: A Key Role in Male Fertility and Successful Cryopreservation. Drug Metab. Lett. 2007, 1, 121–126. [Google Scholar] [CrossRef] [PubMed]
  67. Bauché, F.; Fouchard, M.-H.; Jegou, B. Antioxidant system in rat testicular cells. FEBS Lett. 1994, 349, 392–396. [Google Scholar] [CrossRef] [Green Version]
  68. Jeulin, C.; Soufir, J.C.; Weber, P.; Laval-Martin, D.; Calvayrac, R. Catalase activity in human spermatozoa and seminal plasma. Gamete Res. 1989, 24, 185–196. [Google Scholar] [CrossRef]
  69. Moretti, E.; Castellini, C.; Mourvaki, E.; Capitani, S.; Geminiani, M.; Renieri, T.; Collodel, C. Distribution of α- and δ-tocopherols in seminal plasma and sperm fractions of men with normal and abnormal semen parameters. J. Androl. 2011, 32, 232–239. [Google Scholar] [CrossRef]
  70. Panner Selvam, M.K.; Agarwal, A.; Henkel, R.; Finelli, R.; Robert, K.A.; Iovine, C.; Baskaran, S. The effect of oxidative and reductive stress on semen parameters and functions of physiologically normal human spermatozoa. Free Radic. Biol. Med. 2020, 152, 375–385. [Google Scholar] [CrossRef]
  71. Castagné, V.; Lefevre, K.; Natero, R.; Becker, D.A.; Clarke, P.G.H. An optimal redox status for the survival of axotomized ganglion cells in the developing retina. Neuroscience 1999, 93, 313–320. [Google Scholar] [CrossRef] [PubMed]
  72. Ménézo, Y.; Entezami, F.; Lichtblau, I.; Belloc, S.; Cohen, M.; Dale, B. Oxidative stress and fertility: Incorrect assumptions and ineffective solutions? Zygote 2014, 22, 80–90. [Google Scholar] [CrossRef] [Green Version]
  73. Ménézo, Y., Jr.; Hazout, A.; Panteix, G.; Robert, F.; Rollet, J.; Cohen-Bacrie, P.; Chapuis, F.; Clément, P.; Benkhalifa, M. Antioxidants to reduce sperm DNA fragmentation: An unexpected adverse effect. Reprod. Biomed. Online 2007, 14, 418–421. [Google Scholar] [CrossRef] [PubMed]
  74. Henkel, R.; Sandhu, I.S.; Agarwal, A. The excessive use of antioxidant therapy: A possible cause of male infertility? Andrologia 2019, 51, e13162. [Google Scholar] [CrossRef] [Green Version]
  75. Aitken, R.J.; Jones, K.; Robertson, S. Reactive Oxygen Species and Sperm Function—In Sickness and In Health. J. Androl. 2012, 33, 1096–1106. [Google Scholar] [CrossRef] [PubMed]
  76. 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] [Green Version]
  77. 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] [PubMed]
  78. Hosseinzadeh Colagar, A.; Karimi, F.; Jorsaraei, S.G.A. Correlation of sperm parameters with semen lipid peroxidation and total antioxidants levels in astheno- and oligoasheno- teratospermic men. Iran. Red Crescent Med. J. 2013, 15, 780–785. [Google Scholar] [CrossRef] [Green Version]
  79. Koppers, A.J.; Mitchell, L.A.; Wang, P.; Lin, M.; Aitken, R.J. Phosphoinositide 3-kinase signalling pathway involvement in a truncated apoptotic cascade associated with motility loss and oxidative DNA damage in human spermatozoa. Biochem. J. 2011, 436, 687–698. [Google Scholar] [CrossRef]
  80. Lozano, G.M.; Bejarano, I.; Espino, J.; González, D.; Ortiz, Á.; García, J.F.; Rodríguez, A.B.; Pariente, J.A. Relationship between Caspase Activity and Apoptotic Markers in Human Sperm in Response to Hydrogen Peroxide and Progesterone. J. Reprod. Dev. 2009, 55, 615–621. [Google Scholar] [CrossRef] [Green Version]
  81. Ranawat, P.; Bansal, M.P. Apoptosis induced by modulation in selenium status involves p38 MAPK and ROS: Implications in spermatogenesis. Mol. Cell. Biochem. 2009, 330, 83–95. [Google Scholar] [CrossRef] [PubMed]
  82. Mitchell, L.A.; De Iuliis, G.; Aitken, R.J. The TUNEL assay consistently underestimates DNA damage in human spermatozoa and is influenced by DNA compaction and cell vitality: Development of an improved methodology. Int. J. Androl. 2011, 34, 2–13. [Google Scholar] [CrossRef] [PubMed]
  83. Sotolongo, B.; Huang, T.T.F.; Isenberger, E.; Ward, W.S. An Endogenous Nuclease in Hamster, Mouse, and Human Spermatozoa Cleaves DNA into Loop-Sized Fragments. J. Androl. 2005, 26, 272–280. [Google Scholar] [CrossRef] [PubMed]
  84. Sakkas, D.; Seli, E.; Bizzaro, D.; Tarozzi, N.; Manicardi, G.C. Abnormal spermatozoa in the ejaculate: Abortive apoptosis and faulty nuclear remodelling during spermatogenesis. Reprod. Biomed. Online 2003, 7, 428–432. [Google Scholar] [CrossRef] [PubMed]
  85. Henkel, R.; Samanta, L.; Agarwal, A. (Eds.) Oxidants, Antioxidants and Impact of the Oxidative Status in Male Reproduction; Academic Press: Cambridge, MA, USA, 2018. [Google Scholar] [CrossRef]
  86. Kao, S.H.; Chao, H.T.; Wei, Y.-H. Multiple deletions of mitochondrial DNA are associated with the decline of motility and fertility of human spermatozoa. Mol. Hum. Reprod. 1998, 4, 657–666. [Google Scholar] [CrossRef]
  87. Santiso, R.; Tamayo, M.; Gosálvez, J.; Meseguer, M.; Garrido, N.; Fernández, J.L. Simultaneous determination in situ of DNA fragmentation and 8-oxoguanine in human sperm. Fertil. Steril. 2010, 93, 314–318. [Google Scholar] [CrossRef]
  88. De Iuliis, G.N.; Thomson, L.K.; Mitchell, L.A.; Finnie, J.M.; Koppers, A.J.; Hedges, A.; Nixon, B.; Aitken, R.J. DNA damage in human spermatozoa is highly correlated with the efficiency of chromatin remodeling and the formation of 8-hydroxy-2′-deoxyguanosine, a marker of oxidative stress. Biol. Reprod. 2009, 81, 517–524. [Google Scholar] [CrossRef] [Green Version]
  89. Koppers, A.J.; De Iuliis, G.N.; Finnie, J.M.; McLaughlin, E.A.; Aitken, R.J. Significance of Mitochondrial Reactive Oxygen Species in the Generation of Oxidative Stress in Spermatozoa. J. Clin. Endocrinol. Metab. 2008, 93, 3199–3207. [Google Scholar] [CrossRef] [Green Version]
  90. Agarwal, A.; Majzoub, A.; Baskaran, S.; Panner Selvam, M.K.; Cho, C.L.; Henkel, R.; Finelli, R.; Leisegang, K.; Sengupta, P.; Barbarosie, C.; et al. Sperm DNA Fragmentation: A New Guideline for Clinicians. World J. Men’s Health 2020, 38, 412–471. [Google Scholar] [CrossRef]
  91. Andrés, C.M.C.; de la Lastra, J.M.P.; Juan, C.A.; Plou, F.J.; Pérez-Lebeña, E. Chemistry of Hydrogen Peroxide Formation and Elimination in Mammalian Cells, and Its Role in Various Pathologies. Stresses 2022, 2, 256–274. [Google Scholar] [CrossRef]
  92. Łuczaj, W.; Skrzydlewska, E. DNA damage caused by lipid peroxidation products. Cell Mol. Biol. Lett. 2003, 8, 391–413. [Google Scholar] [PubMed]
  93. Moazamian, R.; Polhemus, A.; Connaughton, H.; Fraser, B.; Whiting, S.; Gharagozloo, P.; Aitken, R.J. Oxidative stress and human spermatozoa: Diagnostic and functional significance of aldehydes generated as a result of lipid peroxidation. Mol. Hum. Reprod. 2015, 21, 502–515. [Google Scholar] [CrossRef]
  94. Badouard, C.; Ménézo, Y.; Panteix, G.; Ravanat, J.; Douki, T.; Cadet, J.; Favier, A. Determination of new types of DNA lesions in human sperm. Zygote 2008, 16, 9–13. [Google Scholar] [CrossRef] [PubMed]
  95. Enciso, M.; Sarasa, J.; Agarwal, A.; Fernández, J.L.; Gosálvez, J. A two-tailed Comet assay for assessing DNA damage in spermatozoa. Reprod. Biomed. Online 2009, 18, 609–616. [Google Scholar] [CrossRef] [PubMed]
  96. Simon, L.; Carrell, D.T. Sperm DNA Damage Measured by Comet Assay. Methods Mol. Biol. 2013, 927, 137–146. [Google Scholar]
  97. Lord, T.; Aitken, R.J. Fertilization stimulates 8-hydroxy-2′-deoxyguanosine repair and antioxidant activity to prevent mutagenesis in the embryo. Dev. Biol. 2015, 406, 1–13. [Google Scholar] [CrossRef] [Green Version]
  98. Ribas-Maynou, J.; Benet, J. Single and Double Strand Sperm DNA Damage: Different Reproductive Effects on Male Fertility. Genes 2019, 10, 105. [Google Scholar] [CrossRef] [Green Version]
  99. Kowalczykowski, S.C. Initiation of genetic recombination and recombination-dependent replication. Trends Biochem. Sci. 2000, 25, 156–165. [Google Scholar] [CrossRef]
  100. Jackson, S.P.; Bartek, J. The DNA-damage response in human biology and disease. Nature 2009, 461, 1071–1078. [Google Scholar] [CrossRef] [Green Version]
  101. Sharma, R.; Iovine, C.; Agarwal, A.; Henkel, R. TUNEL assay—Standardized method for testing sperm DNA fragmentation. Andrologia 2020, 53, e13738. [Google Scholar] [CrossRef]
  102. Evenson, D.P. Sperm Chromatin Structure Assay (SCSA®). Methods Mol. Biol. 2013, 927, 147–164. [Google Scholar]
  103. Fernández, J.L.; Muriel, L.; Rivero, M.T.; Goyanes, V.; Vazquez, R.; Alvarez, J.G. The sperm chromatin dispersion test: A simple method for the determination of sperm DNA fragmentation. J. Androl. 2003, 24, 59–66. [Google Scholar]
  104. Casanovas, A.; Ribas-Maynou, J.; Lara-Cerrillo, S.; Jimenez-Macedo, A.R.; Hortal, O.; Benet, J.; Carrera, J.; García-Peiró, A. Double-stranded sperm DNA damage is a cause of delay in embryo development and can impair implantation rates. Fertil. Steril. 2019, 111, 699–707.e1. [Google Scholar] [CrossRef]
  105. Simon, L.; Aston, K.I.; Emery, B.R.; Hotaling, J.; Carrell, D.T. Sperm DNA damage output parameters measured by the alkaline Comet assay and their importance. Andrologia 2016, 49, e12608. [Google Scholar] [CrossRef]
  106. Cortés-Gutiérrez, E.I.; Fernández, J.L.; Dávila-Rodríguez, M.I.; López-Fernández, C.; Gosálvez, J. Two-Tailed Comet Assay (2T-Comet): Simultaneous Detection of DNA Single and Double Strand Breaks. Methods Mol. Biol. 2017, 1560, 285–293. [Google Scholar] [CrossRef]
  107. Spanò, M.; Bonde, J.P.; Hjøllund, H.I.; Kolstad, H.A.; Cordelli, E.; Leter, G. Sperm chromatin damage impairs human fertility. The Danish First Pregnancy Planner Study Team. Fertil. Steril. 2000, 73, 43–50. [Google Scholar] [CrossRef]
  108. Zini, A.; Libman, J. Sperm DNA damage: Clinical significance in the era of assisted reproduction. Can. Med. Assoc. J. 2006, 175, 495–500. [Google Scholar] [CrossRef] [Green Version]
  109. Tunc, O.; Thompson, J.; Tremellen, K. Improvement in sperm DNA quality using an oral antioxidant therapy. Reprod. Biomed. Online 2009, 18, 761–768. [Google Scholar] [CrossRef]
  110. Greco, E.; Iacobelli, M.; Rienzi, L.; Ubaldi, F.; Ferrero, S.; Tesarik, J. Reduction of the Incidence of Sperm DNA Fragmentation by Oral Antioxidant Treatment. J. Androl. 2005, 26, 349–353. [Google Scholar] [CrossRef]
  111. Samavat, J.; Cantini, G.; Lotti, F.; Di Franco, A.; Tamburrino, L.; Degl’Innocenti, S.; Maseroli, E.; Filimberti, E.; Facchiano, E.; Lucchese, M.; et al. Massive Weight Loss Obtained by Bariatric Surgery Affects Semen Quality in Morbid Male Obesity: A Preliminary Prospective Double-Armed Study. Obes. Surg. 2017, 28, 69–76. [Google Scholar] [CrossRef]
  112. Faure, C.; Dupont, C.; Baraibar, M.A.; Ladouce, R.; Cedrin-Durnerin, I.; Wolf, J.P.; Lévy, R. In Subfertile Couple, Abdominal Fat Loss in Men Is Associated with Improvement of Sperm Quality and Pregnancy: A Case-Series. PLoS ONE 2014, 9, e86300. [Google Scholar] [CrossRef] [PubMed]
  113. Mostafa, R.M.; Nasrallah, Y.S.; Hassan, M.M.; Farrag, A.F.; Majzoub, A.; Agarwal, A. The effect of cigarette smoking on human seminal parameters, sperm chromatin structure and condensation. Andrologia 2017, 50, e12910. [Google Scholar] [CrossRef]
  114. Sánchez-Martín, P.; Sánchez-Martín, F.; González-Martínez, M.; Gosálvez, J. Increased pregnancy after reduced male abstinence. Syst. Biol. Reprod. Med. 2013, 59, 256–260. [Google Scholar] [CrossRef] [PubMed]
  115. Hess, R.A.; Renato de Franca, L. Spermatogenesis and cycle of the seminiferous epithelium. Adv. Exp. Med. Biol. 2008, 636, 1–15. [Google Scholar] [PubMed]
  116. Heller, C.H.; Clermont, Y. Kinetics of the Germinal Epithelium in Man. Recent Prog. Horm. Res. 1964, 20, 545–575. [Google Scholar]
  117. Misell, L.M.; Holochwost, D.; Boban, D.; Santi, N.; Shefi, S.; Hellerstein, M.K.; Turek, P.J. A stable isotope-mass spectrometric method for measuring human spermatogenesis kinetics in vivo. J. Urol. 2006, 175, 242–246; discussion 246. [Google Scholar] [CrossRef]
  118. Amann, R.P. The cycle of the seminiferous epithelium in humans: A need to revisit? J. Androl. 2008, 29, 469–487. [Google Scholar] [CrossRef]
  119. Kerr, J.F. The Cytology of the Human Testis. A.; Burger, H., de Kretser, D.M., Eds.; The Testis; Raven Press: New York, NY, USA, 1981. [Google Scholar]
  120. Steger, K.; Pauls, K.; Klonisch, T.; Franke, F.E.; Bergmann, M. Expression of protamine-1 and -2 mRNA during human spermiogenesis. Mol. Hum. Reprod. 2000, 6, 219–225. [Google Scholar] [CrossRef] [Green Version]
  121. Ward, W.S. Function of sperm chromatin structural elements in fertilization and development. Mol. Hum. Reprod. 2010, 16, 30–36. [Google Scholar] [CrossRef] [Green Version]
  122. Den Boer, P.J.; Poot, M.; Verkerk, A.; Jansen, R.; Mackenbach, P.; Grootegoed, J.A. Glutathione-dependent defence mechanisms in isolated round spermatids from the rat. Int. J. Androl. 1990, 13, 26–38. [Google Scholar] [CrossRef]
  123. Shahidi, F.; Ambigaipalan, P. Omega-3 Polyunsaturated Fatty Acids and Their Health Benefits. Annu. Rev. Food Sci. Technol. 2018, 9, 345–381. [Google Scholar] [CrossRef]
  124. Gao, L.; Yin, H.; Milne, G.; Porter, N.A.; Morrow, J.D. Formation of F-ring Isoprostane-like Compounds (F3-Isoprostanes) in Vivo from Eicosapentaenoic Acid. J. Biol. Chem. 2006, 281, 14092–14099. [Google Scholar] [CrossRef] [Green Version]
  125. Gao, L.; Wang, J.; Sekhar, K.R.; Yin, H.; Yared, N.F.; Schneider, S.N.; Sasi, S.; Dalton, T.P.; Anderson, M.E.; Chan, J.Y.; et al. Novel n-3 Fatty Acid Oxidation Products Activate Nrf2 by Destabilizing the Association between Keap1 and Cullin. J. Biol. Chem. 2007, 282, 2529–2537. [Google Scholar] [CrossRef] [Green Version]
  126. Safarinejad, M.R. Effect of omega-3 polyunsaturated fatty acid supplementation on semen profile and enzymatic anti-oxidant capacity of seminal plasma in infertile men with idiopathic oligoasthenoteratospermia: A double-blind, placebo-controlled, randomised study. Andrologia 2011, 43, 38–47. [Google Scholar] [CrossRef]
  127. Hosseini, B.; Nourmohamadi, M.; Hajipour, S.; Taghizadeh, M.; Asemi, Z.; Keshavarz, S.A.; Jafarnejad, S. The Effect of Omega-3 Fatty Acids, EPA, and/or DHA on Male Infertility: A Systematic Review and Meta-analysis. J. Diet. Suppl. 2019, 16, 245–256. [Google Scholar] [CrossRef]
  128. Martínez-Soto, J.C.; Domingo, J.C.; Cordobilla, B.; Nicolás, M.; Fernández, L.; Albero, P.; Gadea, J.; Landeras, J. Dietary supplementation with docosahexaenoic acid (DHA) improves seminal antioxidant status and decreases sperm DNA fragmentation. Syst. Biol. Reprod. Med. 2016, 62, 387–395. [Google Scholar] [CrossRef]
  129. Hasan, M.; Bae, H. An Overview of Stress-Induced Resveratrol Synthesis in Grapes: Perspectives for Resveratrol-Enriched Grape Products. Molecules 2017, 22, 294. [Google Scholar] [CrossRef] [Green Version]
  130. Kundu, J.K.; Surh, Y.-J. Cancer chemopreventive and therapeutic potential of resveratrol: Mechanistic perspectives. Cancer Lett. 2008, 269, 243–261. [Google Scholar] [CrossRef]
  131. Leonard, S.S.; Xia, C.; Jiang, B.-H.; Stinefelt, B.; Klandorf, H.; Harris, G.K.; Shi, X. Resveratrol scavenges reactive oxygen species and effects radical-induced cellular responses. Biochem. Biophys. Res. Commun. 2003, 309, 1017–1026. [Google Scholar] [CrossRef]
  132. Farkhondeh, T.; Folgado, S.L.; Pourbagher-Shahri, A.M.; Ashrafizadeh, M.; Samarghandian, S. The therapeutic effect of resveratrol: Focusing on the Nrf2 signaling pathway. Biomed. Pharmacother. 2020, 127, 110234. [Google Scholar] [CrossRef]
  133. Villeneuve, N.F.; Lau, A.; Zhang, D.D. Regulation of the Nrf2–Keap1 Antioxidant Response by the Ubiquitin Proteasome System: An Insight into Cullin-Ring Ubiquitin Ligases. Antioxid. Redox Signal. 2010, 13, 1699–1712. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Gureev, A.P.; Popov, V.N.; Starkov, A.A. Crosstalk between the mTOR and Nrf2/ARE signaling pathways as a target in the improvement of long-term potentiation. Exp. Neurol. 2020, 328, 113285. [Google Scholar] [CrossRef] [PubMed]
  135. Rubiolo, J.A.; Mithieux, G.; Vega, F.V. Resveratrol protects primary rat hepatocytes against oxidative stress damage: Activation of the Nrf2 transcription factor and augmented activities of antioxidant enzymes. Eur. J. Pharmacol. 2008, 591, 66–72. [Google Scholar] [CrossRef] [PubMed]
  136. Kode, A.; Rajendrasozhan, S.; Caito, S.; Yang, S.-R.; Megson, I.L.; Rahman, I. Resveratrol induces glutathione synthesis by activation of Nrf2 and protects against cigarette smoke-mediated oxidative stress in human lung epithelial cells. Am. J. Physiol. Cell. Mol. Physiol. 2008, 294, L478–L488. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Song, X.; Liu, Z.; Zeng, R.; Shao, J.; Zheng, Y.; Ye, W. Resveratrol Alleviates Vascular Endothelial Damage Caused by Lower-Extremity Ischemia Reperfusion (I/R) through Regulating Keap1/Nrf2 Signaling-Mediated Oxidative Stress. Evid. Based Complement. Altern. Med. 2021, 2021, 5556603. [Google Scholar] [CrossRef]
  138. Dinkova-Kostova, A.T.; Abramov, A.Y. The emerging role of Nrf2 in mitochondrial function. Free Radic. Biol. Med. 2015, 88, 179–188. [Google Scholar] [CrossRef]
  139. Illiano, E.; Trama, F.; Zucchi, A.; Iannitti, R.G.; Fioretti, B.; Costantini, E. Resveratrol-Based Multivitamin Supplement Increases Sperm Concentration and Motility in Idiopathic Male Infertility: A Pilot Clinical Study. J. Clin. Med. 2020, 9, 4017. [Google Scholar] [CrossRef]
  140. Kurland, C.G.; Andersson, S.G.E. Origin and Evolution of the Mitochondrial Proteome. Microbiol. Mol. Biol. Rev. 2000, 64, 786–820. [Google Scholar] [CrossRef] [Green Version]
  141. Murch, S.J.; Erland, L.A.E. A Systematic Review of Melatonin in Plants: An Example of Evolution of Literature. Front. Plant Sci. 2021, 12, 1016. [Google Scholar] [CrossRef]
  142. Tan, D.-X.; Reiter, R.J.; Manchester, L.C.; Yan, M.-T.; El-Sawi, M.; Sainz, R.M.; Mayo, J.C.; Kohen, R.; Allegra, M.; Hardelan, R. Chemical and Physical Properties and Potential Mechanisms: Melatonin as a Broad Spectrum Antioxidant and Free Radical Scavenger. Curr. Top. Med. Chem. 2002, 2, 181–197. [Google Scholar] [CrossRef] [Green Version]
  143. Bromme, H.; Morke, W.; Peschke, E.; Ebelt, H. Scavenging effect of melatonin on hydroxyl radicals generated by alloxan. J. Pineal Res. 2000, 29, 201–208. [Google Scholar] [CrossRef]
  144. Valko, M.; Morris, H.; Cronin, M.T.D. Metals, toxicity and oxidative stress. Curr. Med. Chem. 2005, 12, 1161–1208. [Google Scholar] [CrossRef] [Green Version]
  145. Limson, J.; Nyokong, T.; Daya, S. The interaction of melatonin and its precursors with aluminium, cadmium, copper, iron, lead, and zinc: An adsorptive voltammetric study. J. Pineal Res. 1998, 24, 15–21. [Google Scholar] [CrossRef]
  146. Galano, A.; Medina, M.E.; Tan, D.X.; Reiter, R.J. Melatonin and its metabolites as copper chelating agents and their role in inhibiting oxidative stress: A physicochemical analysis. J. Pineal Res. 2014, 58, 107–116. [Google Scholar] [CrossRef]
  147. Rodriguez, C.; Mayo, J.C.; Sainz, R.M.; Antolin, I.; Herrera, F.; Martin, V.; Reiter, R.J. Regulation of antioxidant enzymes: A significant role for melatonin. J. Pineal Res. 2004, 36, 1–9. [Google Scholar] [CrossRef]
  148. Sun, T.-C.; Li, H.-Y.; Li, X.-Y.; Yu, K.; Deng, S.-L.; Tian, L. Protective effects of melatonin on male fertility preservation and reproductive system. Cryobiology 2020, 95, 1–8. [Google Scholar] [CrossRef]
  149. Rosen, J.; Ni Than, N.; Koch, D.; Poeggeler, B.; Laatsch, H.; Hardeland, R. Interactions of melatonin and its metabolites with the ABTS cation radical: Extension of the radical scavenger cascade and formation of a novel class of oxidation products, C2-substituted 3-indolinones. J. Pineal Res. 2006, 41, 374–381. [Google Scholar] [CrossRef]
  150. Tan, D.-X.; Manchester, L.C.; Terron, M.P.; Flores, L.J.; Reiter, R.J. One molecule, many derivatives: A never-ending interaction of melatonin with reactive oxygen and nitrogen species? J. Pineal Res. 2007, 42, 28–42. [Google Scholar] [CrossRef]
  151. Mannino, G.; Pernici, C.; Serio, G.; Gentile, C.; Bertea, C.M. Melatonin and Phytomelatonin: Chemistry, Biosynthesis, Metabolism, Distribution and Bioactivity in Plants and Animals—An Overview. Int. J. Mol. Sci. 2021, 22, 9996. [Google Scholar] [CrossRef]
  152. Simopoulos, A.P.; Tan, D.-X.; Manchester, L.C.; Reiter, R.J. Purslane: A plant source of omega-3 fatty acids and melatonin. J. Pineal Res. 2005, 39, 331–332. [Google Scholar] [CrossRef]
  153. García, J.J.; López-Pingarrón, L.; Almeida-Souza, P.; Tres, A.; Escudero, P.; García-Gil, F.A.; Tan, D.-X.; Reiter, R.J.; Ramírez, J.M.; Bernal-Pérez, M. Protective effects of melatonin in reducing oxidative stress and in preserving the fluidity of biological membranes: A review. J. Pineal Res. 2014, 56, 225–237. [Google Scholar] [CrossRef] [PubMed]
  154. Deng, W.-S.; Xu, Q.; Liu, Y.; Jiang, C.-H.; Zhou, H.; Gu, L. Effects of melatonin on liver function and lipid peroxidation in a rat model of hepatic ischemia/reperfusion injury. Exp. Ther. Med. 2016, 11, 1955–1960. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Mollaoglu, H.; Topal, T.; Ozler, M.; Uysal, B.; Reiter, R.J.; Korkmaz, A.; Oter, S. Antioxidant effects of melatonin in rats during chronic exposure to hyperbaric oxygen. J. Pineal Res. 2007, 42, 50–54. [Google Scholar] [CrossRef] [PubMed]
  156. Waseem, M.; Tabassum, H.; Parvez, S. Neuroprotective effects of melatonin as evidenced by abrogation of oxaliplatin induced behavioral alterations, mitochondrial dysfunction and neurotoxicity in rat brain. Mitochondrion 2016, 30, 168–176. [Google Scholar] [CrossRef] [PubMed]
  157. Reiter, R.J.; Tan, D.X.; Kim, S.J.; Qi, W. Melatonin as a pharmacological agent against oxidative damage to lipids and DNA. Proc. West. Pharmacol. Soc. 1998, 41, 229–236. [Google Scholar]
  158. Chua, S.; Lee, F.-Y.; Chiang, H.-J.; Chen, K.-H.; Lu, H.-I.; Chen, Y.-T.; Yang, C.-C.; Lin, K.-C.; Chen, Y.-L.; Kao, G.-S.; et al. The cardioprotective effect of melatonin and exendin-4 treatment in a rat model of cardiorenal syndrome. J. Pineal Res. 2016, 61, 438–456. [Google Scholar] [CrossRef]
  159. Mayo, J.C.; Sainz, R.M.; González-Menéndez, P.; Hevia, D.; Cernuda-Cernuda, R. Melatonin transport into mitochondria. Cell. Mol. Life Sci. 2017, 74, 3927–3940. [Google Scholar] [CrossRef]
  160. Yamamoto, H.-A.; Tang, H.-W. Preventive effect of melatonin against cyanide-induced seizures and lipid peroxidation in mice. Neurosci. Lett. 1996, 207, 89–92. [Google Scholar] [CrossRef]
  161. Reiter, R.J.; Rosales-Corral, S.; Tan, D.X.; Jou, M.J.; Galano, A.; Xu, B. Melatonin as a mitochondria-targeted antioxidant: One of evolution’s best ideas. Cell. Mol. Life Sci. 2017, 74, 3863–3881. [Google Scholar] [CrossRef]
  162. Malmir, M.; Noreini, S.N.; Ghafarizadeh, A.; Faraji, T.; Asali, Z. Ameliorative effect of melatonin on apoptosis, DNA fragmentation, membrane integrity and lipid peroxidation of spermatozoa in the idiopathic asthenoteratospermic men: In vitro. Andrologia 2020, 53, e13944. [Google Scholar] [CrossRef]
  163. Nordlund, J.J.; Lerner, A.B. The Effects of Oral Melatonin on Skin Color and on the Release of Pituitary Hormones. J. Clin. Endocrinol. Metab. 1977, 45, 768–774. [Google Scholar] [CrossRef]
  164. Jan, J.E.; Hamilton, D.; Seward, N.; Fast, D.K.; Freeman, R.D.; Laudon, M. Clinical trials of controlled-release melatonin in children with sleep-wake cycle disorders. J. Pineal Res. 2000, 29, 34–39. [Google Scholar] [CrossRef] [Green Version]
  165. Seabra, M.D.L.V.; Bignotto, M.; Pinto, L.R., Jr.; Tufik, S. Randomized, double-blind clinical trial, controlled with placebo, of the toxicology of chronic melatonin treatment. J. Pineal Res. 2000, 29, 193–200. [Google Scholar] [CrossRef] [Green Version]
  166. Agarwal, A.; Mulgund, A.; Sharma, R.; Sabanegh, E. Mechanisms of oligozoospermia: An oxidative stress perspective. Syst. Biol. Reprod. Med. 2014, 60, 206–216. [Google Scholar] [CrossRef] [Green Version]
  167. Desai, N.; Sharma, R.; Makker, K.; Sabanegh, E.; Agarwal, A. Physiologic and pathologic levels of reactive oxygen species in neat semen of infertile men. Fertil. Steril. 2009, 92, 1626–1631. [Google Scholar] [CrossRef]
  168. Mahfouz, R.; Sharma, R.; Thiyagarajan, A.; Kale, V.; Gupta, S.; Sabanegh, E.; Agarwal, A. Semen characteristics and sperm DNA fragmentation in infertile men with low and high levels of seminal reactive oxygen species. Fertil. Steril. 2010, 94, 2141–2146. [Google Scholar] [CrossRef]
  169. Gharagozloo, P.; Aitken, R.J. The role of sperm oxidative stress in male infertility and the significance of oral antioxidant therapy. Hum. Reprod. 2011, 26, 1628–1640. [Google Scholar] [CrossRef]
  170. Iommiello, V.M.; Albani, E.; Di Rosa, A.; Marras, A.; Menduni, F.; Morreale, G.; Levi, S.L.; Pisano, B.; Levi-Setti, P.E. Ejaculate Oxidative Stress Is Related with Sperm DNA Fragmentation and Round Cells. Int. J. Endocrinol. 2015, 2015, 321901. [Google Scholar] [CrossRef]
  171. Showell, M.G.; Mackenzie-Proctor, R.; Brown, J.; Yazdani, A.; Stankiewicz, M.T.; Hart, R.J. Antioxidants for male subfertility. Cochrane Database Syst. Rev. 2014, 12, CD007411. [Google Scholar] [CrossRef]
  172. Steiner, A.Z.; Hansen, K.R.; Barnhart, K.T.; Cedars, M.I.; Legro, R.S.; Diamond, M.P.; Krawetz, S.A.; Usadi, R.; Baker, V.L.; Coward, R.M.; et al. The effect of antioxidants on male factor infertility: The Males, Antioxidants, and Infertility (MOXI) randomized clinical trial. Fertil. Steril. 2020, 113, 552–560.e3. [Google Scholar] [CrossRef]
  173. Jannatifar, R.; Parivar, K.; Roodbari, N.H.; Nasr-Esfahani, M.H. Effects of N-acetyl-cysteine supplementation on sperm quality, chromatin integrity and level of oxidative stress in infertile men. Reprod. Biol. Endocrinol. 2019, 17, 24. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Alahmar, A.T.; Sengupta, P. Impact of Coenzyme Q10 and Selenium on Seminal Fluid Parameters and Antioxidant Status in Men with Idiopathic Infertility. Biol. Trace Elem. Res. 2020, 199, 1246–1252. [Google Scholar] [CrossRef] [PubMed]
  175. Stenqvist, A.; Oleszczuk, K.; Leijonhufvud, I.; Giwercman, A. Impact of antioxidant treatment on DNA fragmentation index: A double-blind placebo-controlled randomized trial. Andrology 2018, 6, 811–816. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Kessopoulou, E.; Powers, H.J.; Sharma, K.K.; Pearson, M.J.; Russell, J.M.; Cooke, I.D.; Barratt, C.L.R. A double-blind randomized placebo cross-over controlled trial using the antioxidant vitamin E to treat reactive oxygen species associated male infertility. Fertil. Steril. 1995, 64, 825–831. [Google Scholar] [CrossRef] [PubMed]
  177. Hawkes, W.C.; Alkan, Z.; Wong, K. Selenium Supplementation Does Not Affect Testicular Selenium Status or Semen Quality in North American Men. J. Androl. 2009, 30, 525–533. [Google Scholar] [CrossRef]
  178. Moilanen, J.; Hovatta, O.; Lindroth, L. Vitamin E levels in seminal plasma can be elevated by oral administration of vitamin E in infertile men. Int. J. Androl. 1993, 16, 165–166. [Google Scholar] [CrossRef]
  179. Rolf, C.; Cooper, T.G.; Yeung, C.H.; Nieschlag, E. Antioxidant treatment of patients with asthenozoospermia or moderate oligoasthenozoospermia with high-dose vitamin C and vitamin E: A randomized, placebo-controlled, double-blind study. Hum. Reprod. 1999, 14, 1028–1033. [Google Scholar] [CrossRef]
  180. Sigman, M.; Glass, S.; Campagnone, J.; Pryor, J.L. Carnitine for the treatment of idiopathic asthenospermia: A randomized, double-blind, placebo-controlled trial. Fertil. Steril. 2006, 85, 1409–1414. [Google Scholar] [CrossRef]
Figure 1. A number of environmental and lifestyle factors have been linked to increased levels of ROS, being the possible starting point of OS in spermatozoa. Regardless of their origin, an increased ROS concentration could be the driving factor for the establishment of mitochondrial ROS production, resulting in a self-perpetuating cycle that ultimately produces several cellular dysfunctions and constitutes the basis of the pathophysiology that underlies the association between ROS levels and male infertility. MDA: malondialdehyde; 4-HNE: 4-hydroxynonenal. This figure is the author’s original work using Mind the Graph Scientific Illustrator.
Figure 1. A number of environmental and lifestyle factors have been linked to increased levels of ROS, being the possible starting point of OS in spermatozoa. Regardless of their origin, an increased ROS concentration could be the driving factor for the establishment of mitochondrial ROS production, resulting in a self-perpetuating cycle that ultimately produces several cellular dysfunctions and constitutes the basis of the pathophysiology that underlies the association between ROS levels and male infertility. MDA: malondialdehyde; 4-HNE: 4-hydroxynonenal. This figure is the author’s original work using Mind the Graph Scientific Illustrator.
Biomedicines 10 03058 g001
Figure 2. Proposed antioxidant regimen: (a) Long chain ω-3 polyunsaturated fatty acids; (b) Resveratrol; (c) Melatonin.
Figure 2. Proposed antioxidant regimen: (a) Long chain ω-3 polyunsaturated fatty acids; (b) Resveratrol; (c) Melatonin.
Biomedicines 10 03058 g002
Figure 3. Timeframes in human spermatogenesis and proposed antioxidant intervention. Ad, Dark type A spermatogonium; Ap, pale type A spermatogonium; B, type B spermatogonium; II, secondary spermatocyte; L, leptotene spermatocyte; P, pachytene spermatocyte; R, resting or preleptotene primary spermatocyte; Rb, residual body; Sa (a), Sb1 (b1), Sb2 (b2), Sc (c), Sd1 (d1), Sd2 (d2), spermatids; Z, zygotene spermatocyte. (Adapted from Servier Medical Art; Campbell-Walsh-Wein Urology, Part VI: Reproductive and Sexual Function, page 1398; and Henry TD, Porucznik CA, Honda TJ et al. Differential impacts of particulate air pollution exposure on early and late stages of spermatogenesis. Ecotoxicol Environ Saf. 2021; 220:112419).
Figure 3. Timeframes in human spermatogenesis and proposed antioxidant intervention. Ad, Dark type A spermatogonium; Ap, pale type A spermatogonium; B, type B spermatogonium; II, secondary spermatocyte; L, leptotene spermatocyte; P, pachytene spermatocyte; R, resting or preleptotene primary spermatocyte; Rb, residual body; Sa (a), Sb1 (b1), Sb2 (b2), Sc (c), Sd1 (d1), Sd2 (d2), spermatids; Z, zygotene spermatocyte. (Adapted from Servier Medical Art; Campbell-Walsh-Wein Urology, Part VI: Reproductive and Sexual Function, page 1398; and Henry TD, Porucznik CA, Honda TJ et al. Differential impacts of particulate air pollution exposure on early and late stages of spermatogenesis. Ecotoxicol Environ Saf. 2021; 220:112419).
Biomedicines 10 03058 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ávila, C.; Vinay, J.I.; Arese, M.; Saso, L.; Rodrigo, R. Antioxidant Intervention against Male Infertility: Time to Design Novel Strategies. Biomedicines 2022, 10, 3058. https://doi.org/10.3390/biomedicines10123058

AMA Style

Ávila C, Vinay JI, Arese M, Saso L, Rodrigo R. Antioxidant Intervention against Male Infertility: Time to Design Novel Strategies. Biomedicines. 2022; 10(12):3058. https://doi.org/10.3390/biomedicines10123058

Chicago/Turabian Style

Ávila, Cristóbal, José Ignacio Vinay, Marzia Arese, Luciano Saso, and Ramón Rodrigo. 2022. "Antioxidant Intervention against Male Infertility: Time to Design Novel Strategies" Biomedicines 10, no. 12: 3058. https://doi.org/10.3390/biomedicines10123058

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

Article Metrics

Back to TopTop