aitken 2016 causes and consequences of oxidative stress in spermatozoa

11
See discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/285619565 Causes and consequences of oxidative stress in spermatozoa Article in Reproduction Fertility and Development · January 2016 DOI: 10.1071/RD15325 CITATIONS 7 READS 566 5 authors, including: Some of the authors of this publication are also working on these related projects: small noncoding RNAs in the male reproductive system View project Mammalian Sperm Membrane Protein Complexes View project Zamira Gibb University of Newcastle 29 PUBLICATIONS 266 CITATIONS SEE PROFILE Mark A Baker University of Newcastle 97 PUBLICATIONS 3,491 CITATIONS SEE PROFILE Joel R Drevet Université Clermont Auvergne 115 PUBLICATIONS 2,621 CITATIONS SEE PROFILE Parviz Gharagozloo CellOxess Biotechnology 27 PUBLICATIONS 682 CITATIONS SEE PROFILE All content following this page was uploaded by Parviz Gharagozloo on 07 December 2015. The user has requested enhancement of the downloaded file. All in-text references underlined in blue are added to the original document and are linked to publications on ResearchGate, letting you access and read them immediately.

Upload: gloria-maria-levano-sanchez

Post on 05-Apr-2017

186 views

Category:

Technology


0 download

TRANSCRIPT

Page 1: Aitken 2016 causes and consequences of oxidative stress in spermatozoa

Seediscussions,stats,andauthorprofilesforthispublicationat:https://www.researchgate.net/publication/285619565

Causesandconsequencesofoxidativestressinspermatozoa

ArticleinReproductionFertilityandDevelopment·January2016

DOI:10.1071/RD15325

CITATIONS

7

READS

566

5authors,including:

Someoftheauthorsofthispublicationarealsoworkingontheserelatedprojects:

smallnoncodingRNAsinthemalereproductivesystemViewproject

MammalianSpermMembraneProteinComplexesViewproject

ZamiraGibb

UniversityofNewcastle

29PUBLICATIONS266CITATIONS

SEEPROFILE

MarkABaker

UniversityofNewcastle

97PUBLICATIONS3,491CITATIONS

SEEPROFILE

JoelRDrevet

UniversitéClermontAuvergne

115PUBLICATIONS2,621CITATIONS

SEEPROFILE

ParvizGharagozloo

CellOxessBiotechnology

27PUBLICATIONS682CITATIONS

SEEPROFILE

AllcontentfollowingthispagewasuploadedbyParvizGharagozlooon07December2015.

Theuserhasrequestedenhancementofthedownloadedfile.Allin-textreferencesunderlinedinblueareaddedtotheoriginaldocument

andarelinkedtopublicationsonResearchGate,lettingyouaccessandreadthemimmediately.

Page 2: Aitken 2016 causes and consequences of oxidative stress in spermatozoa

Causes and consequences of oxidative stressin spermatozoa

Robert John AitkenA,D, Zamira GibbA, Mark A. BakerA, Joel DrevetB

and Parviz GharagozlooC

APriority Research Centre in Reproductive Science and Hunter Medical Research Institute,

Faculty of Science and IT, University of Newcastle, Callaghan, NSW 2308, Australia.BGReD laboratory, CNRS UMR6293-INSERM U1103-Clermont Universite, 63171 BP80006,

Aubiere cedex, France.CCellOxess LLC, 15 Roszel Street, Princeton, NJ 08540, USA.DCorresponding author. Email: [email protected]

Abstract. Spermatozoa are highly vulnerable to oxidative attack because they lack significant antioxidant protectiondue to the limited volume and restricted distribution of cytoplasmic space in which to house an appropriate armoury of

defensive enzymes. In particular, sperm membrane lipids are susceptible to oxidative stress because they aboundin significant amounts of polyunsaturated fatty acids. Susceptibility to oxidative attack is further exacerbated by thefact that these cells actively generate reactive oxygen species (ROS) in order to drive the increase in tyrosinephosphorylation associated with sperm capacitation. However, this positive role for ROS is reversed when spermatozoa

are stressed. Under these conditions, they default to an intrinsic apoptotic pathway characterised by mitochondrial ROSgeneration, loss of mitochondrial membrane potential, caspase activation, phosphatidylserine exposure and oxidativeDNA damage. In responding to oxidative stress, spermatozoa only possess the first enzyme in the base excision repair

pathway, 8-oxoguanine DNA glycosylase. This enzyme catalyses the formation of abasic sites, thereby destabilising theDNA backbone and generating strand breaks. Because oxidative damage to sperm DNA is associated with bothmiscarriage and developmental abnormalities in the offspring, strategies for the amelioration of such stress, including

the development of effective antioxidant formulations, are becoming increasingly urgent.

Additional keywords: apoptosis, fertilizing potential, lipid peroxidation, male germ line, oxidative DNA damage,ROS generation.

Introduction

Traditionally, spermatozoa are regarded as highly specialised

cells that have but one function in life: to achieve fertilisationand deliver the paternal component of the embryonic genometo an MII oocyte. Although defective sperm function has longbeen recognised as a major cause of human infertility (Hull

et al. 1985), this condition has conventionally been equatedwith the ability of spermatozoa to achieve fertilisation (Aitkenet al. 1987; Aitken 2006). With the passage of time, we have

come to understand that the functional competence of sper-matozoa cannot be defined merely in terms of the ability ofthese cells to fertilise an oocyte; it also needs to incorporate an

assessment of their ability to program a normal pattern ofembryonic development. Spermatozoa may affect embryonicdevelopment via both genetic and a variety of epigeneticmechanisms involving the methylation profile of the DNA,

the post-translational modification of nuclear histones and thecomposition of a variety of coding and non-coding RNAspecies that are integrated into these cells, to find ultimate

expression in the zygote and early embryo (Aitken 1999;Ostermeier et al. 2005; Prescott et al. 2012; Hosken and

Hodgson 2014; Metzler-Guillemain et al. 2015; Soubry 2015).These sperm-borne epigenetic marks are, in turn, affectedby a variety of paternal factors, including genotype, age,obesity, smoking and exposure to environmental contaminants

(Aitken 2014).The mechanisms by which such environmental and lifestyle

factors affect mammalian spermatozoa, to define both their

potential for fertilisation and the subsequent initiation of embry-onic development, are poorly understood. The central hypothe-sis outlined in this article is that all such factors ultimately

converge to induce a high level of oxidative stress in the malegermline. Oxidative stress is known to interfere with thefertilising capacity of spermatozoa, to damage sperm nuclearDNA and to affect the epigenetic profile of these cells (Aitken

et al. 2014b). Herein, we review the evidence relating to theorigins and consequences of such stress and consider potentialstrategies for its remediation.

CSIRO PUBLISHING

Reproduction, Fertility and Development, 2016, 28, 1–10

http://dx.doi.org/10.1071/RD15325

Journal compilation � IETS 2016 www.publish.csiro.au/journals/rfd

Page 3: Aitken 2016 causes and consequences of oxidative stress in spermatozoa

Oxidative stress and fertilisation potential

The notion that sperm function may be compromised by theonset of oxidative stress can be traced back to the early studies ofEvans (1947), who observed that heavily irradiated seawater

impaired the fertilising capacity of sea urchin spermatozoa. Heconcluded that the irradiation process had generated hydrogenperoxide (H2O2) and that this powerful oxidising agent wasdamaging to spermatozoa. Although this conclusion was not

subsequently supported by Barron et al. (1949), these authorsdid generate unequivocal data indicating that H2O2 is extremelydamaging to sperm function. Around the same time, Tosic and

Walton (1946) demonstrated that the metabolite generated bybovine spermatozoa in the presence of egg yolk-based cryo-preservatives was H2O2 and that this oxidant actively sup-

pressed their respiration. Furthermore, these authors identifiedthe source of the H2O2 to be an L-amino acid oxidase with anaffinity for aromatic amino acids, particularly phenylalanine,

which is abundant in egg yolk (Tosic and Walton 1950).MacLeod (1943) also demonstrated that human spermatozoalost motility at high oxygen tensions via mechanisms that couldbe reversed by catalase, again suggesting that H2O2 generation

was causally involved in the loss of sperm function. The par-ticular destructive power of H2O2 relative to any other reactiveoxygen species (ROS) was later emphasised in studies revealing

that catalase, but not superoxide dismutase, was able to relievethe detrimental effect of ROS generated by the xanthine oxidasefree radical-generating system on human spermmotility in vitro

(Aitken et al. 1993a).The possibility that excess ROS generationmay be associated

with defective sperm function in vivo was highlighted by two

papers that appeared in 1987 and demonstrated that the sperma-tozoa of infertilemales were characterised by high levels of ROSgeneration and the induction of lipid peroxidation (Aitken andClarkson 1987; Alvarez et al. 1987). The susceptibility of human

spermatozoa to lipid peroxidation had previously been highlight-ed by Thaddeus Mann (Jones et al. 1979), who pointed out thatthese cells contain exceptionally high levels of polyunsaturated

fatty acids (PUFA; particularly docosahexanoic acid), which arevulnerable to free radical attack, generating lipid peroxides andaldehydes that have a direct inhibitory action on sperm move-

ment. These early studies have subsequently been confirmedin many independent laboratories, all of which agree on thefundamental tenet that defective sperm function is frequentlyinduced by oxidative stress, affecting the motility of these cells,

their DNA integrity and their competence for sperm–oocytefusion (e.g. Aitken et al. 1991, 2010; Zalata et al. 1995; Sanockaet al. 1996; Sharma and Agarwal 1996; Nakamura et al. 2002;

Kao et al. 2008; Sakamoto et al. 2008; Bejarano et al. 2014;Morielli and O’Flaherty 2015).

Oxidative stress and lipid peroxidation

The way in which oxidative stress suppresses sperm motilityappears to be directly related to the induction of lipid per-

oxidation. When ROS attack the PUFA that abound in humanspermatozoa, a variety of lipid metabolites is generated,including lipid peroxyl radicals, alkoxyl radicals and variousaldehydes, such as malondialdehyde, 4-hydroxynonenal (4HNE)

and acrolein (Jones et al. 1978; Moazamian et al. 2015). Theaddition of both lipid peroxides and lipid aldehydes to popula-

tions of human spermatozoa results in the rapid immobilisationof these cells via different mechanisms. Lipid peroxyl radicalsdestabilise the sperm plasma membrane by virtue of their ten-

dency to abstract hydrogen atoms from adjacent PUFA toachieve a measure of stabilisation as the corresponding lipidhydroperoxide. This process creates carbon-centred lipid radi-

cals that combine with oxygen to generate more peroxyl radi-cals, which, in turn, abstract hydrogen from adjacent PUFA tostabilise, generating additional lipid radicals and promoting thepropagation of the lipid peroxidation chain reaction (Fig. 1a).

The lipid peroxides generated in this process destabilise theplasma membrane by becoming targets for phospholipase A2,which moves into the plasma membranes to cleave out the lipid

peroxides so they can be further processed by glutathione per-oxidase (van Kuijk et al. 1985). This process, in turn, generateslysophospholipids that destabilise the sperm plasma membrane,

affecting the microarchitecture of this structure and changingthe functions of integral membrane proteins that are critical tothe maintenance of sperm motility, such as ATP-dependent ionpumps and voltage-regulated ion channels (Nishikawa et al.

1989; Lundbaek and Andersen 1994). The disruptive effect ofperoxidative damage on lipid membrane architecture alsoaffects the ability of spermatozoa to participate in themembrane

fusion events associated with fertilisation (Aitken et al. 1989,1993b, 1993c).

The lipid peroxidation chain reactions initiated in spermato-

zoa may also result in the formation of a cascade of aldehydeby-products that include alkanals, such asmalondialdehyde, andalkenals, such as 4HNE and acrolein. These compounds, partic-

ularly 4HNE and acrolein, are powerful electrophiles that formadducts with several proteins within the spermatozoa that, inturn, affect sperm function. For example, the formation ofadductswith the flagellar axonemal protein, dynein heavy chain,

may explain the effect of these aldehydes on sperm movement(Baker et al. 2015; Moazamian et al. 2015). In addition, 4HNEhas been shown to bind to mitochondrial proteins in human

spermatozoa, triggering electron leakage and the formation ofROS (Fig. 1b). The oxidative stress associated with the latterthen forces the spermatozoa to enter the intrinsic apoptotic

cascade, beginning with a loss of mitochondrial membranepotential and terminating in oxidative DNA adduct formation,DNA strand breakage and cell death (Aitken et al. 2012).

Sources of ROS and oxidative stress in spermatozoa

With oxidative stress being such a major factor in the aetiology

of defective human sperm function, resolving the possiblecauses of this condition is critical. In considering thismatter, it isimportant to emphasise that spermatozoa are not only vulnerable

to oxidative stress because of the targets they offer for freeradical attack in the form of PUFA, proteins and nucleic acids,but they are also lacking significant intracellular antioxidant

protection, including ROS-metabolising enzymes, such ascatalase and glutathione peroxidase, by virtue of the limitedvolume and restricted distribution of cytoplasmic space inwhichto house such mediators of cell survival. Furthermore, these

2 Reproduction, Fertility and Development R. J. Aitken et al.

Page 4: Aitken 2016 causes and consequences of oxidative stress in spermatozoa

cells actively generate physiological levels of ROS in order to

drive the tyrosine phosphorylation events associated with spermcapacitation (Aitken and Nixon 2013). The involvement of ROSin the capacitation of mammalian spermatozoa has been

appreciated since the pioneering studies of Claude Gagnon inthe 1990s (de Lamirande and Gagnon 1993a). The ROSresponsible for sperm capacitation have been variously reported

as H2O2 (Bise et al. 1991; Aitken et al. 1995, 1996; Rivlin et al.2004) superoxide anion (de Lamirande and Gagnon 1993b) andthe peroxynitrite radical generated by the reaction of superoxide

anion with another free radical species, namely nitric oxide(Herrero et al. 2001; Rodriguez and Beconi 2009). In reality, theinterconversion of these various ROS and reactive nitrogenspecies is very rapid and it is probable that several different

redox entities are involved in various aspects of the capacitationprocess, including the suppression of tyrosine phosphatase

activity and the stimulation of cAMP generation (Aitken and

Nixon 2013). It has recently been hypothesised that the con-tinued generation of ROS, particularly peroxynitrite, to achievecapacitation ultimately overwhelms the limited antioxidant

defences of these cells and precipitates a state of apoptosis.According to this concept, capacitation and the intrinsic apo-ptotic cascade are the opposite ends of a metabolic continuum

driven by ROS (Aitken et al. 2015b).If ROS are so important for sperm function, what is the

subcellular source of these molecules? In mammalian sperma-

tozoa there can be little doubt that the major sources of ROS arethe mitochondria. Human sperm mitochondria are particularlyactive in the generation of ROS via mechanisms that are notdependent on a loss of mitochondrial membrane potential

(Koppers et al. 2008). Activation of ROS generation at ComplexIII was found to stimulate the rapid release of H2O2 into the

(a) PUFA

R

Freeradicalattack

H

O2

Initiation

OO •

R

Lipid radical

Mitochondrial ROS generation

Lipid radical

PUFA Peroxyl radical Lipid hydroperoxide

R

R

H

R

Hydrogenabstraction

R

OOHOO •

R

Peroxyl radical

Lipid aldehydes4HNE/acrolein

(c)H2O2

(b)

� H2O•OH

Nucleus

Mitochondria

Adduction ofmitochondrial

proteins

Lipidperoxidation

Fig. 1. Oxidative stress in mammalian spermatozoa. (a) Spermatozoa are susceptible to oxidative stress because they contain high

concentrations of polyunsaturated fatty acids (PUFA). Free radical attack leads to the formation of lipid radicals that then combinewith the

universal electron acceptor, oxygen, to generate a lipid peroxyl radical. In order to stabilise as a hydroperoxide, the latter extracts hydrogen

atoms from adjacent lipids, generating lipid radicals that then perpetuate the peroxidation cascade. (b) Lipid aldehydes generated as a

consequence of lipid peroxidation, such as 4-hydroxynonenal (4HNE), bind to mitochondrial proteins, including succinic acid

dehydrogenase and stimulate yet more free radical generation, further enhancing lipid peroxidation in a self-propagating cycle that

propels spermatozoa towards an apoptotic fate. (c) The unusual architecture of spermatozoameans that nucleases activated in themidpiece

cytoplasm, or released from the mitochondria, cannot enter the nuclear compartment. The only product of apoptosis that can pass from the

midpiece to the sperm head to damage the DNA is H2O2; this is why most DNA damage in spermatozoa is oxidative.

Oxidative stress in spermatozoa Reproduction, Fertility and Development 3

Page 5: Aitken 2016 causes and consequences of oxidative stress in spermatozoa

extracellular space, but no detectable peroxidative damage.Conversely, the induction of ROS on the matrix side of

the inner mitochondrial membrane at Complex I resulted inperoxidative damage to the midpiece and a loss of spermmovement that could be prevented by the concomitant presence

of a-tocopherol (Koppers et al. 2008). Defective human sper-matozoa spontaneously generate mitochondrial ROS in a man-ner that is negatively correlated with motility (Koppers et al.

2008). Indeed, simultaneous measurement of total cellular ROSwith dihydroethidium indicated that 68% of the variability insuch measurements could be explained by differences in mito-chondrial ROS production (Koppers et al. 2008).

Another potential source of ROS are the NADPH oxidaseenzymes (NOX), including the calcium-dependent NOX5,which are known to be present in the spermatozoa of certain

species, including human (Banfi et al. 2001), although otherspecies, such as the mouse, do not possess this enzyme. Expos-ing spermatozoa to NADPH can trigger a redox response that is

detectable with the redox probe lucigenin and inhibitable bydiphenylene iodonium (DPI), a flavoprotein inhibitor (Aitkenet al. 1997; Vernet et al. 2001). However, this lucigenin-dependent activity was subsequently shown to be due to the

direct enzymatic reduction of the probe by cytochrome P450reductase (Baker et al. 2004) and cytochrome b5 reductase(Baker et al. 2005) when the electron donors were NADPH

and NADH, respectively. In contrast, using luminol as aROS probe, clear evidence has been obtained for a calcium-dependent increase in ROS generation, which is particularly

marked in the spermatozoa of infertile patients and potentiallyreflective of an involvement of NOX5 in the aetiology ofdefective sperm function (Aitken and Clarkson 1987). There

is even some evidence to suggest that NOX5 may be overrepre-sented in the defective spermatozoa recovered from patientsexhibiting teratozoospermia (Ghani et al. 2013). However,definitive proof that NOX5 is the source of ROS under such

circumstances is currently lacking. There is a possibility that thecalcium-dependent signals observed with unfractionated spermsuspensions are the result of low-level leucocyte contamination

(Aitken and Clarkson 1987; Aitken et al. 1992). The abilityof the NOX inhibitor apocynin to suppress the ROS signalsgenerated by human sperm suspensions (Dona et al. 2011) could

also be accounted for by leucocyte contamination because thisreagent prevents assembly of the key cytosolic componentsof the NADPH oxidase system (p40phox, p47phox and p67phox),which is not necessary for NOX5 to be active. A detailed study

of the NOX species present in human spermatozoa is currentlylacking and the role of these enzymes in the creation of oxidativestress within the germline remains unresolved. One possibility

that cannot be excluded is that the NOX enzymes present inmammalian spermatozoa play no role at all in the regulation ofsperm function, but rather function much earlier in germ cell

production, controlling spermatogonial stem cell proliferation(Morimoto et al. 2013).

Finally, L-amino acid oxidases with a particular affinity for

phenylalanine have been identified in bull, horse, human andram spermatozoa (Tosic and Walton 1946; Aitken et al. 2015a;Houston et al. 2015). In the case of equine spermatozoa, whichare heavily dependent on oxidative phosphorylation (Gibb et al.

2014), the primary role for this amino acid oxidase may be tosupport the energy metabolism of these cells through the

oxidative deamination of aromatic amino acids, generating ketoacids that are then processed by the sperm mitochondria.However, in the case of human spermatozoa, oxidative phos-

phorylation appears to play a minor role in sperm metabolismbecause these cells are largely dependent of glycolysis to meettheir energy needs (du Plessis et al. 2015). In these cells, the

L-amino acid oxidase (interleukin 4 induced protein 1, IL4I1)seems to have acquired a new biological function in supplyingthe redox drive to sperm capacitation (Houston et al. 2015).

Role of oxidative stress in DNA damage

One of the major complications associated with male infertility

is the presence of high levels of DNA damage in the sperma-tozoa. Such damage can arise as a consequence of infertility(Irvine et al. 2000; Aitken and Curry 2011) age (Singh et al.

2003) smoking (Fraga et al. 1996) antioxidant deficiency (Fragaet al. 1991, 1996), obesity (Fariello et al. 2012) exposure toinfection (Reichart et al. 2000; Burrello et al. 2004), heat(De Iuliis et al. 2009a; Santiso et al. 2012) acidic pH (Santiso

et al. 2012), metals, particularly transition metals such as ironand copper (Aitken et al. 2014a), radiofrequency electromag-netic radiation (De Iuliis et al. 2009a), ionising radiation (Singh

and Stephens 1998), environmental toxicants such as acrylam-ide (Katen and Roman 2015), chemotherapeutic agents (Delbeset al. 2010), air pollution, plasticisers, pesticides (Evenson and

Wixon 2005; O’Flaherty 2014) and chloracetanilide herbicidessuch as alachlor (Grizard et al. 2007).

There can be little doubt that most of these factors affect the

integrity of sperm chromatin through the induction of oxidativestress. Such stress results in the generation of oxidised DNAbase adducts such as 8-hydroxy-20-deoxyguanosine (8OHdG),particularly in areas of the genome that are not heavily prota-

minated (De Iuliis et al. 2009b; Noblanc et al. 2013). Spermato-zoa only possess one enzyme in the base excision repair (BER)pathway, 8-oxoguanine DNA glycosylase (OGG1). This glyco-

sylase is associated with the sperm nucleus and mitochondriaand can actively excise 8OHdG, releasing this base adduct intothe extracellular space. Remarkably, spermatozoa do not pos-

sess the downstream components of the BER pathway, namelyapurinic endonuclease 1 (APE1) and X-ray repair complement-ing defective repair in Chinese hamster cells 1 (XRCC1). Thenet result of this truncated DNA repair capacity is to generate

abasic sites at locations that have been affected by 8OHdGformation. Such abasic sites destabilise the ribose–phosphatebackbone, leading to a b-elimination or a ring opening reaction

of the ribose unit and a consequential strand break. This type ofDNA chemistry has been identified as being central to theinitiation of cancer in other cell types. Therefore, oxidative

DNA base lesions are not only potentially mutagenic but,importantly, also contribute indirectly to the DNA fragmenta-tion observed in the patient population (Ohno et al. 2014).

An oxidative involvement in DNA damage to mammalianspermatozoa has been observed in relation to infertility (Shenand Ong 2000; Aitken et al. 2010), heat (De Iuliis et al. 2009a),antioxidant deficiency (Fraga et al. 1991), age (Weir and

4 Reproduction, Fertility and Development R. J. Aitken et al.

Page 6: Aitken 2016 causes and consequences of oxidative stress in spermatozoa

Robaire 2007; Smith et al. 2013a), smoking (Fraga et al. 1991),obesity (Bakos et al. 2011), radiofrequency electromagnetic

radiation (De Iuliis et al. 2009a), herbicides (Grizard et al.

2007), plasticisers (Erkekoglu et al. 2010; Zhou et al. 2010) andchemotherapeutic agents (Ghosh et al. 2002). Indeed, it would

appear that most DNA damage in mammalian spermatozoais the result of an oxidative insult generated as a result of eitherimpaired antioxidant protection because of endogenous (e.g. age)

or exogenous (e.g. phthalate esters) factors or changes in theredox status of spermatozoa because of internal (e.g. mitochon-drial electron leakage) or external (e.g. radiation or alachlor)influences. However, it is also undeniable that not every

spermatozoon afflicted with DNA damage shows signs of oxida-tive stress. Under these conditions, it has been suggested thatnuclease-mediated DNA fragmentation must occur as a result of

spermatozoa defaulting to an apoptotic state rather than oxidativestress (Muratori et al. 2015). This interesting hypothesis isdifficult to reconcile with the fact that apoptosis invariably

involves the induction of oxidative stress (Koppers et al. 2011),so it is difficult to imagine how these phenomena can be separatedin vivo. A possible resolution of this dilemma is set out below.

Role of apoptosis in DNA damage

It is well known that testicular precursor germ cells can undergoapoptosis as part of a physiological process designed to optimisegerm cell : Sertoli cell ratios and to bring a measure of qualitycontrol to the spermatogenic process, ensuring that no defective

germ cells are allowed to differentiate into spermatozoa (Shuklaet al. 2012). Apoptosis may also occur during spermatogenesisin response to adverse circumstances, including heat shock,

ionising radiation, growth factor deprivation and chemothera-peutic agents. The apoptotic process is largely, but not exclu-sively, targeted to spermatocytes and both the intrinsic

mitochondrial pathway and the extrinsic p53/Fas system havebeen implicated as key modulators of this process (Boekelheide2005; Lagos-Cabre and Moreno 2012). However, the focus ofthis discussion is the spermatozoa.

Spermatozoa are highly differentiated, transcriptionallysilent cells that, by virtue of their inert nuclear constitutionand highly specialised architecture, cannot undergo apoptosis in

the conventional sense. Nevertheless, they can undergo atruncated version of this process. One of the key features ofsperm cell biology is that we do not have to expend energy

searching for factors that will induce these cells to undergoapoptosis. Rather, these cells are designed to undergo apoptosis;it is their default position. Spermatozoa are the ultimate symbol

of disposable cell types; indeed, all these cells are destined to diea lonely apoptotic death in the male or female tract. Thefortunate exceptions to this rule are the handful of individualgametes that manage to fertilise an oocyte and, in so doing,

achieve potential immortality for the genotype they carry.Whenapoptosis does eventually occur, it is generally the intrinsicapoptotic cascade that is induced, mediated by the sperm

mitochondria. Although receptor-mediated extrinsic apoptosisremains a theoretical possibility in spermatozoa, no ligands havebeen convincingly described to date that are capable of eliciting

such a response in the fully differentiated gamete. There has

been a claim that bacterial lipopolysaccharide (LPS) can elicitapoptosis in spermatozoa by interacting with Toll-like receptor

(TLR) 2 and TLR4 on the sperm surface (Fujita et al. 2011).However these data have not yet been independently validatedand our research group has not yet been able to achieve apoptosis

using commercially available LPS (R. J. Aitken, unpubl. obs.).As a result, our view is that the mature gamete has very littlecapacity to activate the extrinsic apoptotic cascade, but is

extremely vulnerable to its intrinsic counterpart.Spermatozoa are normally prevented from entering the

intrinsic apoptotic pathway by virtue of the continuing activityof phosphatidylinositol 3-kinase (PI3K; Koppers et al. 2011). If

PI3K is inhibited, then the spermatozoa default to an apoptoticcascade characterised by rapid loss of motility, generation ofROS, caspase activation in the cytosol, annexin V binding to the

cell surface, cytoplasmic vacuolisation and oxidative DNAdamage. The anti-apoptotic action of PI3K appears to dependon its ability to promote the phosphorylation of another kinase,

AKT, which, in turn, is responsible for phosphorylating anti-apoptotic effector proteins such as Bcl-2-associated death pro-moter (BAD). Phosphorylation of the latter is essential for BADto remain associatedwith its cytoplasmic keeper protein, 14-3-3.

However, dephosphorylation allows BAD to orchestrate anapoptotic process that has many similarities with the intrinsicapoptotic cascade observed in somatic cells. There are two

major points of difference between apoptosis in spermatozoaand somatic cells, as follows:

1. Mammalian spermatozoa are structurally different fromsomatic cells in that all the mitochondria and most of thecytoplasm are compartmentalised in the mid-piece of the

cell, physically separated from the DNA in the spermnucleus. As a result, even if apoptosis is activated in thesecells, the endonucleases released from the mitochondria

(e.g. endonuclease G) or activated in the cytoplasm(e.g. caspase-activated DNAse) are physically impeded fromattacking the sperm nucleus (Koppers et al. 2011). The onlyelement of the apoptotic cascade that can exit from the sperm

midpiece and penetrate the nuclear compartment is H2O2. Itis for this reason that most of the DNA damage present inhuman spermatozoa appears to be oxidatively induced

(Fig. 1c; Aitken et al. 2010).2. Spermatozoa are also characterised by a severely truncated

BER pathway, as discussed above, that stalls after OGG1 has

removed the oxidised base to create abasic sites that have tobe further processed by the oocyte following fertilisation.One consequence of spermatozoa lacking the next enzyme

in this pathway, namely APE1, is that these cells cannotcreate the 30-OH termini that are required by the terminaldeoxyribonucleotidyl transferase-mediated dUTP–digoxigeninnick end-labelling (TUNEL) assay. As a result, TUNEL is a

very insensitive methodology for assessing DNA damage inspermatozoa. Under circumstances where DNA damage isinduced by, for example, exposure to H2O2, intracellular and

extracellular 8OHdG can be clearly detected in the affectedsperm suspension and DNA strand breakage can be detectedwith the sperm chromatin structure assay (SCSA); however,

TUNEL signals are not apparent (Smith et al. 2013b).

Oxidative stress in spermatozoa Reproduction, Fertility and Development 5

Page 7: Aitken 2016 causes and consequences of oxidative stress in spermatozoa

The later do eventually appear when the cells are close todeath. At this point, it is possible that a DNAse does become

activated in the spermatozoa (Sotolongo et al. 2005). For thereasons given above, such a nuclease would have to beincorporated into the sperm chromatin, not released from

themitochondria or activated in the cytoplasm. The nature ofthis DNAse (topoisomerase? DNAse 1?) and its mechanismsof activation are not known. It is possible such nuclease

activity may be activated by a rise in intracellular calcium ascells die, because calcium-dependent nuclease activity hasbeen described in this cell type on several independentoccasions by independent groups (Sotolongo et al. 2005;

Sibirtsev et al. 2011).

Apoptosis or oxidative stress causes DNA damage

We can conclude from the foregoing discussion that there are

two mechanisms for damaging DNA in mammalian spermato-zoa. It is now generally acknowledged that a wide variety ofdifferent intrinsic and extrinsic factors converge to generate a

state of oxidative stress in the germline. Once such stress hasbeen initiated, it tends to become accentuated because the lipidaldehydes generated during the peroxidative process bind to

proteins in the mitochondrial electron transport chain, particu-larly succinic acid dehydrogenase, stimulating the generation ofyet more free radicals, more DNA damage and more lipid per-oxidation to continue the downward spiral towards apoptosis

(Fig. 1b; Aitken et al. 2012). There is also an associationbetween oxidative DNA damage in the male germline and poorchromatin protamination during spermiogenesis (De Iuliis et al.

2009b). This relationship may reflect a certain vulnerabilitytowards oxidative stress as a consequence of the failure of spermnuclear DNA to adequately compact. However, it may also be a

consequence of inadequate protamination, because these small,basic proteins are thought to protect the DNA by acting assacrificial antioxidants and by chelating redox active metalssuch as copper (Liang et al. 1999).

The relationship between oxidative stress and apoptosis iscomplex. Clearly, spermatozoa do express the classical markersof apoptosis, such as ROS generation, phosphatidylserine expo-

sure, caspase activation and DNA fragmentation (Koppers et al.2011). If PI3K activity is inhibited with wortmannin, thenhuman spermatozoa rapidly default to the intrinsic apoptotic

cascade, displaying all of the above features, including highlevels of mitochondrial ROS generation (Koppers et al. 2011).Therefore, entry of spermatozoa into the senescence-driven

apoptotic pathway as a consequence of compromised PI3Kactivity inevitably results in an apoptotic cascade involvingthe stimulation of mitochondrial ROS generation and the induc-tion of oxidative DNA damage. Under these circumstances,

apoptosis and oxidative DNA damage are inextricably linked.Similarly, when spermatozoa are exposed to xenobiotics such asalachlor, the induction of apoptosis is inextricably linked with

the induction of oxidative stress (Grizard et al. 2007).Within theinfertile population, DNA damage is again associated with thesimultaneous appearance of oxidative stress and apoptosis

(Wang et al. 2003; Aitken et al. 2010) via mechanisms thatcan be reversed by the sustained administration of antioxidants

such as melatonin (Bejarano et al. 2014). Similarly, cryostorageleads to the induction of oxidatively driven DNA damage and

apoptosis that can be reversed by the presence of the antioxidantquercetin (Zribi et al. 2012). However, there are occasionalcircumstances where DNA damage can be visualised in the

absence of any evidence that the cells have been subjected tooxidative stress (Muratori et al. 2015). Under these circum-stances, it is possible that the apoptosis involves stimulation of a

DNAse that is integrated into the sperm chromatin and somehowbecomes activated when these cells are under stress.

Developmental consequences of oxidative damagein the germ line

Whatever the causes of oxidative stress in the male germline,there can be no doubt that this pathophysiological mechanismleads to both impaired fertility and disrupted embryonic

development. At high levels of oxidative stress, fertilisation isprevented because the damage to the sperm plasma membraneimpairs both the motility of these cells and their competence for

fusion with the oocyte. At lower levels of oxidative stress thespermatozoa can retain their capacity for fertilisation while theDNA in their nuclei is still oxidatively damaged (Aitken et al.

1998). The developmental consequences of fertilising eggs withspermatozoa exhibiting oxidative DNA damage has beenexplored using the glutathione peroxidase 5 (GPx5)-knockoutmouse. In this mouse model, the spermatozoa suffer from

oxidative areas as they descend the epididymis (Chabory et al.

2009). The level of stress experienced by these spermatozoadoes not impair their fertilising capacity, but does induce high

levels of oxidative DNA damage in the sperm nuclei. Theconsequence of this damage can be seen in the developmentalstatus of the embryos when Gpx5-null males are mated with

wild-type females, because such unions are accompanied by asignificant increase in the incidence of miscarriage and devel-opmental abnormalities (Chabory et al. 2009). In related studiesin which mouse spermatozoa were oxidatively damaged by

exposing them to H2O2, several developmental abnormalitieswere observed in the offspring, including a delay in embryonicdevelopment rates, a decrease in the ratio of inner cell mass cells

in the resulting blastocyst and a reduction in implantation rates(Lane et al. 2014). Crown–rump length at Day 18 of gestationwas also reduced in offspring produced from H2O2-treated

spermatozoa. Female offspring from peroxide-treated sperma-tozoa were smaller, became glucose intolerant and accumulatedincreased levels of adipose tissue compared with control female

offspring. Interestingly, the male offspring phenotype was lesssevere, with increases in fat depots only seen at 4 weeks of age,which returned to control levels later in life (Lane et al. 2014).Studies in primates (Burruel et al. 2013) and cattle (Simoes et al.

2013) have confirmed the effect of oxidative DNA damage inspermatozoa on the developmental potential of fertilised ova.Given the developmental significance of this oxidative sperm

DNA damage, it is important that strategies are developed toreduce such pathological changes as a matter of good practicein assisted conception programs and as a matter of good con-

science in couples contemplating parenthood by natural means.The data accumulated to date are encouraging in that the

6 Reproduction, Fertility and Development R. J. Aitken et al.

Page 8: Aitken 2016 causes and consequences of oxidative stress in spermatozoa

oxidative damage in spermatozoa associated with obesity, forexample, can clearly be reversed by a combination of diet and

exercise (Palmer et al. 2012). The use of antioxidants has alsomet with some success in treating idiopathic oxidative stress inspermatozoa (Gharagozloo and Aitken 2011; Showell et al.

2014). The advantages of using an effective oral antioxidanttreatment are that, in cases where oxidative stress has beendiagnosed in the male germline, it should result in increases in

both the fertilising potential of human spermatozoa and theirgenetic integrity. The disadvantage of using antioxidants, par-ticularly as an empirical treatment for patients where there is noevidence of oxidative stress, is that it may create a reductive

stress (Chen et al. 2013).We still await the results of randomiseddouble-blind cross-over trials to definitively establish thetherapeutic value of different antioxidant formulations in the

treatment of male patients exhibiting high levels of oxidativeDNA damage in their spermatozoa. Such studies are eagerlyanticipated.

References

Aitken, R. J. (1999). TheAmoroso Lecture. The human spermatozoon: a cell

in crisis? J. Reprod. Fertil. 115, 1–7. doi:10.1530/JRF.0.1150001

Aitken, R. J. (2006). Sperm function tests and fertility. Int. J. Androl. 29,

69–75. doi:10.1111/J.1365-2605.2005.00630.X

Aitken, R. J. (2014). Age, the environment and our reproductive future:

bonking baby boomers and the future of sex.Reproduction 147, S1–S11.

doi:10.1530/REP-13-0399

Aitken, R. J., and Clarkson, J. S. (1987). Cellular basis of defective sperm

function and its association with the genesis of reactive oxygen species

by human spermatozoa. J. Reprod. Fertil. 81, 459–469. doi:10.1530/

JRF.0.0810459

Aitken, R. J., and Curry, B. J. (2011). Redox regulation of human sperm

function: from the physiological control of sperm capacitation to the

etiology of infertility andDNAdamage in the germ line.Antioxid. Redox

Signal. 14, 367–381. doi:10.1089/ARS.2010.3186

Aitken, R. J., and Nixon, B. (2013). Sperm capacitation: a distant landscape

glimpsed but unexplored.Mol. Hum. Reprod. 19, 785–793. doi:10.1093/

MOLEHR/GAT067

Aitken, R. J., Thatcher, S., Glasier, A. F., Clarkson, J. S., Wu, F. C., and

Baird, D. T. (1987). Relative ability of modified versions of the hamster

oocyte penetration test, incorporating hyperosmotic medium or the

ionophore A23187, to predict IVF outcome. Hum. Reprod. 2, 227–231.

Aitken, R. J., Clarkson, J. S., and Fishel, S. (1989). Generation of reactive

oxygen species, lipid peroxidation, and human sperm function. Biol.

Reprod. 41, 183–197. doi:10.1095/BIOLREPROD41.1.183

Aitken, R. J., Irvine, D. S., and Wu, F. C. (1991). Prospective analysis of

sperm–oocyte fusion and reactive oxygen species generation as criteria

for the diagnosis of infertility. Am. J. Obstet. Gynecol. 164, 542–551.

doi:10.1016/S0002-9378(11)80017-7

Aitken, R. J., Buckingham, D., West, K., Wu, F. C., Zikopoulos, K., and

Richardson, D. W. (1992). Differential contribution of leucocytes and

spermatozoa to the generation of reactive oxygen species in the ejacu-

lates of oligozoospermic patients and fertile donors. J. Reprod. Fertil.

94, 451–462. doi:10.1530/JRF.0.0940451

Aitken, R. J., Buckingham, D., and Harkiss, D. (1993a). Use of a xanthine

oxidase free radical generating system to investigate the cytotoxic

effects of reactive oxygen species on human spermatozoa. J. Reprod.

Fertil. 97, 441–450. doi:10.1530/JRF.0.0970441

Aitken, R. J., Harkiss, D., and Buckingham, D. (1993b). Relationship

between iron-catalysed lipid peroxidation potential and human sperm

function. J. Reprod. Fertil. 98, 257–265. doi:10.1530/JRF.0.0980257

Aitken, R. J., Harkiss, D., and Buckingham, D.W. (1993c). Analysis of lipid

peroxidation mechanisms in human spermatozoa.Mol. Reprod. Dev. 35,

302–315. doi:10.1002/MRD.1080350313

Aitken, R. J., Paterson,M., Fisher, H., Buckingham,D.W., and vanDuin,M.

(1995). Redox regulation of tyrosine phosphorylation in human sperma-

tozoa and its role in the control of human sperm function. J. Cell Sci. 108,

2017–2025.

Aitken, R. J., Buckingham, D. W., Harkiss, D., Paterson, M., Fisher, H., and

Irvine, D. S. (1996). The extragenomic action of progesterone on human

spermatozoa is influenced by redox regulated changes in tyrosine

phosphorylation during capacitation. Mol. Cell. Endocrinol. 117,

83–93. doi:10.1016/0303-7207(95)03733-0

Aitken, R. J., Fisher, H.M., Fulton, N., Gomez, E., Knox,W., Lewis, B., and

Irvine, S. (1997). Reactive oxygen species generation by human sper-

matozoa is induced by exogenous NADPH and inhibited by the flavo-

protein inhibitors diphenylene iodonium and quinacrine. Mol. Reprod.

Dev. 47, 468–482. doi:10.1002/(SICI)1098-2795(199708)47:4,468::

AID-MRD14.3.0.CO;2-S

Aitken, R. J., Gordon, E., Harkiss, D., Twigg, J. P., Milne, P., Jennings, Z.,

and Irvine, D. S. (1998). Relative impact of oxidative stress on the

functional competence and genomic integrity of human spermatozoa.

Biol. Reprod. 59, 1037–1046. doi:10.1095/BIOLREPROD59.5.1037

Aitken, R. J., De Iuliis, G. N., Finnie, J.M., Hedges, A., andMcLachlan, R. I.

(2010). Analysis of the relationships between oxidative stress, DNA

damage and sperm vitality in a patient population: development of

diagnostic criteria. Hum. Reprod. 25, 2415–2426. doi:10.1093/

HUMREP/DEQ214

Aitken, R. J., Whiting, S., De Iuliis, G. N., McClymont, S., Mitchell, L. A.,

and Baker, M. A. (2012). Electrophilic aldehydes generated by sperm

metabolism activate mitochondrial reactive oxygen species generation

and apoptosis by targeting succinate dehydrogenase. J. Biol. Chem. 287,

33 048–33 060. doi:10.1074/JBC.M112.366690

Aitken, R. J., Finnie, J. M., Muscio, L., Whiting, S., Connaughton, H. S.,

Kuczera, L., Rothkirch, T. B., and De Iuliis, G. N. (2014a). Potential

importance of transition metals in the induction of DNA damage by

sperm preparation media. Hum. Reprod. 29, 2136–2147. doi:10.1093/

HUMREP/DEU204

Aitken, R. J., Smith, T. B., Jobling, M. S., Baker, M. A., and De Iuliis, G. N.

(2014b). Oxidative stress and male reproductive health. Asian J. Androl.

16, 31–38. doi:10.4103/1008-682X.122203

Aitken, J. B., Naumovski, N., Grupen, C. G., Gibb, Z., and Aitken, R. J.

(2015a). Characterization of an L-amino acid oxidase in equine sperma-

tozoa. Biol. Reprod. 92, 125. doi:10.1095/BIOLREPROD.114.126052

Aitken, R. J., Baker, M. A., and Nixon, B. (2015b). Are sperm capacitation

and apoptosis the opposite ends of a continuum driven by oxidative

stress? Asian J. Androl. 17, 633–639. doi:10.4103/1008-682X.153850

Alvarez, J. G., Touchstone, J. C., Blasco, L., and Storey, B. T. (1987).

Spontaneous lipid peroxidation and production of hydrogen peroxide

and superoxide in human spermatozoa. Superoxide dismutase as major

enzyme protectant against oxygen toxicity. J. Androl. 8, 338–348.

Baker, M. A., Krutskikh, A., Curry, B. J., McLaughlin, E. A., and Aitken,

R. J. (2004). Identification of cytochrome P450-reductase as the enzyme

responsible for NADPH-dependent lucigenin and tetrazolium salt reduc-

tion in rat epididymal sperm preparations. Biol. Reprod. 71, 307–318.

doi:10.1095/BIOLREPROD.104.027748

Baker, M. A., Krutskikh, A., Curry, B. J., Hetherington, L., and Aitken, R. J.

(2005). Identification of cytochrome-b5 reductase as the enzyme respon-

sible for NADH-dependent lucigenin chemiluminescence in human

spermatozoa. Biol. Reprod. 73, 334–342. doi:10.1095/BIOLREPROD.

104.037960

Baker, M. A., Weinberg, A., Hetherington, L., Villaverde, A. I., Velkov, T.,

Baell, J., and Gordon, C. P. (2015). Defining the mechanisms by which

the reactive oxygen species by-product, 4-hydroxynonenal, affects

Oxidative stress in spermatozoa Reproduction, Fertility and Development 7

Page 9: Aitken 2016 causes and consequences of oxidative stress in spermatozoa

human sperm cell function. Biol. Reprod. 92, 108. doi:10.1095/BIOL

REPROD.114.126680

Bakos, H. W., Mitchell, M., Setchell, B. P., and Lane, M. (2011). The effect

of paternal diet-induced obesity on sperm function and fertilization in a

mouse model. Int. J. Androl. 34, 402–410. doi:10.1111/J.1365-2605.

2010.01092.X

Banfi, B., Molnar, G., Maturana, A., Steger, K., Hegedus, B., Demaurex, N.,

and Krause, K. H. (2001). A Ca(2þ)-activated NADPH oxidase in

testis, spleen, and lymph nodes. J. Biol. Chem. 276, 37 594–37 601.

doi:10.1074/JBC.M103034200

Barron, E. S. G., Flood, V., and Gasvoda, B. (1949). The effect of

hydrogen peroxide and of X-ray irradiated sea water on the respiration

of sea urchin sperm and eggs. Biol. Bull. 97, 51–56. doi:10.2307/

1538093

Bejarano, I., Monllor, F., Marchena, A. M., Ortiz, A., Lozano, G., Jimenez,

M. I., Gaspar, P., Garcıa, J. F., Pariente, J. A., Rodrıguez, A. B., and

Espino, J. (2014). Exogenous melatonin supplementation prevents

oxidative stress-evoked DNA damage in human spermatozoa. J. Pineal

Res. 57, 333–339. doi:10.1111/JPI.12172

Bize, I., Santander, G., Cabello, P., Driscoll, D., and Sharpe, C. (1991).

Hydrogen peroxide is involved in hamster sperm capacitation in vitro.

Biol. Reprod. 44, 398–403. doi:10.1095/BIOLREPROD44.3.398

Boekelheide, K. (2005). Mechanisms of toxic damage to spermatogenesis.

J. Natl Cancer Inst. Monogr. 2005, 6–8. doi:10.1093/JNCIMONO

GRAPHS/LGI006

Burrello, N., Calogero, A. E., Perdichizzi, A., Salmeri, M., D’Agata, R., and

Vicari, E. (2004). Inhibition of oocyte fertilization by assisted reproduc-

tive techniques and increased spermDNA fragmentation in the presence

of Candida albicans: a case report. Reprod. Biomed. Online 8, 569–573.

doi:10.1016/S1472-6483(10)61104-2

Burruel, V., Klooster, K. L., Chitwood, J., Ross, P. J., and Meyers, S. A.

(2013). Oxidative damage to rhesus macaque spermatozoa results in

mitotic arrest and transcript abundance changes in early embryos. Biol.

Reprod. 89, 72. doi:10.1095/BIOLREPROD.113.110981

Chabory, E., Damon, C., Lenoir, A., Kauselmann, G., Kern, H., Zevnik, B.,

Garrel, C., Saez, F., Cadet, R., Henry-Berger, J., Schoor, M., Gottwald,

U., Habenicht, U., Drevet, J. R., and Vernet, P. (2009). Epididymis

seleno-independent glutathione peroxidase 5 maintains sperm DNA

integrity in mice. J. Clin. Invest. 119, 2074–2085.

Chen, S. J., Allam, J. P., Duan, Y. G., and Haidl, G. (2013). Influence of

reactive oxygen species on human sperm functions and fertilizing

capacity including therapeutical approaches. Arch. Gynecol. Obstet.

288, 191–199. doi:10.1007/S00404-013-2801-4

De Iuliis, G. N., Newey, R. J., King, B. V., and Aitken, R. J. (2009a). Mobile

phone radiation induces reactive oxygen species production and

DNA damage in human spermatozoa in vitro. PLoS One 4, e6446.

doi:10.1371/JOURNAL.PONE.0006446

De Iuliis, G. N., Thomson, L. K., Mitchell, L. A., Finnie, J. M., Koppers, A.

J., Hedges, A., Nixon, B., and Aitken, R. J. (2009b). DNA damage in

human spermatozoa is highly correlatedwith the efficiency of chromatin

remodeling and the formation of 8-hydroxy-20-deoxyguanosine, amarker

of oxidative stress. Biol. Reprod. 81, 517–524. doi:10.1095/BIOLRE

PROD.109.076836

de Lamirande, E., and Gagnon, C. (1993a). Human sperm hyperactivation

and capacitation as parts of an oxidative process. Free Radic. Biol. Med.

14, 157–166. doi:10.1016/0891-5849(93)90006-G

deLamirande, E., andGagnon,C. (1993b). A positive role for the superoxide

anion in triggering hyperactivation and capacitation of human sperma-

tozoa. Int. J. Androl. 16, 21–25. doi:10.1111/J.1365-2605.1993.

TB01148.X

Delbes, G., Hales, B. F., andRobaire, B. (2010). Toxicants and human sperm

chromatin integrity. Mol. Hum. Reprod. 16, 14–22. doi:10.1093/

MOLEHR/GAP087

Dona, G., Fiore, C., Andrisani, A., Ambrosini, G., Brunati, A., Ragazzi, E.,

Armanini, D., Bordin, L., and Clari, G. (2011). Evaluation of correct

endogenous reactive oxygen species content for human sperm capacita-

tion and involvement of the NADPH oxidase system. Hum. Reprod. 26,

3264–3273. doi:10.1093/HUMREP/DER321

du Plessis, S. S., Agarwal, A., Mohanty, G., and van der Linde, M. (2015).

Oxidative phosphorylation versus glycolysis: what fuel do spermatozoa

use? Asian J. Androl. 17, 230–235. doi:10.4103/1008-682X.135123

Erkekoglu, P., Rachidi, W., Yuzugullu, O. G., Giray, B., Favier, A., Ozturk,

M., andHincal, F. (2010). Evaluation of cytotoxicity and oxidativeDNA

damaging effects of di(2-ethylhexyl)-phthalate (DEHP) and mono

(2-ethylhexyl)-phthalate (MEHP) onMA-10Leydig cells and protection

by selenium. Toxicol. Appl. Pharmacol. 248, 52–62. doi:10.1016/

J.TAAP.2010.07.016

Evans, T. C. (1947). Effects of hydrogen peroxide produced in the medium

by radiation of spermatozoa of Arbacia punctulata. Biol. Bull. 92,

99–109. doi:10.2307/1538160

Evenson, D. P., andWixon, R. (2005). Environmental toxicants cause sperm

DNA fragmentation as detected by the sperm chromatin structure assay

(SCSA). Toxicol. Appl. Pharmacol. 207(Suppl.), 532–537. doi:10.1016/

J.TAAP.2005.03.021

Fariello, R. M., Pariz, J. R., Spaine, D. M., Cedenho, A. P., Bertolla, R. P.,

and Fraietta, R. (2012). Association between obesity and alteration of

spermDNA integrity andmitochondrial activity.BJU Int. 110, 863–867.

doi:10.1111/J.1464-410X.2011.10813.X

Fraga, C. G.,Motchnik, P. A., Shigenaga,M.K., Helbock,H. J., Jacob,R. A.,

and Ames, B. N. (1991). Ascorbic acid protects against endogenous

oxidative DNA damage in human sperm. Proc. Natl Acad. Sci. USA

88, 11 003–11 006. doi:10.1073/PNAS.88.24.11003

Fraga, C. G., Motchnik, P. A., Wyrobek, A. J., Rempel, D. M., and Ames,

B. N. (1996). Smoking and low antioxidant levels increase oxidative

damage to sperm DNA. Mutat. Res. 351, 199–203. doi:10.1016/0027-

5107(95)00251-0

Fujita, Y., Mihara, T., Okazaki, T., Shitanaka, M., Kushino, R., Ikeda, C.,

Negishi, H., Liu, Z., Richards, J. S., and Shimada, M. (2011). Toll-like

receptors (TLR) 2 and 4 on human sperm recognize bacterial endotoxins

and mediate apoptosis. Hum. Reprod. 26, 2799–2806. doi:10.1093/

HUMREP/DER234

Ghani, E., Keshtgar, S., Habibagahi, M., Ghannadi, A., and Kazeroni, M.

(2013). Expression of NOX5 in human teratozoospermia compared to

normozoospermia. Andrologia 45, 351–356. doi:10.1111/AND.12023

Gharagozloo, P., and Aitken, R. J. (2011). The role of sperm oxidative stress

in male infertility and the significance of oral antioxidant therapy.Hum.

Reprod. 26, 1628–1640. doi:10.1093/HUMREP/DER132

Ghosh, D., Das, U. B., and Misro, M. (2002). Protective role of alpha-

tocopherol–succinate (provitamin-E) in cyclophosphamide induced tes-

ticular gametogenic and steroidogenic disorders: a correlative approach

to oxidative stress. Free Radic. Res. 36, 1209–1218. doi:10.1080/

1071576021000016472

Gibb, Z., Lambourne, S. R., and Aitken, R. J. (2014). The paradoxical

relationship between stallion fertility and oxidative stress. Biol. Reprod.

91, 77. doi:10.1095/BIOLREPROD.114.118539

Grizard, G., Ouchchane, L., Roddier, H., Artonne, C., Sion, B., Vasson,M. P.,

and Janny, L. (2007). In vitro alachlor effects on reactive oxygen species

generation, motility patterns and apoptosis markers in human spermato-

zoa.Reprod. Toxicol.23, 55–62. doi:10.1016/J.REPROTOX.2006.08.007

Herrero, M. B., de Lamirande, E., and Gagnon, C. (2001). Tyrosine nitration

in human spermatozoa: a physiological function of peroxynitrite, the

reaction product of nitric oxide and superoxide. Mol. Hum. Reprod. 7,

913–921. doi:10.1093/MOLEHR/7.10.913

Hosken, D. J., and Hodgson, D. J. (2014). Why do sperm carry RNA?

Relatedness, conflict, and control. Trends Ecol. Evol. 29, 451–455.

doi:10.1016/J.TREE.2014.05.006

8 Reproduction, Fertility and Development R. J. Aitken et al.

Page 10: Aitken 2016 causes and consequences of oxidative stress in spermatozoa

Houston, B., Curry, B., and Aitken, R. J. (2015). Human spermatozoa

possess an IL4I1 L-amino acid oxidase with a potential role in sperm

function. Reproduction 149, 587–596. doi:10.1530/REP-14-0621

Hull, M. G. R., Glazener, C. M. A., Kelly, N. J., Conway, D. I., Foster, P. A.,

Hunton, R. A., Coulson, C., Lambert, P. A., Watt, E. M., and Desai,

K. M. (1985). Population study of causes, treatment and outcome of

infertility. Br. Med. J. (Clin. Res. Ed.) 291, 1693–1697. doi:10.1136/

BMJ.291.6510.1693

Irvine,D. S., Twigg, J. P., Gordon, E. L., Fulton,N.,Milne, P.A., andAitken,

R. J. (2000). DNA integrity in human spermatozoa: relationships with

semen quality. J. Androl. 21, 33–44.

Jones, R.,Mann, T., and Sherins, R. J. (1978).Adverse effects of peroxidized

lipid on human spermatozoa. Proc. R. Soc. Lond. B Biol. Sci. 201,

413–417. doi:10.1098/RSPB.1978.0053

Jones, R., Mann, T., and Sherins, R. J. (1979). Peroxidative breakdown of

phospholipids in human spermatozoa: spermicidal effects of fatty acid

peroxides and protective action of seminal plasma. Fertil. Steril. 31,

531–537.

Kao, S.H., Chao, H. T., Chen, H.W., Hwang, T. I., Liao, T. L., andWei, Y. H.

(2008). Increase of oxidative stress in human sperm with lower motility.

Fertil. Steril. 89, 1183–1190. doi:10.1016/J.FERTNSTERT.2007.05.029

Katen, A. L., andRoman, S. D. (2015). The genetic consequences of paternal

acrylamide exposure and potential for amelioration. Mutat. Res. 777,

91–100. doi:10.1016/J.MRFMMM.2015.04.008

Koppers, A. J., De Iuliis, G.N., Finnie, J.M.,McLaughlin, E. A., andAitken,

R. J. (2008). Significance of mitochondrial reactive oxygen species in

the generation of oxidative stress in spermatozoa. J. Clin. Endocrinol.

Metab. 93, 3199–3207. doi:10.1210/JC.2007-2616

Koppers, A. J., Mitchell, L. A., Wang, P., Lin, M., and Aitken, R. J. (2011).

Phosphoinositide 3-kinase signalling pathway involvement in a

truncated apoptotic cascade associated with motility loss and oxidative

DNA damage in human spermatozoa. Biochem. J. 436, 687–698.

doi:10.1042/BJ20110114

Lagos-Cabre, R., and Moreno, R. D. (2012). Contribution of environmental

pollutants to male infertility: a working model of germ cell apoptosis

induced by plasticizers. Biol. Res. 45, 5–14. doi:10.4067/S0716-

97602012000100001

Lane, M., McPherson, N. O., Fullston, T., Spillane, M., Sandeman, L.,

Kang, W. X., and Zander-Fox, D. L. (2014). Oxidative stress in mouse

sperm impairs embryo development, fetal growth and alters adiposity

and glucose regulation in female offspring. PLoS One 9, e100832.

doi:10.1371/JOURNAL.PONE.0100832

Liang, R., Senturker, S., Shi, X., Bal, W., Dizdaroglu, M., and Kasprzak,

K. S. (1999). Effects of Ni(II) and Cu(II) on DNA interaction with the

N-terminal sequence of human protamine P2: enhancement of binding

and mediation of oxidative DNA strand scission and base damage.

Carcinogenesis 20, 893–898. doi:10.1093/CARCIN/20.5.893

Lundbaek, J. A., and Andersen, O. S. (1994). Lysophospholipids modulate

channel function by altering the mechanical properties of lipid bilayers.

J. Gen. Physiol. 104, 645–673. doi:10.1085/JGP.104.4.645

MacLeod, J. (1943). The role of oxygen in the metabolism and motility of

human spermatozoa. Am. J. Physiol. 138, 512–518.

Metzler-Guillemain,C.,Victorero,G., Lepoivre,C.,Bergon,A.,Yammine,M.,

Perrin, J., Sari-Minodier, I., Boulanger, N., Rihet, P., and Nguyen, C.

(2015). Sperm mRNAs and microRNAs as candidate markers for the

impact of toxicants on human spermatogenesis: an application to tobacco

smoking. Syst Biol Reprod Med 61, 139–149. doi:10.3109/19396368.

2015.1022835

Moazamian, R., Polhemus, A., Connaughton, H., Fraser, B., Whiting, S.,

Gharagozloo, P., and Aitken, R. J. (2015). Oxidative stress and human

spermatozoa: diagnostic and functional significance of aldehydes gen-

erated as a result of lipid peroxidation.Mol. Hum. Reprod. 21, 502–515.

doi:10.1093/MOLEHR/GAV014

Morielli, T., and O’Flaherty, C. (2015). Oxidative stress impairs function

and increases redox protein modifications in human spermatozoa.

Reproduction 149, 113–123. doi:10.1530/REP-14-0240

Morimoto, H., Iwata, K., Ogonuki, N., Inoue, K., Atsuo, O., Kanatsu-

Shinohara, M., Morimoto, T., Yabe-Nishimura, C., and Shinohara, T.

(2013). ROS are required for mouse spermatogonial stem cell self-

renewal.Cell Stem Cell 12, 774–786. doi:10.1016/J.STEM.2013.04.001

Muratori, M., Tamburrino, L., Marchiani, S., Cambi, M., Olivito, B.,

Azzari, C., Forti, G., and Baldi, E. (2015). Investigation on the origin

of sperm DNA fragmentation: role of apoptosis, immaturity and oxida-

tive stress.Mol. Med. 21, 109–122. doi:10.2119/MOLMED.2014.00158

Musset, B., Clark,R.A.,DeCoursey, T.E., Petheo,G.L.,Geiszt,M., Chen,Y.,

Cornell, J. E., Eddy, C.A., Brzyski,R.G., andEl Jamali,A. (2012).NOX5

in human spermatozoa: expression, function, and regulation. J. Biol.

Chem. 287, 9376–9388. doi:10.1074/JBC.M111.314955

Nakamura, H., Kimura, T., Nakajima, A., Shimoya, K., Takemura, M.,

Hashimoto, K., Isaka, S., Azuma, C., Koyama, M., and Murata, Y.

(2002). Detection of oxidative stress in seminal plasma and fractionated

sperm from subfertile male patients. Eur. J. Obstet. Gynecol. Reprod.

Biol. 105, 155–160. doi:10.1016/S0301-2115(02)00194-X

Nishikawa, T., Tomori, Y., Yamashita, S., and Shimizu, S. (1989). Inhibition

of Naþ,Kþ-ATPase activity by phospholipase A2 and several lysopho-

spholipids: possible role of phospholipase A2 in noradrenaline release

from cerebral cortical synaptosomes. J. Pharm. Pharmacol. 41, 450–458.

doi:10.1111/J.2042-7158.1989.TB06499.X

Noblanc,A.,Damon-Soubeyrand,C., Karrich, B., Henry-Berger, J., Cadet,R.,

Saez, F., Guiton, R., Janny, L., Pons-Rejraji, H., Alvarez, J. G., Jr,

Drevet, J. R., and Kocer, A. (2013). DNA oxidative damage in mamma-

lian spermatozoa: where and why the male nucleus is impacted? Free

Radic. Biol. Med. 65, 719–723. doi:10.1016/J.FREERADBIOMED.

2013.07.044

O’Flaherty, C. (2014). Iatrogenic genetic damage of spermatozoa. Adv. Exp.

Med. Biol. 791, 117–135. doi:10.1007/978-1-4614-7783-9_8

Ohno,M., Sakumi,K., Fukumura, R., Furuichi,M., Iwasaki,Y.,Hokama,M.,

Ikemura, T., Tsuzuki, T., Gondo, Y., and Nakabeppu, Y. (2014).

8-Oxoguanine causes spontaneous de novo germline mutations in mice.

Sci. Rep. 4, 4689. doi:10.1038/SREP04689

Ostermeier, G. C., Goodrich, R. J., Moldenhauer, J. S., Diamond, M. P., and

Krawetz, S. A. (2005). A suite of novel human spermatozoal RNAs.

J. Androl. 26, 70–74.

Palmer, N. O., Bakos, H. W., Owens, J. A., Setchell, B. P., and Lane, M.

(2012). Diet and exercise in an obese mouse fed a high-fat diet improve

metabolic health and reverse perturbed sperm function. Am. J. Physiol.

Endocrinol. Metab. 302, E768–E780. doi:10.1152/AJPENDO.00401.

2011

Prescott, J., Du,M.,Wong, J. Y., Han, J., andDeVivo, I. (2012). Paternal age

at birth is associated with offspring leukocyte telomere length in the

Nurses’ Health Study. Hum. Reprod. 27, 3622–3631. doi:10.1093/

HUMREP/DES314

Reichart, M., Kahane, I., and Bartoov, B. (2000). In vivo and in vitro

impairment of human and ram sperm nuclear chromatin integrity by

sexually transmitted Ureaplasma urealyticum infection. Biol. Reprod.

63, 1041–1048. doi:10.1095/BIOLREPROD63.4.1041

Rivlin, J., Mendel, J., Rubinstein, S., Etkovitz, N., and Breitbart, H. (2004).

Role of hydrogen peroxide in sperm capacitation and acrosome reaction.

Biol. Reprod. 70, 518–522. doi:10.1095/BIOLREPROD.103.020487

Rodriguez, P. C., and Beconi, M. T. (2009). Peroxynitrite participates in

mechanisms involved in capacitation of cryopreserved cattle. Anim.

Reprod. Sci. 110, 96–107. doi:10.1016/J.ANIREPROSCI.2007.12.017

Sakamoto, Y., Ishikawa, T., Kondo, Y., Yamaguchi, K., and Fujisawa, M.

(2008). The assessment of oxidative stress in infertile patients with

varicocele. BJU Int. 101, 1547–1552. doi:10.1111/J.1464-410X.2008.

07517.X

Oxidative stress in spermatozoa Reproduction, Fertility and Development 9

Page 11: Aitken 2016 causes and consequences of oxidative stress in spermatozoa

Sanocka, D., Miesel, R., Jedrzejczak, P., and Kurpisz, M. K. (1996).

Oxidative stress and male infertility. J. Androl. 17, 449–454.

Santiso,R., Tamayo,M.,Gosalvez, J., Johnston, S.,Marino,A., Fernandez,C.,

Losada, C., and Fernandez, J. L. (2012). DNA fragmentation

dynamics allows the assessment of cryptic sperm damage in human:

evaluation of exposure to ionizing radiation, hyperthermia, acidic pH

and nitric oxide. Mutat. Res. 734, 41–49. doi:10.1016/J.MRFMMM.

2012.03.006

Sharma, R. K., and Agarwal, A. (1996). Role of reactive oxygen species in

male infertility. Urology 48, 835–850. doi:10.1016/S0090-4295(96)

00313-5

Shen, H., and Ong, C. (2000). Detection of oxidative DNA damage in

human sperm and its associationwith sperm function andmale infertility.

Free Radic. Biol. Med. 28, 529–536. doi:10.1016/S0891-5849(99)

00234-8

Showell, M. G., Mackenzie-Proctor, R., Brown, J., Yazdani, A., Stankiewicz,

M. T., andHart, R. J. (2014). Antioxidants for male subfertility.Cochrane

Database Syst. Rev. 12, CD007411.

Shukla, K. K., Mahdi, A. A., and Rajender, S. (2012). Apoptosis, spermato-

genesis and male infertility. Front. Biosci. (Elite Ed.) E4, 746–754.

doi:10.2741/E415

Sibirtsev, J. T., Shastina, V. V., Menzorova, N. I., Makarieva, T. N., and

Rasskazov, V. (2011). A Ca2þ, Mg2þ-dependent DNase involvement in

apoptotic effects in spermatozoa of sea urchin Strongylocentrotus

intermedius induced by two-headed sphingolipid, rhizochalin. Mar.

Biotechnol. (NY) 13, 536–543. doi:10.1007/S10126-010-9324-9

Simoes, R., Feitosa, W. B., Siqueira, A. F., Nichi, M., Paula-Lopes, F. F.,

Marques, M. G., Peres, M. A., Barnabe, V. H., Visintin, J. A., and

Assumpcao, M. E. (2013). Influence of bovine sperm DNA fragmenta-

tion and oxidative stress on early embryo in vitro development outcome.

Reproduction 146, 433–441. doi:10.1530/REP-13-0123

Singh, N. P., and Stephens, R. E. (1998). X-Ray induced DNA double-

strand breaks in human sperm. Mutagenesis 13, 75–79. doi:10.1093/

MUTAGE/13.1.75

Singh, N. P., Muller, C. H., and Berger, R. E. (2003). Effects of age on DNA

double-strand breaks and apoptosis in human sperm. Fertil. Steril. 80,

1420–1430. doi:10.1016/J.FERTNSTERT.2003.04.002

Smith, T. B., De Iuliis, G. N., Lord, T., and Aitken, R. J. (2013a). The

senescence-accelerated mouse prone 8 as a model for oxidative stress

and impaired DNA repair in the male germ line. Reproduction 146,

253–262. doi:10.1530/REP-13-0186

Smith, T. B., Dun,M. D., Smith, N. D., Curry, B. J., Connaughton, H. S., and

Aitken, R. J. (2013b). The presence of a truncated base excision repair

pathway in human spermatozoa that is mediated by OGG1. J. Cell Sci.

126, 1488–1497. doi:10.1242/JCS.121657

Sotolongo, B., Huang, T. T., Isenberger, E., and Ward, W. S. (2005). An

endogenous nuclease in hamster, mouse, and human spermatozoa

cleaves DNA into loop-sized fragments. J. Androl. 26, 272–280.

Soubry, A. (2015). Epigenetic inheritance and evolution: a paternal per-

spective on dietary influences. Prog. Biophys. Mol. Biol. 118, 79–85.

doi:10.1016/J.PBIOMOLBIO.2015.02.008

Tosic, J., and Walton, A. (1946). Formation of hydrogen peroxide by

spermatozoa and its inhibitory effect on respiration. Nature 158, 485.

doi:10.1038/158485A0

Tosic, J., andWalton, A. (1950).Metabolism of spermatozoa. The formation

and elimination of hydrogen peroxide by spermatozoa and effects on

motility and survival. Biochem. J. 47, 199–212.

van Kuijk, F. J., Handelman, G. J., and Dratz, E. A. (1985). Consecutive

action of phospholipase A2 and glutathione peroxidase is required for

reduction of phospholipid hydroperoxides and provides a convenient

method to determine peroxide values in membranes. J. Free Radic. Biol.

Med. 1, 421–427. doi:10.1016/0748-5514(85)90156-4

Vernet, P., Fulton, N., Wallace, C., and Aitken, R. J. (2001). Analysis of

reactive oxygen species generating systems in rat epididymal spermato-

zoa. Biol. Reprod. 65, 1102–1113. doi:10.1095/BIOLREPROD65.4.1102

Wang, X., Sharma, R. K., Sikka, S. C., Thomas, A. J., Jr, Falcone, T., and

Agarwal, A. (2003). Oxidative stress is associated with increased

apoptosis leading to spermatozoa DNA damage in patients with male

factor infertility. Fertil. Steril. 80, 531–535. doi:10.1016/S0015-0282

(03)00756-8

Weir, C. P., and Robaire, B. (2007). Spermatozoa have decreased antioxi-

dant enzymatic capacity and increased reactive oxygen species produc-

tion during aging in the Brown Norway rat. J. Androl. 28, 229–240.

doi:10.2164/JANDROL.106.001362

Zalata, A., Hafez, T., Mahmoud, A., and Comhaire, F. (1995). Relationship

between resazurin reduction test, reactive oxygen species generation,

and gamma-glutamyltransferase. Hum. Reprod. 10, 1136–1140.

Zhou, D., Wang, H., Zhang, J., Gao, X., Zhao, W., and Zheng, Y. (2010).

Di-n-butyl phthalate (DBP) exposure induces oxidative damage in testes

of adult rats. Syst. Biol. Reprod. Med. 56, 413–419. doi:10.3109/

19396368.2010.509902

Zribi, N., Chakroun, N. F., Ben Abdallah, F., Elleuch, H., Sellami, A.,

Gargouri, J., Rebai, T., Fakhfakh, F., and Keskes, L. A. (2012). Effect of

freezing–thawing process and quercetin on human sperm survival and

DNA integrity. Cryobiology 65, 326–331. doi:10.1016/J.CRYOBIOL.

2012.09.003

www.publish.csiro.au/journals/rfd

10 Reproduction, Fertility and Development R. J. Aitken et al.

View publication statsView publication stats