Abstract

Mammalian oocytes begin meiosis in the fetal ovary, but only complete it when fertilized in the adult reproductive tract. This review examines the cell biology of this protracted process: from entry of primordial germ cells into meiosis to conception. The defining feature of meiosis is two consecutive cell divisions (meiosis I and II) and two cell cycle arrests: at the germinal vesicle (GV), dictyate stage of prophase I and at metaphase II. These arrests are spanned by three key events, the focus of this review: (i) passage from mitosis to GV arrest during fetal life, regulated by retinoic acid; (ii) passage through meiosis I and (iii) completion of meiosis II following fertilization, both meiotic divisions being regulated by cyclin-dependent kinase (CDK1) activity. Meiosis I in human oocytes is associated with an age-related high rate of chromosomal mis-segregation, such as trisomy 21 (Down’s syndrome), resulting in aneuploid conceptuses. Although aneuploidy is likely to be multifactorial, oocytes from older women may be predisposed to be becoming aneuploid as a consequence of an age-long decline in the cohesive ties holding chromosomes together. Such loss goes undetected by the oocyte during meiosis I either because its ability to respond and block division also deteriorates with age, or as a consequence of being inherently unable to respond to the types of segregation defects induced by cohesion loss.

Introduction

The main objective of this review is to provide an account of the events in the life of an oocyte, from its creation in the fetal ovary to eventual fusion with a spermatozoon following ovulation. The review aims to focus on the etiology of aneuploidy, given this process has such important clinical implications.

Defining meiosis

The defining feature of meiosis, in females as well as males alike, is that chromosome replication in S-phase is followed by two consecutive cell divisions to produce haploid gametes (see life cycle of oocyte in Fig. 1). The replicated pairs of chromosomes, known as homologs, recombine allowing exchange of genetic information between sister chromatids of originally paternally and maternally-derived homologs before the first meiotic division (Coop and Przeworski, 2007). Paradoxically, the large numbers of spermatozoa produced in humans are aneuploid at a rate of 1–2%, which is at least 10-fold lower than that calculated for oocytes (Hassold and Hunt, 2001; Pacchierotti et al., 2007) whose number is most probably limited to those established in fetal life. The creation of oocytes in adult life from hemopoietic stem cells appears possible (Johnson et al., 2004, 2005), however, they have been reported not to contribute to any of the cohort of mature ovulated oocytes (Eggan et al., 2006); therefore their function remains enigmatic and is not discussed further. Other reviews have concentrated on defining the incidence and exact nature of aneuploidy in human oocytes (Rosenbusch, 2004; Delhanty, 2005; Pellestor et al., 2005, 2006) and these aspects are therefore not discussed in detail here.

Figure 1:

The life of a mammalian oocyte

The figures depicts the life of an oocyte beginning with its inception from oogonia following PGC colonization of the ovary (left) to pronucleus formation following fertilization (right), which marks the completion of meiosis and entry into the first embryonic cell cycle. See text for details

Abnormalities which result in numerical changes to the 46+XY chromosome number in humans are common and may be caused by non-disjunction of chromosomes during either of the two meiotic divisions or by premature separation of sister chromatids (so called pre-division) (Fig. 2); although their relative contribution to human aneuploidy is still debated (Wolstenholme and Angell, 2000; Rosenbusch, 2004; Pellestor et al., 2005). It is however likely that pre-division is associated more with the age-related incidence of aneuploidy (Angell, 1991; Pellestor et al., 2003; Vialard et al., 2006, 2007).

Figure 2:

Aneuploidy in oocytes during meiosis I

A homolog pair, aligned and under tension on a metaphase I spindle: (A, C, D and E) these pairs have a single crossover event, i.e. located interstitial (A); distal (C) or proximal (D and E) in relation to the kinetochore (sister kinetochores are drawn as one for simplicity). (B) Homologs are achiasmate and so segregate independently. (E) Sister chromatids segregate during meiosis I (pre-division) due to both kinetochores of the sisters being attached to poles. In (A) the homologs segregate reductionally; in (D) the proximal exchange has prevented homolog segregation; and in (C) the pair has undergone premature separation due to the weak association of homologs

Aneuploidy and reproduction

Aneuploidy is a major obstacle in achieving reproductive success. About 20% of all human oocytes are regarded as being aneuploid; although this figure can vary widely, from ∼10% to as high as ∼40–60% (Kuliev et al., 2005; Rosenbusch and Schneider, 2006; Pacchierotti et al., 2007), and is thought to rise with increasing maternal age. It is important to consider the difficulty in getting a true reflection of aneuploidy rates in the general population, and how this may change with maternal age, given the data available are from a very selective population, failed-to-fertilize oocytes from female patients undergoing IVF (Pacchierotti et al., 2007). The vast majority of embryos formed from aneuploid oocytes are non-viable and are lost at some point before term. Trisomy 21 (Down syndrome, DS) is one of the few autosomal aneuploidies to produce viable embryos, but even here it is estimated that 80% of conceptuses are spontaneously aborted (Hassold and Jacobs, 1984). The increased use of various types of antenatal screening has been shown to detect and reduce the live DS birth rate in a number of countries despite increasing trends for older women to have children (Lai et al., 2002; Egan et al., 2004; Khoshnood et al., 2004a, b; Sipek et al., 2004; Jou et al., 2005; O’Leary et al., 2006; Coory et al., 2007), although in some reports the increased detection by screening appears to have been offset by the increased age-related incidence (Bell et al., 2003).

The success of antenatal screening and the overall incidence of DS in live births, ∼1:1000 (Morris et al., 2002; Egan et al., 2004), belies the large spectrum of clinical symptoms associated with the syndrome that have major influence on the health and welfare of affected individuals, their carers and on health service provision. Furthermore, with the increasing global first world trend to have children in later life coupled to the increased incidence of DS and many aneuploidies with maternal age (Delhanty, 2005; Pellestor et al., 2005; Sherman et al., 2005), it is likely to remain an important health issue over the next few decades.

The ‘two-hit’ hypothesis of aneuploidy

Current thinking on the derivation of aneuploidy in oocytes is dominated by the two-hit hypothesis (Lamb et al., 1996; Warren and Gorringe, 2006). These two ‘hits’ both need to occur in order to produce mis-segregated chromosomes. The first hit is a non-recombined or poorly recombined homolog pair (Lamb et al., 1996, 1997), although the derivation of this is discussed later. The second hit is an inability of the oocyte to detect and respond to this recombination failure, and in so doing allow chromosomes to mis-segregate during their meiotic division (Hawley et al., 1994). However, this second hit could in fact be the accumulation of various influences with maternal age, and these together, when summated, constitute the second hit.

It is important to consider the time interval between these two events; the first hit happens in fetal life during recombination, whereas the second hit is in the adult, in a periovulatory period during either the first or second meiotic divisions (MI and MetII, respectively). The long time interval allows the rate of aneuploidy to be affected by maternal age; theoretically by exposure to extrinsic, environmental insults (e.g. smoking; health; pollutants, etc…) which would act to accelerate the intrinsic process of aging accrued through cellular damage, e.g. by reactive oxygen species (Eichenlaub-Ritter et al., 2004). One attractive idea therefore is that such accrued damage in some way makes oocytes less able to sense recombination errors that arose in prophase during fetal life. However, it is also possible that oocytes remain poor at this detection throughout life, and instead it is the ties which hold chromosomes together that simply get weaker with age, damaged by aging and unable to be properly repaired (Wolstenholme and Angell, 2000).

Passage from mitosis to GV stage arrest

From PGC to oocyte

The tenet of the ‘two-hit’ model is that fetal recombination errors constitute the first hit. Therefore it is important to discuss how primordial germ cells (PGCs) undergo entry into meiosis, and this will be done in mice where detail is most known. PGCs appear not to enter meiosis completely synchronously and such findings may have implications for the quality of oocytes produced as is discussed later.

In mouse fetal life at 7 days post-coitum (days pc), ∼45 PGCs (Lawson and Hage, 1994) migrate from the base of the allantois to a structure that becomes the genital ridge (McLaren, 2003; Wilhelm et al., 2007). Here the now 3000 PGCs, termed gonocytes, continue to divide mitotically until 13.5 days pc (Monk and McLaren, 1981; McLaren, 2000), when in females they commit to enter meiosis. Retinoic acid (RA) produced by the associated mesenopheros appears to be critical in inducing entry into meiosis of female gonocytes (oogonia) (Bowles et al., 2006; Koubova et al., 2006); however, PGCs on entering the genital ridge are initially unable to respond to the RA signal because of the action of the cyp26b1 gene, which produces a RA-metabolizing P450 cytochrome enzyme (Bowles et al., 2006). CYP26B1 expression falls off in oogonia, such that by 12.5 days pc its expression is male-specific. Loss of cyp26b1 expression in females between 10–12.5 days pc, allows RA to stimulate oogonia commitment to meiosis. One important target gene for RA is Stimulated by Retinoic Acid gene 8 (Stra8), since in Stra8 knockouts, oogonia fail to undergo any of the critical events of entry into the pre-meiotic S-phase which eventually lead to recombination (Baltus et al., 2006). An important feature of RA-induced entry into meiosis is the source of the RA, which is located in the associated mesenopheros (Bowles et al., 2006), this would form a gradient of RA within the developing gonad. Such a gradient probably underlies the observed spatial wave, anterior to posterior, of meiosis entry in mouse fetal ovaries, judged by the switch off of germ cell markers such as Oct-4, and the switch on of meiotic markers such as Stra8, Dmc1 and Sycp3 (Menke et al., 2003; Bullejos and Koopman, 2004). A similar wave of meiosis entry probably also occurs in the human ovary, judged by the finding that oogonia committing to leptotene first at around 9 week pc are closest to the mesonephros, although oogenesis appears less strictly timed (Bendsen et al., 2006).

Meiotic prophase: from leptotene to dictyate-arrest

The process of meiotic recombination in oocytes begins on 13–14 days pc in mouse (Speed, 1982) and in the 10–11th week of gestation in humans (Kurilo, 1981; Garcia et al., 1987). During leptotene, homolog pairs align and then synapse with the formation of the synaptonemal complex during zygotene (Zickler and Kleckner, 1998; Walker and Hawley, 2000; Page and Hawley, 2004). The synaptonemal complex marks the site of crossovers and probably provides the structural framework in which to execute double strand breaks. Double strand breaks are initiated by the topoisomerase-like enzyme Spo11 (Keeney and Neale, 2006), whose loss in mouse blocks this process leading to sterile animals (Baudat et al., 2000; Romanienko and Camerini-Otero, 2000) and whose mutation may be responsible for some infertility in humans (Mori et al., 2006). The proteins involved in synapsis, crossover exchange and synaptonemal complex formation are partly known in mammals, and such identification is often aided by having yeast homologs. Invariably, knockout of these genes produces sterile mice (Pittman et al., 1998; Yuan et al., 2002; Li and Schimenti, 2007). The creation of double strand breaks requires oocytes to have a mechanism for a meiosis-specific DNA break repair which uses as template the homolog rather than the sister strand (Kleckner, 2006). It also necessitates a surveillance mechanism which monitors for unrepaired double strand breaks at a time when recombination should have been completed. Defects in genes (eg dmc1 and Msh5) involved in the processing of meiotic double strand breaks produces a severe loss of oocytes shortly before and after birth (Di Giacomo et al., 2005; Morelli and Cohen, 2005).

The synaptonemal complex dissolves leaving the physical manifestations at points where crossover has occurred, chiasmata, visible in diplotene. The oocyte then arrests with decondensed chromatin at this stage, often referred to as dictyate-stage arrest. In mouse, most oocytes have reached the dictyate stage by 20 days pc (Speed, 1982), while it is less synchronous in humans with dictyate oocytes observed from 14–15 weeks pc onwards (Kurilo, 1981; Garcia et al., 1987). In mouse, entry into meiosis is not synchronous: oogonia first enter meiosis at ∼13.5 days pc and this process continues in a wave which is nearly complete about 3 days later (Peters, 1970; McLaren, 2000). Parturition is at ∼20 dcp and oocytes have completed prophase by Day 1 post-partum (Speed, 1982). In humans, the passage of oogonia into meiosis appears far more spread during gestation. Here leptotene stage oocytes can first be detected at ∼10 weeks and these oocytes appear to take around 5 weeks to reach the dictyate stage (Kurilo, 1981; Garcia et al., 1987). However, oocytes continue to be observed at all stage of meiosis throughout gestation.

Entry into meiotic prophase and aneuploidy

It is important to note that crossover happens for each homolog pair, even the X and Y chromosomes in spermatogenesis, which pair through their pseudoautosomal regions (Graves et al., 1998; Blaschke and Rappold, 2006). The crossovers, in conjunction with sister chromatid cohesion, allow homologs to stay paired when pulling forces from microtubules are generated during MI (Fig. 2A). A failure to recombine means that homologs cannot remain associated together during MI, they therefore segregate randomly (Fig. 2B). The generation of achiasmate homologs is likely to account for around 50% of all DS conceptuses whose origin is in a non-disjunction during MI (Lamb et al., 1997). If the exchange is near the end of the chromosomes it follows that during MI it would only be the small distal exchange regions that could generate the counter-tension from microtubule pulling forces. A priori, exchange at sites near the telomeres of chromosomes should generate recombined homologs which are sensitive to non-disjunction (Fig. 2C). This has indeed been borne out by studies in yeast artificial chromosomes (Ross et al., 1996) and flies (Koehler et al., 1996). Exchange in a pericentromeric location also appears to generate a configuration susceptible to non-disjunction (Lamb et al., 1997; Hassold et al., 2000). Here the pericentromeric crossover may interfere with centromeric cohesion and/or kinetochore attachment to allow both homologs to move to the same pole (Fig. 2D). Or pericentromeric crossover may interfere with the essential block in MI to sister kinetochrores being attached to both spindle poles, which therefore allows sister chromatids to be separated in MI, rather than during metaphase II, MetII (so called ‘pre-division’; Fig. 2E).

In the two-hit model, susceptible patterns of recombination forming the first hit can be generated stochastically during prophase because it is the second hit which is believed to be age related. However, in mouse it has long been observed that chiasmata frequency in oocytes decreases with maternal age (Henderson and Edwards, 1968; Luthardt et al., 1973; Polani and Jagiello, 1976). Since the recombination pattern of homologs is established in fetal life, Henderson and Edwards (1968), who were the first to identify this phenomenon, suggested that such observations were explained by fetal oocytes entering meiosis in an order which mirrored their exit in the adult. Oocytes first to enter meiosis are those that mature first in the adult so resembling a production line. The further proposition being that the late entry of oocytes into meiosis negatively affects their ability to recombine.

Because the molecular details of sister attachment is now known, it is probable that the visible loss of chiasmata observed previously is likely due to an age-associated loss of sister chromatid cohesion (Fig. 3). Therefore no production line has to be invoked to account for the age-related decrease in the incidence of chiasmata. However, it remains an interesting unresolved question as to whether the oocytes first to commit to meiosis in the fetus are the first to be ovulated in the adult. It seems clear that oocytes do enter meiosis in a wave (as detailed above), however, is that wave similarly mirrored in ovulations during adult life? Experiments designed directly to demonstrate the phenomenon of ‘first in–first out’ have not been extensive and unfortunately remain equivocal (Polani and Crolla, 1991; Hirshfield, 1992; Meredith and Doolin, 1997). An in vitro labeling of pre-meiotic S-phase mouse oogonia by radiolabeled nucleotide, followed by transplantation, demonstrated that early labeled oogonia go on to mature early (Polani and Crolla, 1991). Complementary findings in the rat using an in vivo method of delivery demonstrated that late labeled oogonia were not observed in growing follicles of adult rats and instead formed a pool of the most quiescent primordial follicles (Hirshfield, 1992). However, a similar approach again in rat failed to verify a strict production line with early entry oocytes being found in growing follicles at all ages (Meredith and Doolin, 1997), but it did show late entry late entry oocytes were more quiescent than those with early entry.

Figure 3:

Loss of cohesin with age resolves chiasmata

A single crossover event has happened in fetal life for this homolog pair (A). With age the cohesive ties holding homologs weaken (B). This distal exchange has been affected by the loss of distal cohesin such that the homolog pair are achiasmate and will segregate randomly at meiosis I

For the production line hypothesis to be important here, in the derivation of aneuploidy, then one would have to assume the process of meiosis is hindered somehow by late entry. Interestingly, it is of note that late entry oogonia would have likely undergone more replications than the early oogonia. During this replication period one would anticipate cells maintain high telomerase activity; the enzyme responsible for maintaining adequate telomere length to the end of chromosomes. Indeed, telomerase activity is very high in the fetal ovary (Wright et al., 1996). Defective telomeres, however, in yeasts, reduce recombination rates during meiosis (Cooper et al., 1998; Nimmo et al., 1998; Rockmill and Roeder, 1998) and it has therefore been suggested that telomere shortening in late entry oogonia may be a determinant in aneuploidy (Keefe et al., 2006, 2007). It remains to be established if telomere shortening is observed in oocytes from older women.

The cohesin complex

It is important to consider the ties which hold chromosomes together. These ties if perturbed could potentially account for mis-segregation. As it will be discussed, they are set up in fetal life during prophase and as such offer a long-standing target for the generation of aneuploidy by non-disjunction. First, to understand the process in meiosis we have to examine how sisters are held together during mitosis. One important feature of mitosis is that to ensure fidelity in sharing genetic information, chromosomes are first aligned on a metaphase spindle and only once alignment is achieved does anaphase onset occur (Mitchison and Salmon, 2001; Tanaka, 2002). Newly replicated sister chromatids are held together in S-phase onwards by cohesin complexes (Fig. 4), which are believed to form ring-like structures embracing both sisters (Gruber et al., 2003; Uhlmann, 2004; Nasmyth and Haering, 2005). A heterodimeric pair SMC1 and SMC3, members from the structural maintenance of chromosomes (SMC) family, and SCC1/RAD21/MCD1 and SCC3, which has two vertebrate isoforms STAG1/SA1 and STAG2/SA2, constitute the cohesin complex. The SMC1–SMC3 heterodimer joins in a head-to-head and tail-to-tail configuration, with their tails forming a hinge-like structure (Hirano, 2006). At their head they associate with SCC1 and SCC3, proteins which are modified when cohesin’s hold on sisters needs be released during the mitotic division.

Figure 4:

The cohesin complex

Sister chromatids are held together by the cohesin complex, which embraces them in a ring-like structure. The components of the cohesin complex are labeled and their meiotic counterparts tabulated. Note that at anaphase-onset separase proteolytically cleaves Sister chromatin cohesion protein 1 (SCC1), allowing sisters to separate under pulling forces from spindle microtubules. SMC, structural maintenance of chromosomes protein

In meiosis some of the mitotic components of the cohesin complex are replaced (Fig. 4). Therefore SCC1 is largely replaced by REC8 (Klein et al., 1999; Parisi et al., 1999; Watanabe and Nurse, 1999); SA1/SA2 by STAG3 (Pezzi et al., 2000; Prieto et al., 2001) and SMC1 (now better named SMC1α) by SMC1β (Revenkova et al., 2001). Having a meiotic variant is likely to confer some specialized meiotic function: e.g. in fission yeast REC8 can rescue mitosis loss of SCC1, but SCC1 cannot rescue meiotic loss of REC8 (Watanabe and Nurse, 1999); similarly, mammalian REC8 can bind to SMC1β but not SMC1α (Lee et al., 2003). Loss of REC8 in mouse leads to homolog pairing but no homolog synapsis (Xu et al., 2005). One possible further meiotic function for REC8 observed in yeasts and Arabidopsis is in the establishment and maintenance of monopolar orientation of the MI sister kinetochores (Chelysheva et al., 2005; Yokobayashi and Watanabe, 2005). This is necessary only in MI because sister kinetochores act as a functional unit binding to the same pole, a process in budding yeast known to be dependent on the Aurora B kinase (Monje-Casas et al., 2007) and the Monopolin complex (Toth et al., 2000; Rabitsch et al., 2001), whose mammalian homologs intriguingly remain unknown.

Passage through meiosis I

Loss of oocytes during early life

Oocytes from newborn animals are often connected by cytoplasmic bridges in aggregates known as ‘cysts’ (Pepling and Spradling, 1998). These cysts rapidly brake down in the days following birth and oocytes individualize (Pepling and Spradling, 2001); this is discussed in an account elsewhere (Pepling, 2006). Following birth, oocyte numbers drop substantially and such loss of oocytes may be similar to the so-called pachytene, or meiotic recombination, checkpoint in budding yeasts (Roeder and Bailis, 2000). The oocyte within primordial and primary follicles does have the capacity to detect DNA damage through the p53-like gene, TA-p63α (Suh et al., 2006), a process which intuitively should be active to get rid of any DNA-damaged oocytes during their period of dictyate arrest. However, it is also clear that oocytes also have a DNA damage independent mechanism of cell death (Di Giacomo et al., 2005).

Signalling between oocyte and follicular cells

Outside the scope of this review is the cyclical recruitment of ovarian follicles from the non-growing pool (McGee and Hsueh, 2000; Webb et al., 2003; Knox, 2005; Visser and Themmen, 2005). Important in the health of the oocyte during this period of arrest is bi-directional communication between the oocyte and the follicular granulosa cells (Eppig, 2001). Various factors are likely provided to the oocyte by the follicular somatic cells which support growth, and here interestingly the cumulus-derived sterol FF-MAS may act to suppress pre-division (Cukurcam et al., 2007). Conversely, oocyte-derived members of the transforming growth factor (TGF)β superfamily act as paracrine factors to influence follicle development (Drummond, 2005), such as Growth Differentiation Factor-9 whose loss blocks growth of primary follicles (Dong et al., 1996). The close relationship between oocyte and surrounding granulosa cells is highlighted by the fact that oocytes are deficient in their ability to generate ATP through the glycolytic pathway; instead they secrete fibroblast growth factor and bone morphogenetic factor 15, a member of the TGFβ superfamily, to stimulate glycolytic activity in adjoining granulosa cells and in so doing provide them with nutrients (Sugiura et al., 2007). Oocyte-derived TGFβ paracrine factors are also involved in stimulating granulosa cell growth, preventing apoptosis and play a role in the process of cumulus cell expansion which is associated with oocyte maturation (Hussein et al., 2005; Gilchrist et al., 2006; Dragovic et al., 2007; Hutt and Albertini, 2007).

From GV arrest to polar body extrusion

Having undergone S-phase during early fetal life as discussed earlier oocytes remain arrested within growing follicles. The transition from the dictyate stage into diakinesis, which occurs physiologically within large antral follicles, forms part of a process termed ‘oocyte maturation’ or ‘meiotic maturation’, which encompasses other cellular events that prepare the oocyte for fertilization, such as a maturation of the cortical granules responsible for the block to polyspermy (Ducibella et al., 1988; Abbott et al., 1999). Associated loss of the nuclear envelope (called ‘germinal vesicle’, GV, in oocytes) gives the start of the process the name of germinal besicle breakdown (GVB or GVBD). Nuclear maturation is the name given to the events controlling the actual cell cycle progression leading to eventual arrest at MetII, whereas cytoplasmic maturation refers to events and processes that are not chromatin based, such as the aforementioned cortical granule maturation. Here the review will concentrate only on nuclear maturation, which is driven by rising cyclin-dependent kinase (CDK1) activity.

Oocyte maturation finishes with a small first polar body (PB1) being extruded. This unequal cell division results from a cortical migration of the spindle dependent on actin microfilaments (Verlhac et al., 2000), and the actin-binding protein formin-2 (Leader et al., 2002; Dumont et al., 2007). As the spindle nears the oolemma, a region above the spindle accumulates actin, myosin, Par6 and the small GTPases Rho, Rac and Cdc42 (Vinot et al., 2004; Na and Zernicka-Goetz, 2006; Deng et al., 2007; Halet and Carroll, 2007), which are essential for polarity in a number of systems. This cortical actin cap is therefore associated with the spindle pole nearest the plasma membrane, and with its associated set of separated homologs is cleaved from the oocyte by a myosin ring constriction. Presumably, the purpose of an unequal cell division is to ensure retention of maternal RNA and protein useful for MetII and early embryogenesis within the oocyte, as well as providing a too small a target for any fertilizing sperm that may fuse with the polar body rather than the oocyte proper.

cAMP in maintaining GV arrest

Once follicles become fully grown, continued dictyate arrest requires high protein kinase A (PKA) activity inside the oocyte. Therefore, removal from the ovary necessitates addition of cyclic AMP (cAMP) analogs or phosphodiesterase inhibitors to the culture media to maintain arrest (Eppig, 1989; Tornell et al., 1991; Conti et al., 1998; Dekel, 2005; Mehlmann, 2005). In vivo a G-protein coupled receptor on the plasma membrane of the oocyte is a constitutively-active receptor providing adequate cAMP to maintain GV arrest (Mehlmann et al., 2002, 2004). Whether the periovulatory surge in LH switches off this receptor or up-regulates phosphodiesterases in order to lower PKA activity remains to be established (Mehlmann et al., 2006). Also still under investigation are the intermediaries which provide the necessary signals from the rise in LH to oocyte, since oocytes lack LH receptors. Possibly involved is a cumulus cell-derived sterol FF-MAS (Byskov et al., 1995, 2002), but the physiological relevance of FF-MAS to LH-induced meiotic resumption is now questioned and other mechanisms likely (Tsafriri et al., 2005; Conti et al., 2006). Despite all these uncertainties, it is generally agreed that further progression through MI is mediated by a drop in intra-oocyte cAMP.

CDK1 and GVB

Critical to maintenance of GV arrest is low CDK1 activity (a.k.a maturation or M-phase promoting factor; cdc2; p34cdc2), a kinase implicated in dissolution of the nuclear envelope and condensation of chromatin (Doree and Hunt, 2002; Jones, 2004): two events associated with entry into the first meiotic division. Key to maintaining low CDK1 is regulation of the kinase and phosphatase responsible for phosphorylating/dephosphorylating CDK1, respectively (Fig. 5). Thus during GV arrest the activity of wee1B, the kinase which negatively affects CDK1 activity through phosphorylation is high, while cdc25B, the phosphatase responsible for activating CDK1 is low; probably by direct phosphorylation through PKA (Duckworth et al., 2002; Lincoln et al., 2002; Han et al., 2005). The ability of PKA to negatively affect CDK1 activity, and maintain it in a so-called pre-MPF state, would be the physiological basis of the need to maintain a high cAMP drive in oocytes for GV arrest. Interestingly, although multiple members of the wee1 and cdc25 family exist, specific isoforms are responsible for regulating CDK1 in mammalian oocytes (Lincoln et al., 2002; Han et al., 2005). For example, oocytes from cdc25b knockout mice lack the ability to undergo GVB, even though they contain cdc25a and cdc25c (Lincoln et al., 2002).

Figure 5:

Germinal vesicle stage arrest

(A) Oocytes from mature follicles are prevented from undergoing GVB by high intra-oocyte cAMP levels. cAMP acts via protein kinase A to keep wee1B active and cdc25B inactive. In so doing, maturation-promoting factor (MPF) is maintained in a phosphorylated and inactive state (Pre-MPF). A reduction in cAMP is believed to be the physiological trigger for inducing GVB. Also during GV arrest levels of cyclin B1 (CyB), the regulatory partner of CDK1, are controlled through degradation mediated by the anaphase-promoting factor bound to its co-activator cdh1 (APCcdh1), which is itself regulated by Emi1. During GV arrest cyclin B1 levels must be maintained by some mRNA translation to nullify its continual degradation. (B) Loss of cAMP at GVB allows MPF activity to rise. This negatively affects wee1B activity but stimulates that of cdc25B. Rising MPF may also switch on more cyclin B1 translation and will act eventually to switch off APCcdh1 activity (although this latter event does not happen until several hours after GVB). See text for further details

Although in vivo release from GV arrest is likely mediated through regulation of CDK1, cyclin B1 regulation also appears important for maintenance of GV arrest. This is because it is possible to raise CDK1 activity by overexpressing cyclin B1, the regulatory partner of CDK1, and in so doing induce GVB even in the presence of high cAMP (Ledan et al., 2001; Marangos and Carroll, 2004; Reis et al., 2006). Thus there is turnover of cyclin B1 in the GV-arrested oocytes to prevent precocious GVB (Reis et al., 2006). This is achieved through the anaphase-promoting complex/cyclosome (APC), a large multisubunit complex which acts as an E3 ubiquitin ligase, able to bind specific substrates such as cyclin B1 and earmark them for proteolysis through the 26S proteasome by a process of polyubiquitination (Peters, 2006). The APC is active in GV-arrested mouse oocytes, through association with one of its two key activators, cdh1, and this APCcdh1 activity is needed to maintain oocytes in GV arrest (Reis et al., 2006).

APCcdh1, the regulator of cyclin B1 levels in oocytes, is itself regulated during GV arrest by association with its inhibitor Emi1. Emi1 was first described as a G2-phase cell cycle regulator of APC—acting to prevent precocious APC activity before mitosis (Reimann et al., 2001). Therefore if Emi1 is knocked-down in oocytes, it increases APCcdh1 activity and makes oocytes refractory for GVB, presumably by lowering cyclin B1 (Marangos et al., 2007). Conversely, overexpressing Emi1 leads to GVB, presumably by inhibition of APCcdh1 allowing cyclin B1 levels to accumulate.

Oocytes from preantral follicle of an adult mouse or oocytes from young mice (Day 15 or younger) fail to undergo maturation in culture (Erickson and Sorensen, 1974; Sorensen and Wassarman, 1976). This is most likely due to insufficient levels or insufficient interaction of CDK1 and cyclin B1 (de Vant’ery et al., 1996, 1997; Mitra and Schultz, 1996; Kanatsu-Shinohara et al., 2000). Indeed, overexpression of both CDK1 and cyclin B1 is needed to induce GVB in oocytes from young mice (de Vantery et al., 1997). It remains a challenge to refine in vitro culture systems for human oocytes derived from preantral stages of follicle growth, in order to deliver oocytes that are capable of undergoing in vitro maturation and eventual fertilization (Picton et al., 2003). Especially so, since it is during oocyte growth that various mRNAs are made and stored, disruption of this process severely affecting later oocyte maturation and fertilization (Yu et al., 2004), and is also when imprints are laid down for proper control of gene expression (Obata and Kono, 2002).

Interestingly, frog eggs contain an activator of CDK1 which is not cyclin B (Ferby et al., 1999). This protein, called RINGO (rapid inducer of G2/M progression in oocytes) is able to bind CDK1 so inducing GVB and its loss from frog oocytes blocks this transition. Its role in the process of GVB in mammalian oocytes is little explored. Frog RINGO can induce GVB in mouse and pig oocytes but blocks them in MI (Terret et al., 2001; Kume et al., 2007), whereas pig RINGO allows oocytes to complete MI (Kume et al., 2007). This suggests that mammalian RINGO homologs may have a physiological role in GVB. Future experiments are therefore required to assess the relative contribution of cyclin B1 and RINGO in the activation of CDK1 to trigger GVB.

Cohesin loss: meiosis V mitosis

The completion of the first meiotic division, from GVB to the extrusion of the PB1 marks the period in which the second hit, important in the derivation of aneuploidy, is thought to operate. In essence, it is regarded as an age-related decline in the ability of the maturing oocyte during MI to respond to errors in non- or poorly recombined homologs. It is first worth examining how faithful segregation of chromosomes (sister chromatids) are achieved in mitosis and then examining if the same mechanisms are known to operate in oocytes during this first, reductional division. During prometaphase in mitosis, non-centromeric cohesin complexes are removed by phosphorylation rather than proteolysis: involving Aurora B and polo (Losada et al., 2002; Sumara et al., 2002; Gimenez-Abian et al., 2004). Polo-kinase directly phosphorylates SCC3 subunits (Hauf et al., 2005). However, centromeric SCC3 is protected from phosphorylation by the protein Shugoshin (Sgo) and the phosphatase PP2A until the metaphase–anaphase transition (McGuinness et al., 2005; Kitajima et al., 2006; Riedel et al., 2006). Loss of centromeric cohesin at anaphase is caused by proteolysis of its SCC1 subunit by the protease separase (Uhlmann et al., 1999, 2000; Waizenegger et al., 2000; Hauf et al., 2001), possibly aided by degradation of Sgo (Salic et al., 2004; Fu et al., 2007). Premature activity of separase is prevented through binding the protein securin; which in mammals was first described as pituitary tumour transforming gene1 (Zou et al., 1999; Waizenegger et al., 2000). Therefore when sister chromatids have fully aligned on a metaphase plate, timely separation is achieved by securin proteolysis, which allows separase to act on Scc1, a process first described in yeasts (Ciosk et al., 1998; Uhlmann et al., 1999).

In meiosis, cohesin is lost in a step-like manner, with non-centromeric cohesin removed during MI; and centromeric cohesin, which was protected by Sgo1, during MetII. Both the non-centromeric loss of cohesin in MI and the centromeric loss in MetII seem to be dependent on REC8 proteolysis by separase (Huo et al., 2006; Kudo et al., 2006; Lee et al., 2006). This therefore appears to be different to mitosis where the non-centromeric loss would be driven not by proteolysis, but by polo-kinase (Hauf et al., 2005).

Regulation of the mitotic division by the SAC

In order to understand how segregation of homologs may be regulated during MI, we first must examine how sister segregation is regulated in mitosis. At the metaphase–anaphase transition mitotic cells need to decrease their CDK1 activity (Peters, 2002, 2006; Pines, 2006). This is achieved by degradation of CDK1’s regulatory partner cyclin B1 (Jones, 2004). Loss of cyclin B1, and also securin to free separase and so cleave SCC1 (or REC8), is through APC-dependent proteolysis. The high CDK1 activity prevents cdh1 association with the APC (Zachariae et al., 1998; Jaspersen et al., 1999; Blanco et al., 2000), instead it is another WD-repeat protein cdc20 which is required for APC activity at metaphase (Peters, 2002, 2006; Pines, 2006). Before a mitotic cell wants to commit to segregating its chromosomes, APCcdc20 activity is held low by activation of the spindle assembly checkpoint (SAC) (Musacchio and Salmon, 2007). The SAC pathway is active until chromosomes are fully aligned and under tension, with sister pairs attached to opposite spindle poles, a condition called bi-orientation. There are many protein members of the SAC that were initially identified from studies in yeasts in genetic screens of cells failing to arrest in mitosis in the presence of spindle poisons that disrupted assembly of the spindle (Hoyt et al., 1991; Li and Murray, 1991; Li et al., 1993; Weiss and Winey, 1996). These include the mitotic arrest deficient gene products Mad1, Mad2 and Mad3 (BubR1 in mammalian cells) and the budding uninhibited by benzimidazole gene product Bub1, Bub2 and Bub3 as well as the protein Mps1.

Members of the SAC are thought to be able to respond to a lack of tension and/or attachment of microtubules to the kinetochore by blocking the ability of cdc20 to bind and so activate the APC (Hwang et al., 1998; Kim et al., 1998). The relative contribution of tension and attachment to activation of the SAC is much debated and difficult to ascertain because of their inter-dependence (Pinsky and Biggins, 2005). A lack of attachment consequently leads to a lack of tension; which can also weaken attachment, a necessary process in establishing correct attachment of sister kinetochores to opposite spindle poles. However, a mitotic checkpoint complex containing Mad2, BubR1 and Bub3, in addition to cdc20, appears to be the effector of the SAC and is thought to require kinetochores for its assembly (Sudakin et al., 2001; Musacchio and Salmon, 2007).

The SAC in meiosis

Attempts to examine if SAC components are absent or somehow non-functional in oocytes have in fact often reached the surprising conclusion that they are present and active. Rodent oocytes are known to contain Mad2 (Wassmann et al., 2003; Zhang et al., 2004; Homer et al., 2005a, b; Wang et al., 2007) and Bub1 (Brunet et al., 2003; Yin et al., 2006); and importantly knocking down or interfering with their function, as well as that of BubR1, accelerates passage through MI and, when examined, leads to mis-segregated homologs (Tsurumi et al., 2004; Zhang et al., 2004; Homer et al., 2005a, b; Yin et al., 2006; Wang et al., 2007). This suggests that normally Mad2, Bub1 and BubR1 all function in slowing passage through MI, and by implication constitute an active SAC.

It is important to consider what precisely are the triggers for activation of the SAC during MI. Each sister kinetochore pair must act as a functional unit to establish bi-orientation of the bivalent. It is therefore interesting that oocytes from female mice carrying a univalent X fail to show any metaphase arrest (LeMaire-Adkins et al., 1997). In these oocytes it is possible that univalents evade the SAC by bi-orientating their sister kinetochores and so undergoing pre-division (Fig. 2), as observed in oocytes lacking Sycp3, a component of the synaptonemal complex (Kouznetsova et al., 2007). However, the univalent X is reported to segregate reductionally, suggesting that tension across the univalent is not detected. Interpreting these observations and with those of the Mad2, Bub1 and BubR1 data described above, suggests that oocytes do possess a SAC-like mechanism which can halt segregation of chromosomes during MI, however, it does not necessarily monitor homolog bi-orientation. Recent work in budding yeast during MI suggest that Mad2, and by implication the SAC, may only be important in re-orientating homologs that have distal sites of crossover from the kinetochore (Lacefield and Murray, 2007). The argument being that with proximal exchange the bivalent has a more rigid geometry which predisposes it to bi-orientation. However in distal crossovers the bivalent could be considered more lithe and open to mono-orientation of sister kinetochores.

It has been suggested that an early prometaphase SAC-like mechanism, dependent on Mad2 and BubR1 but independent of kinetochores, is present in cells which can delay passage through mitosis (Meraldi et al., 2004). Such a delay could be due to an ability to bind and so sequester cdc20 (Burton and Solomon, 2007). This early mechanism is then followed by a true SAC, dependent on kinetochores, later in prometaphase (Musacchio and Salmon, 2007). It follows that loss of SAC proteins such as Mad2 and BubR1 could accelerate MI, without there being a proper kinetochore-dependent SAC in oocytes. Similarly, overexpression of SAC proteins such as Mad2, known to block oocytes at metaphase I (Homer et al. 2005a, b), would act due to an ability to bind and sequester cdc20 from the APC, rather than by imposition of a SAC. However the simplest experiment, addition of spindle poisons, to oocytes have all demonstrated directly the existence of a SAC by blocking oocytes in MI (Wassmann et al., 2003; Homer et al., 2005a, b). Clearly, whatever the detail is in the SAC-imposed arrest upon addition of spindle poisons, such a pathway is not executed when univalent homologs are present in MI.

Control of CDK1 loss during MI

One intriguing feature of MI is its extraordinary length. In cultured cells undergoing mitosis, from the time of nuclear envelope breakdown to cytokinesis is ∼20–30 min (Rieder et al., 1994; Meraldi et al., 2004), compared with the several hours from GVB until PB1 extrusion in mammalian oocytes. The extended period of meiosis is reflected in the dynamics of CDK1 activity, which rises abruptly in mitosis but gradually in meiosis, reaching a peak several hours after GVB (Choi et al., 1991; Gavin et al., 1994; Verlhac et al., 1994, 1996). Loss of CDK1 activity is an essential requirement for completion of MI, and specifically degradation of cyclin B1. Also important is degradation of securin to free separase. In mouse, loss of both cyclin B1 and securin occur synchronously in a period which terminates with PB1 extrusion and is dependent on APCcdc20 (Reis et al., 2007), as it is in mitosis. However, early in prometaphase I APCcdh1 is active, made possible by low CDK1 activity at this time, and here it degrades cdc20. This means cdc20 has to be resynthesized in order for oocytes to complete MI and suggests an unusual progression through chromosome segregation not observed with sisters in mitosis. Loss of cdh1 brings forward the period of APCcdc20 activity and consequently the period of cyclin B1 and securin degradation (Reis et al., 2007). This premature metaphase I induces high rates of non-disjunction and leads to a disruption of the integrity of the MetII spindle probably as a consequence of a chromosome crowded spindle if hyperploid. Similar reasoning may be behind the disruptions of the spindle structure that have been observed in human oocytes from older women (Rosenbusch and Schneider, 2006; Shen et al., 2006). Interestingly, the SAC had not been switched off in these cdh1-depleted oocytes (Reis et al., 2007), since spindle poisons were equally effective at blocking exit from MI whether cdh1 was present or not. Therefore the acceleration of MI generated mis-segregated homologs and the SAC was not activated. Again, this suggests the inability of the SAC to monitor homolog bi-orientation.

Loss of CDK1 activity through degradation of cyclin B1 may not be the sole mechanism for regulating CDK1 in oocytes. Free separase, generated by proteolysis of securin, can also bind CDK1 and in so doing inhibit its kinase activity (Stemmann et al., 2001, 2006); in fact, inhibition is mutual since CDK1 binding inhibits separase proteolytic activity too (Gorr et al., 2005). The ability of separase to bind CDK1 also appears essential for PB1, since PB1 extrusion is blocked when separase binding is inhibited (Gorr et al., 2006). The ability of CDK1 activity in oocytes to be regulated by both loss of cyclin B1 and separase binding warrants further investigation given this process is required for completion of MI. Future studies are therefore required to assess their inter-dependence.

The SAC or cohesion deteriorates with age?

It remains possible that SAC function could deteriorate with maternal age to account for the age-related incidence of aneuploidy. Therefore one could argue that the SAC is fully functional in oocytes from younger women, but its ability to respond to mis-alignment of homologs on the MI spindle weakens in oocytes from older females. Mad2 protein and mRNA both seem relatively unstable in MetII oocytes making such a hypothesis attractive if their instability extends to immature oocytes (Ma et al., 2005; Steuerwald et al., 2005). Interestingly, these authors hypothesize that the post-ovulatory loss of Mad2 may be responsible for the increased sister chromatid segregation defects associated with post-ovulatory aging. However, the SAC, although it can be activated by various spindle poisons, is currently thought not to be involved in maintaining the physiological arrest during MetII in mammalian oocytes, so establishing any causal link would be important (Tsurumi et al., 2004; Madgwick et al., 2006; Shoji et al., 2006; Madgwick and Jones, 2007).

Only one study has addressed if SAC components decrease in human oocytes with age. Here it was found that transcript levels of both Mad2 and Bub1 decrease with increasing maternal age (Steuerwald et al., 2001). It will be important to corroborate this study and also provide direct evidence that the SAC is indeed compromised in oocytes from older women during MI. This would address whether the reduced transcript levels affect protein levels and consequently SAC function. Interestingly, levels of hundreds of transcripts, including cell cycle genes, have been been reported to decrease with increased maternal age in mice and women (Hamatani et al., 2004; Steuerwald et al., 2007). This raises the possibility that aneuploidy may result due to falling levels of various components of the cell cycle, not just SAC members.

One possible target rather than the SAC to explain age-related aneuploidy is the cohesin complex itself. Theoretically, if functionality of the complex were to decline with age it would bring together many of the loose threads presented thus far. For example, the reduced recombination frequency, assessed by chiasmata, with age in mouse (Henderson and Edwards, 1968). This was interpreted in terms of a production line because of recombination frequency being established in fetal life. However, one would observe a similar loss in chiasmata if the cohesin ties holding homologs together were loosened during dictyate arrest to the extent of separating homologs. The fact that the cohesin complex holding homologs together is established during fetal S-phase and yet has to remain functional decades later would make it susceptible to age-related damage because it may be difficult or impossible to repair. Mutation of the meiosis-specific cohesin SMC1β makes female mice sterile due to large levels of aneuploidy (Revenkova et al., 2004). Interestingly, these mice display an age-related incidence in aneuploidy with 4-week old mice containing oocytes with bivalents only, whereas essentially no intact homologs could be observed by 4 months (Hodges et al., 2005). The ability of the aneuploidy rate to rise with maternal age suggests that during early oocyte growth and follicle maturation cohesins may deteriorate and need to be replaced. The inability of these SMC1β mutant mice to perform this replacement would then account for the maternal age effect. One would predict that oocytes in which higher than average rates of recombination exist may be in part protected against any age-related loss in sister chromatid cohesion, and indeed human studies examining recombination frequency of offspring against their mother’s age appear to bear this out (Kong et al., 2004).

Age-associated aneuploidy: environmental and hormonal factors

An age-related decline in the cohesive ties holding chromosomes together coupled to an intrinsic inability of oocytes to detect or repair the resulting error is likely to constitute the etiology of some aneuploidy. This process may be exacerbated by environmental exposure to agents which interfere with detection or repair and here it critical to best match environmental exposure to controlled laboratory conditions. Evidence exists that either neonatal or adult exposure to the weak estrogen bisphenol A (BPA) can induce aneuploidy in oocytes, suggesting environmental pollutants could underlie some age-related aneuploidy. Non-disjunction in mouse is an uncommon event in most strains, yet very high rates of aneuploidy can result when mice are exposed to BPA, a common constituent of polycarbonate plastics and resins used to line food cans and make dental sealants. Aneuploidy happens when oocytes are exposed during their maturation in vitro or when oocytes are exposed to excess BPA in the animals diet (Hunt et al., 2003; Can et al., 2005). Also exposure of fetal oocytes to BPA can affect the placement and number of recombination events on chromosomes, this probably generates susceptible patterns of exchange, and so leads to increased non-disjunction in treated animals (Susiarjo et al., 2007). It is suggested that BPA may interfere with the function of estrogen receptor beta (ERβ) at this time, given the same recombination phenotype is observed in ERβ knockout mice (Susiarjo et al., 2007). However, this is unexpected as it suggests a very early stimulatory effect of a fetal estrogen directly on oocyte recombination.

Exposure of maturing mouse oocytes to high levels of the gonadotropin FSH can also induce aneuploidy in vitro (Roberts et al., 2005). Such an observation is interesting from the perspective of circulating gonadotropins being higher in older women with diminishing ovarian reserve, and the observations that increased FSH has been measured in women having DS children (van Montfrans et al., 2002). This latter effect appears to be due to a decreased ovarian reserve. The influence of FSH on aneuploidy therefore requires much further investigation.

Completion of meiosis II at fertilization

The preovulatory LH surge induces oocyte maturation and subsequent re-arrest at metaphase of the second meiotic division just prior to ovulation. Arrest at MetII and subsequent fertilization at this stage of meiosis is universal in mammals, even in Canidae oocytes which are exceptional in being ovulated at the GV stage (Reynaud et al., 2006). Re-arrest is protracted and oocytes of most animals seem to show a very good block to further progression through MetII (Jones, 1998).

Cytostatic factor

Re-arrest is achieved by the development of an activity during oocyte maturation first described as ‘cytostatic factor’ (CSF) (Masui and Markert, 1971), which at molecular level is now thought likely to be expression in maturing oocytes of the APC inhibitor Emi2/Erp1(Liu and Maller, 2005; Rauh et al., 2005; Schmidt et al., 2005; Tung et al., 2005; Madgwick et al., 2006; Shoji et al., 2006). The degradation of this oocyte-specific protein at fertilization is triggered by its phosphorylation through the concerted activity of polo kinase and calmodulin-dependent kinase II (CamKII) (Madgwick et al., 2006; Schmidt et al., 2006). Although polo is already active in unfertilized oocytes, CamKII is not. Thus Emi2 is only fully phosphorylated and so degraded when CamKII is switched on at fertilization (Hansen et al., 2006); which in mammals is through an oscillatory Ca2+ signal induced by a sperm-specific member of the phospholipase C (PLC) family called PLC zeta (Saunders et al., 2002; Kouchi et al., 2004; Swann et al., 2006). Fully phosphorylated Emi2 is then degraded by the E3 ligase complex Skp1-Cullin/F-box protein. The mos/MAPkinase pathway has long been implicated also in MetII arrest (O’Keefe et al., 1989; Sagata et al., 1989). Recently, in frog it was demonstrated that the mos pathway is responsible for stabilizing Emi2 through p90 ribosomal S6 kinase (p90rsk) (Inoue et al., 2007; Nishiyama et al., 2007) so neatly establishing a link between the long-time CSF candidate mos and the much newer contender Emi2. Interestingly, p90rsk plays no part in arresting mouse oocytes at MetII, so it will be important to establish the link between the mos and Emi2 in mammals (Dumont et al., 2005), given mos clearly plays a role in arresting mammalian oocytes at MetII (Colledge et al., 1994; Hashimoto et al., 1994; Verlhac et al., 1996).

Degradation of Emi2 is responsible for the 6–7-fold increase in APC activity observed at fertilization (Nixon et al., 2002) that leads to rapid cyclin B1 proteolysis. Therefore a drop in CDK1 activity is an early event of fertilization. Also likely degraded is securin which is probably re-synthesized during MetII arrest, thus contributing to separase inhibition. However, it is also possible that the high CDK1 activity associated with MetII arrest inhibits separase, as has been observed in other cells (Gorr et al., 2005; Holland and Taylor, 2006). Interestingly, the binding of CDK1 to separase leads to a mutual inhibition of activities, with binding being thought to play an essential role in inhibiting CDK1 during MI (Gorr et al., 2006). Overexpression of non-degradable cyclin B1 in fertilizing oocytes blocks not only the decline in CDK1 activity, but also sister chromatid separation, suggesting that CDK1 can function as a separase inhibitor during MetII arrest (Madgwick et al., 2004); although whether this applies at physiological levels of CDK1 remains to be established.

The ability of ovulated oocytes to maintain a fully functional spindle with bi-orientated sisters is probably finite and would account for the drop in oocyte quality associated with post-ovulatory aging. Interestingly, increasing female age exacerbates the decline in oocyte quality with post-ovulatory aging and is probably accounted for by a decline in their ability to maintain high CDK1 activity (Tatone et al., 2006). Metabolic, biochemical and structural parameters also decline in oocytes undergoing post-ovulatory ageing (Takahashi et al., 2000; Igarashi et al., 2005; Huang et al., 2007) and the fertilization-associated calcium signal seems to switch on apoptotic pathways as oocytes age (Gordo et al., 2002). Finally, there appears important long-term detriment to the animal derived from a post-ovulatory aged oocyte, since both its reproductive fitness as an adult and its longevity are adversely affected (Tarin et al., 2002).

Aneuploidy in oocytes: is it a human condition?

Much of the present review has focussed on the mouse oocyte. Yet, many will argue that the human oocyte is uniquely prone to segregation errors in meiosis and therefore any other model system, including mouse, is inappropriate for the study of aneuploidy and in particular age-related aneuploidy. Such judgement may be erroneous. Even the evolutionary distant but important model organism Drosophila melanogaster can display increased incidence of age-related aneuploidy on a background where sister chromatid cohesion is perturbed (Jeffreys et al., 2003).

It is true that many strains of mice have low rates of aneuploidy and the above age-related aneuploidy in Drosophila requires disruption in the expression of the ORD gene product involved in sister chromatid cohesion. However, high, human-like levels of aneuploidy can exist without perturbation in some mice. For example, crosses of C57BL/6 with SPRET mice, generate F1 female mice with an aneuploidy rate of ∼10% at age 4 weeks, corresponding to a young adult, which then doubles to 20% in mice aged 8–11 months that are at the end of their reproductive life (Koehler et al., 2002). This increase in aneuploidy was associated with a decrease in recombination frequency between homologs and is likely due to sequence divergence in the homologs of these two strains given the same phenomenon can be observed in close strains of yeasts (Hunter et al., 1996). Also some strains of mice such as CBA display higher rates of aneuploidy than other strains (Eichenlaub-Ritter et al., 1988). As an aside it is important to note that achiasmate homologs are a feature of meiosis in a number of organisms, especially in Drosophila (Wolf, 1994; Thomas et al., 2005). One mechanism to account for proper segregation of achiasmate homologs in Drosophila is heterochromatin pairing (Karpen et al., 1996), such that non-disjunction is substantially increased when heterochromatin pairing is reduced by deleting heterochromatin content of an introduced mini-chromosome. It would be interesting to determine if pairing of homologs in mammalian oocytes, independent of chiasmata, play any part in MI as it can in other organisms (Gerton and Hawley, 2005).

Concluding remarks

This review has had two main purposes: first, to give a broad review of the cell biology responsible for the meiotic cell cycle transitions which define the remarkable life of an oocyte; second, to put our knowledge of cell biology into a clinical context by using it to discuss our current understanding of the unique susceptibility of the oocyte to aneuploidy. It is hoped that despite the fact that the etiology of aneuploidy is unlikely to be found in one particular meiotic defect or even pathway, our understanding of its causes will likely come from basic cell biology done on model organisms which are more tractable than human, and such knowledge will eventually feed into a clinical setting by collaborations of basic and applied researchers who together may find newer and better methods for prevention, screening and possible treatment. Major themes basic cell biologist are likely to make substantial progress in the next decade are (i) how germ cells commit to entry into meiosis; (ii) how homologs are held together for the maintenance of cohesion and how they may deteriorate with age; (iii) how the oocyte remains viable during a protracted period of GV arrest; (iv) how the oocyte controls the segregation of homologs during the first meiotic division and (v) how the oocyte maintains MetII arrested and how this is detrimentally affected by post-ovulatory ageing. Pursuit of answers to these questions will likely lead to a better understanding of aneuploidy in oocytes.

Funding

The author would like to acknowledge continued funding from the Wellcome Trust.

References

Abbott
AL
Fissore
RA
Ducibella
T
Incompetence of preovulatory mouse oocytes to undergo cortical granule exocytosis following induced calcium oscillations
Dev Biol
1999
, vol. 
207
 (pg. 
38
-
48
)
Angell
RR
Predivision in human oocytes at meiosis I: a mechanism for trisomy formation in man
Hum Genet
1991
, vol. 
86
 (pg. 
383
-
387
)
Baltus
AE
Menke
DB
Hu
YC
Goodheart
ML
Carpenter
AE
de Rooij
DG
Page
DC
In germ cells of mouse embryonic ovaries, the decision to enter meiosis precedes premeiotic DNA replication
Nat Genet
2006
, vol. 
38
 (pg. 
1430
-
1434
)
Baudat
F
Manova
K
Yuen
JP
Jasin
M
Keeney
S
Chromosome synapsis defects and sexually dimorphic meiotic progression in mice lacking Spo11
Mol Cell
2000
, vol. 
6
 (pg. 
989
-
998
)
Bell
R
Rankin
J
Donaldson
LJ
Down’s syndrome: occurrence and outcome in the north of England, 1985–99
Paediatr Perinat Epidemiol
2003
, vol. 
17
 (pg. 
33
-
39
)
Bendsen
E
Byskov
AG
Andersen
CY
Westergaard
LG
Number of germ cells and somatic cells in human fetal ovaries during the first weeks after sex differentiation
Hum Reprod
2006
, vol. 
21
 (pg. 
30
-
35
)
Blanco
MA
Sanchez-Diaz
A
de Prada
JM
Moreno
S
APCste9/srw1 promotes degradation of mitotic cyclins in G1 and is inhibited by cdc2 phosphorylation
Embo J
2000
, vol. 
19
 (pg. 
3945
-
3955
)
Blaschke
RJ
Rappold
G
The pseudoautosomal regions, SHOX and disease
Curr Opin Genet Dev
2006
, vol. 
16
 (pg. 
233
-
239
)
Bowles
J
Knight
D
Smith
C
Wilhelm
D
Richman
J
Mamiya
S
Yashiro
K
Chawengsaksophak
K
Wilson
MJ
Rossant
J
, et al. 
Retinoid signaling determines germ cell fate in mice
Science
2006
, vol. 
312
 (pg. 
596
-
600
)
Brunet
S
Pahlavan
G
Taylor
S
Maro
B
Functionality of the spindle checkpoint during the first meiotic division of mammalian oocytes
Reproduction
2003
, vol. 
126
 (pg. 
443
-
450
)
Bullejos
M
Koopman
P
Germ cells enter meiosis in a rostro-caudal wave during development of the mouse ovary
Mol Reprod Dev
2004
, vol. 
68
 (pg. 
422
-
428
)
Burton
JL
Solomon
MJ
Mad3p, a pseudosubstrate inhibitor of APCCdc20 in the spindle assembly checkpoint
Genes Dev
2007
, vol. 
21
 (pg. 
655
-
667
)
Byskov
AG
Andersen
CY
Nordholm
L
Thogersen
H
Xia
G
Wassmann
O
Andersen
JV
Guddal
E
Roed
T
Chemical structure of sterols that activate oocyte meiosis
Nature
1995
, vol. 
374
 (pg. 
559
-
562
)
Byskov
AG
Andersen
CY
Leonardsen
L
Role of meiosis activating sterols, MAS, in induced oocyte maturation
Mol Cell Endocrinol
2002
, vol. 
187
 (pg. 
189
-
196
)
Can
A
Semiz
O
Cinar
O
Bisphenol-A induces cell cycle delay and alters centrosome and spindle microtubular organization in oocytes during meiosis
Mol Hum Reprod
2005
, vol. 
11
 (pg. 
389
-
396
)
Chelysheva
L
Diallo
S
Vezon
D
Gendrot
G
Vrielynck
N
Belcram
K
Rocques
N
Marquez-Lema
A
Bhatt
AM
Horlow
C
, et al. 
AtREC8 and AtSCC3 are essential to the monopolar orientation of the kinetochores during meiosis
J Cell Sci
2005
, vol. 
118
 (pg. 
4621
-
4632
)
Choi
T
Aoki
F
Mori
M
Yamashita
M
Nagahama
Y
Kohmoto
K
Activation of p34cdc2 protein kinase activity in meiotic and mitotic cell cycles in mouse oocytes and embryos
Development
1991
, vol. 
113
 (pg. 
789
-
795
)
Ciosk
R
Zachariae
W
Michaelis
C
Shevchenko
A
Mann
M
Nasmyth
K
An ESP1/PDS1 complex regulates loss of sister chromatid cohesion at the metaphase to anaphase transition in yeast
Cell
1998
, vol. 
93
 (pg. 
1067
-
1076
)
Colledge
WH
Carlton
MB
Udy
GB
Evans
MJ
Disruption of c-mos causes parthenogenetic development of unfertilized mouse eggs
Nature
1994
, vol. 
370
 (pg. 
65
-
68
)
Conti
M
Andersen
CB
Richard
FJ
Shitsukawa
K
Tsafriri
A
Role of cyclic nucleotide phosphodiesterases in resumption of meiosis
Mol Cell Endocrinol
1998
, vol. 
145
 (pg. 
9
-
14
)
Conti
M
Hsieh
M
Park
JY
Su
YQ
Role of the epidermal growth factor network in ovarian follicles
Mol Endocrinol
2006
, vol. 
20
 (pg. 
715
-
723
)
Coop
G
Przeworski
M
An evolutionary view of human recombination
Nat Rev Genet
2007
, vol. 
8
 (pg. 
23
-
34
)
Cooper
JP
Watanabe
Y
Nurse
P
Fission yeast Taz1 protein is required for meiotic telomere clustering and recombination
Nature
1998
, vol. 
392
 (pg. 
828
-
831
)
Coory
MD
Roselli
T
Carroll
HJ
Antenatal care implications of population-based trends in Down syndrome birth rates by rurality and antenatal care provider, Queensland, 1990–2004
Med J Aust
2007
, vol. 
186
 (pg. 
230
-
234
)
Cukurcam
S
Betzendahl
I
Michel
G
Vogt
E
Hegele-Hartung
C
Lindenthal
B
Eichenlaub-Ritter
U
Influence of follicular fluid meiosis-activating sterol on aneuploidy rate and precocious chromatid segregation in aged mouse oocytes
Hum Reprod
2007
, vol. 
22
 (pg. 
815
-
828
)
de Vant’ery
C
Gavin
AC
Vassalli
JD
Schorderet-Slatkine
S
An accumulation of p34cdc2 at the end of mouse oocyte growth correlates with the acquisition of meiotic competence
Dev Biol
1996
, vol. 
174
 (pg. 
335
-
344
)
de Vantery
C
Stutz
A
Vassalli
JD
Schorderet-Slatkine
S
Acquisition of meiotic competence in growing mouse oocytes is controlled at both translational and posttranslational levels
Dev Biol
1997
, vol. 
187
 (pg. 
43
-
54
)
Dekel
N
Cellular, biochemical and molecular mechanisms regulating oocyte maturation
Mol Cell Endocrinol
2005
, vol. 
234
 (pg. 
19
-
25
)
Delhanty
JD
Mechanisms of aneuploidy induction in human oogenesis and early embryogenesis
Cytogenet Genome Res
2005
, vol. 
111
 (pg. 
237
-
244
)
Deng
M
Suraneni
P
Schultz
RM
Li
R
The Ran GTPase mediates chromatin signaling to control cortical polarity during polar body extrusion in mouse oocytes
Dev Cell
2007
, vol. 
12
 (pg. 
301
-
308
)
Di Giacomo
M
Barchi
M
Baudat
F
Edelmann
W
Keeney
S
Jasin
M
Distinct DNA-damage-dependent and -independent responses drive the loss of oocytes in recombination-defective mouse mutants
Proc Natl Acad Sci
2005
, vol. 
102
 (pg. 
737
-
742
)
Dong
J
Albertini
DF
Nishimori
K
Kumar
TR
Lu
N
Matzuk
MM
Growth differentiation factor-9 is required during early ovarian folliculogenesis
Nature
1996
, vol. 
383
 (pg. 
531
-
535
)
Doree
M
Hunt
T
From Cdc2 to Cdk1: when did the cell cycle kinase join its cyclin partner? J
Cell Sci
2002
, vol. 
115
 (pg. 
2461
-
2464
)
Dragovic
RA
Ritter
LJ
Schulz
SJ
Amato
F
Thompson
JG
Armstrong
DT
Gilchrist
RB
Oocyte-secreted factor activation of SMAD 2/3 signaling enables initiation of mouse cumulus cell expansion
Biol Reprod
2007
, vol. 
76
 (pg. 
848
-
857
)
Drummond
A
TGFβ signalling in the development of ovarian function
Cell Tissue Res
2005
, vol. 
322
 (pg. 
107
-
115
)
Ducibella
T
Rangarajan
S
Anderson
E
The development of mouse oocyte cortical reaction competence is accompanied by major changes in cortical vesicles and not cortical granule depth
Dev Biol
1988
, vol. 
130
 (pg. 
789
-
792
)
Duckworth
BC
Weaver
JS
Ruderman
JV
G2 arrest in Xenopus oocytes depends on phosphorylation of cdc25 by protein kinase A
Proc Natl Acad Sci USA
2002
, vol. 
99
 (pg. 
16794
-
16799
)
Dumont
J
Million
K
Sunderland
K
Rassinier
P
Lim
H
Leader
B
Verlhac
H
Formin-2 is required for spindle migration and for the late steps of cytokinesis in mouse oocytes
Dev Biol
2007
, vol. 
301
 (pg. 
254
-
265
)
Dumont
J
Umbhauer
M
Rassinier
P
Hanauer
A
Verlhac
MH
p90Rsk is not involved in cytostatic factor arrest in mouse oocytes
J Cell Biol
2005
, vol. 
169
 (pg. 
227
-
231
)
Egan
JF
Benn
PA
Zelop
CM
Bolnick
A
Gianferrari
E
Borgida
AF
Down syndrome births in the United States from 1989 to 2001
Am J Obstet Gynecol
2004
, vol. 
191
 (pg. 
1044
-
1048
)
Eggan
K
Jurga
S
Gosden
R
Min
IM
Wagers
AJ
Ovulated oocytes in adult mice derive from non-circulating germ cells
Nature
2006
, vol. 
441
 (pg. 
1109
-
1114
)
Eichenlaub-Ritter
U
Chandley
AC
Gosden
RG
The CBA mouse as a model for age-related aneuploidy in man: studies of oocyte maturation, spindle formation and chromosome alignment during meiosis
Chromosoma
1988
, vol. 
96
 (pg. 
220
-
226
)
Eichenlaub-Ritter
U
Vogt
E
Yin
H
Gosden
R
Spindles, mitochondria and redox potential in ageing oocytes
Reprod Biomed Online
2004
, vol. 
8
 (pg. 
45
-
58
)
Eppig
JJ
The participation of cyclic adenosine monophosphate (cAMP) in the regulation of meiotic maturation of oocytes in the laboratory mouse
J Reprod Fertil Suppl
1989
, vol. 
38
 (pg. 
3
-
8
)
Eppig
JJ
Oocyte control of ovarian follicular development and function in mammals
Reproduction
2001
, vol. 
122
 (pg. 
829
-
838
)
Erickson
GF
Sorensen
RA
In vitro maturation of mouse oocytes isolated from late, middle, and pre-antral graafian follicles
J Exp Zool
1974
, vol. 
190
 (pg. 
123
-
127
)
Ferby
I
Blazquez
M
Palmer
A
Eritja
R
Nebreda
AR
A novel p34cdc2-binding and activating protein that is necessary and sufficient to trigger G2/M progression in Xenopus oocytes
Genes Dev
1999
, vol. 
13
 (pg. 
2177
-
2189
)
Fu
G
Hua
S
Ward
T
Ding
X
Yang
Y
Guo
Z
Yao
X
D-box is required for the degradation of human Shugoshin and chromosome alignment
Biochem Biophys Res Commun
2007
, vol. 
357
 (pg. 
672
-
678
)
Garcia
M
Dietrich
AJ
Freixa
L
Vink
AC
Ponsa
M
Egozcue
J
Development of the first meiotic prophase stages in human fetal oocytes observed by light and electron microscopy
Hum Genet
1987
, vol. 
77
 (pg. 
223
-
232
)
Gavin
AC
Cavadore
JC
Schorderet-Slatkine
S
Histone H1 kinase activity, germinal vesicle breakdown and M phase entry in mouse oocytes
J Cell Sci
1994
, vol. 
107
 (pg. 
275
-
283
)
Gerton
JL
Hawley
RS
Homologous chromosome interactions in meiosis: diversity amidst conservation
Nat Rev Genet
2005
, vol. 
6
 (pg. 
477
-
487
)
Gilchrist
RB
Ritter
LJ
Myllymaa
S
Kaivo-Oja
N
Dragovic
RA
Hickey
TE
Ritvos
O
Mottershead
DG
Molecular basis of oocyte-paracrine signalling that promotes granulosa cell proliferation
J Cell Sci
2006
, vol. 
119
 (pg. 
3811
-
3821
)
Gimenez-Abian
JF
Sumara
I
Hirota
T
Hauf
S
Gerlich
D
de la Torre
C
Ellenberg
J
Peters
M
Regulation of sister chromatid cohesion between chromosome arms
Curr Biol
2004
, vol. 
14
 (pg. 
1187
-
1193
)
Gordo
AC
Rodrigues
P
Kurokawa
M
Jellerette
T
Exley
GE
Warner
C
Fissore
R
Intracellular calcium oscillations signal apoptosis rather than activation in in vitro aged mouse eggs
Biol Reprod
2002
, vol. 
66
 (pg. 
1828
-
1837
)
Gorr
IH
Boos
D
Stemmann
O
Mutual inhibition of separase and Cdk1 by two-step complex formation
Mol Cell
2005
, vol. 
19
 (pg. 
135
-
141
)
Gorr
IH
Reis
A
Boos
D
Wuhr
M
Madgwick
S
Jones
KT
Stemmann
O
Essential CDK1-inhibitory role for separase during meiosis I in vertebrate oocytes
Nat Cell Biol
2006
, vol. 
8
 (pg. 
1035
-
1037
)
Graves
JA
Wakefield
MJ
Toder
R
The origin and evolution of the pseudoautosomal regions of human sex chromosomes
Hum Mol Gen
1998
, vol. 
7
 (pg. 
1991
-
1996
)
Gruber
S
Haering
CH
Nasmyth
K
Chromosomal cohesin forms a ring
Cell
2003
, vol. 
112
 (pg. 
765
-
777
)
Halet
G
Carroll
J
Rac activity is polarized and regulates meiotic spindle stability and anchoring in mammalian oocytes
Dev Cell
2007
, vol. 
12
 (pg. 
309
-
317
)
Hamatani
T
Falco
G
Carter
MG
Akutsu
H
Stagg
CA
Sharov
AA
Dudekula
DB
VanBuren
V
Ko
MSH
Age-associated alteration of gene expression patterns in mouse oocytes
Hum Mol Genet
2004
, vol. 
13
 (pg. 
2263
-
2278
)
Han
SJ
Chen
R
Paronetto
MP
Conti
M
Wee1B Is an oocyte-specific kinase involved in the control of meiotic arrest in the mouse
Curr Biol
2005
, vol. 
15
 (pg. 
1670
-
1676
)
Hansen
DV
Tung
JJ
Jackson
PK
CaMKII and polo-like kinase 1 sequentially phosphorylate the cytostatic factor Emi2/XErp1 to trigger its destruction and meiotic exit
Proc Natl Acad Sci USA
2006
, vol. 
103
 (pg. 
608
-
613
)
Hashimoto
N
Watanabe
N
Furuta
Y
Tamemoto
H
Sagata
N
Yokoyama
M
Okazaki
K
Nagayoshi
M
Takeda
N
Ikawa
Y
, et al. 
Parthenogenetic activation of oocytes in c-mos-deficient mice
Nature
1994
, vol. 
370
 (pg. 
68
-
71
)
Hassold
TJ
Jacobs
PA
Trisomy in man
Annu Rev Genet
1984
, vol. 
18
 (pg. 
69
-
97
)
Hassold
T
Sherman
S
Hunt
P
Counting cross-overs: characterizing meiotic recombination in mammals
Hum Mol Genet
2000
, vol. 
9
 (pg. 
2409
-
2419
)
Hassold
T
Hunt
P
To err (meiotically) is human: the genesis of human aneuploidy
Nat Rev Genet
2001
, vol. 
2
 (pg. 
280
-
291
)
Hauf
S
Waizenegger
IC
Peters
JM
Cohesin cleavage by separase required for anaphase and cytokinesis in human cells
Science
2001
, vol. 
293
 (pg. 
1320
-
1323
)
Hauf
S
Roitinger
E
Koch
B
Dittrich
CM
Mechtler
K
Peters
JM
Dissociation of cohesin from chromosome arms and loss of arm cohesion during early mitosis depends on phosphorylation of SA2
PLoS Biol
2005
, vol. 
3
 pg. 
e69
 
Hawley
RS
Frazier
JA
Rasooly
R
Separation anxiety: the etiology of nondisjunction in flies and people
Hum Mol Genet
1994
, vol. 
3
 (pg. 
1521
-
1528
)
Henderson
SA
Edwards
RG
Chiasma frequency and maternal age in mammals
Nature
1968
, vol. 
218
 (pg. 
22
-
28
)
Hirano
T
At the heart of the chromosome: SMC proteins in action
Nat Rev Mol Cell Biol
2006
, vol. 
7
 (pg. 
311
-
322
)
Hirshfield
AN
Heterogeneity of cell populations that contribute to the formation of primordial follicles in rats
Biol Reprod
1992
, vol. 
47
 (pg. 
466
-
472
)
Hodges
CA
Revenkova
E
Jessberger
R
Hassold
TJ
Hunt
PA
SMC1beta-deficient female mice provide evidence that cohesins are a missing link in age-related nondisjunction
Nat Genet
2005
, vol. 
37
 (pg. 
1351
-
1355
)
Holland
AJ
Taylor
SS
Cyclin-B1-mediated inhibition of excess separase is required for timely chromosome disjunction
J Cell Sci
2006
, vol. 
119
 (pg. 
3325
-
3336
)
Homer
HA
McDougall
A
Levasseur
M
Murdoch
AP
Herbert
M
Mad2 is required for inhibiting securin and cyclin B degradation following spindle depolymerisation in meiosis I mouse oocytes
Reproduction
2005
, vol. 
a 130
 (pg. 
829
-
843
)
Homer
HA
McDougall
A
Levasseur
M
Yallop
K
Murdoch
AP
Herbert
M
Mad2 prevents aneuploidy and premature proteolysis of cyclin B and securin during meiosis I in mouse oocytes
Genes Dev
2005
, vol. 
b 19
 (pg. 
202
-
207
)
Hoyt
MA
Totis
L
Roberts
BT
S. cerevisiae genes required for cell cycle arrest in response to loss of microtubule function
Cell
1991
, vol. 
66
 (pg. 
507
-
517
)
Huang
J-C
Yan
L-Y
Lei
Z-L
Miao
Y-L
Shi
L-H
Yang
J-W
Wang
Q
Ouyang
Y-C
Sun
Q-Y
Chen
Y
Changes in histone acetylation during postovulatory aging of mouse oocyte
Biol Reprod
2007
, vol. 
77
 (pg. 
666
-
670
)
Hunt
PA
Koehler
KE
Susiarjo
M
Hodges
CA
Ilagan
A
Voigt
RC
Thomas
S
Thomas
BF
Hassold
TJ
Bisphenol A exposure causes meiotic aneuploidy in the female mouse
Curr Biol
2003
, vol. 
13
 (pg. 
546
-
553
)
Hunter
N
Chambers
SR
Louis
EJ
Borts
RH
The mismatch repair system contributes to meiotic sterility in an interspecific yeast hybrid
Embo J
1996
, vol. 
15
 (pg. 
1726
-
1733
)
Huo
L-J
Zhong
Z-S
Liang
C-G
Wang
Q
Yin
S
Ai
J-S
Yu
L-Z
Chen
D-Y
Schatten
H
Sun
Y
Degradation of securin in mouse and pig oocytes is dependent on ubiquitin-proteasome pathway and is required for proteolysis of the cohesion subunit, Rec8, at the metaphase-to-anaphase transition
Front Biosci
2006
, vol. 
11
 (pg. 
2193
-
2202
)
Hussein
TS
Froiland
DA
Amato
F
Thompson
JG
Gilchrist
RB
Oocytes prevent cumulus cell apoptosis by maintaining a morphogenic paracrine gradient of bone morphogenetic proteins
J Cell Sci
2005
, vol. 
118
 (pg. 
5257
-
5268
)
Hutt
KJ
Albertini
DF
An oocentric view of folliculogenesis and embryogenesis
Reprod Biomed Online
2007
, vol. 
14
 (pg. 
758
-
764
)
Hwang
LH
Lau
LF
Smith
DL
Mistrot
CA
Hardwick
KG
Hwang
ES
Amon
A
Murray
AW
Budding yeast Cdc20: a target of the spindle checkpoint
Science
1998
, vol. 
279
 (pg. 
1041
-
1044
)
Igarashi
H
Takahashi
T
Takahashi
E
Tezuka
N
Nakahara
K
Takahashi
K
Kurachi
H
Aged mouse oocytes fail to readjust intracellular adenosine triphosphates at fertilization
Biol Reprod
2005
, vol. 
72
 (pg. 
1256
-
1261
)
Inoue
D
Ohe
M
Kanemori
Y
Nobui
T
Sagata
N
A direct link of the Mos-MAPK pathway to Erp1/Emi2 in meiotic arrest of Xenopus laevis eggs
Nature
2007
, vol. 
446
 (pg. 
1100
-
1104
)
Jaspersen
SL
Charles
JF
Morgan
DO
Inhibitory phosphorylation of the APC regulator Hct1 is controlled by the kinase Cdc28 and the phosphatase Cdc14
Curr Biol
1999
, vol. 
9
 (pg. 
227
-
236
)
Jeffreys
CA
Burrage
PS
Bickel
SE
A model system for increased meiotic nondisjunction in older oocytes
Curr Biol
2003
, vol. 
13
 (pg. 
498
-
503
)
Johnson
J
Canning
J
Kaneko
T
Pru
JK
Tilly
JL
Germline stem cells and follicular renewal in the postnatal mammalian ovary
Nature
2004
, vol. 
428
 (pg. 
145
-
150
)
Johnson
J
Bagley
J
Skaznik-Wikiel
M
Lee
H-J
Adams
GB
Niikura
Y
Tschudy
KS
Tilly
JC
Cortes
ML
Forkert
R
, et al. 
Oocyte generation in adult mammalian ovaries by putative germ cells in bone marrow and peripheral blood
Cell
2005
, vol. 
122
 (pg. 
303
-
315
)
Jones
KT
Ca2+ oscillations in the activation of the egg and development of the embryo in mammals
Int J Dev Biol
1998
, vol. 
42
 (pg. 
1
-
10
)
Jones
KT
Turning it on and off: M-phase promoting factor during meiotic maturation and fertilization
Mol Hum Reprod
2004
, vol. 
10
 (pg. 
1
-
5
)
Jou
HJ
Kuo
YS
Hsu
JJ
Shyu
MK
Hsieh
TT
Hsieh
FJ
The evolving national birth prevalence of Down syndrome in Taiwan. A study on the impact of second-trimester maternal serum screening
Prenat Diagn
2005
, vol. 
25
 (pg. 
665
-
670
)
Kanatsu-Shinohara
M
Schultz
RM
Kopf
GS
Acquisition of meiotic competence in mouse oocytes: absolute amounts of p34cdc2, cyclin B1, cdc25C, and wee1 in meiotically incompetent and competent oocytes
Biol Reprod
2000
, vol. 
63
 (pg. 
1610
-
1616
)
Karpen
GH
Le
MH
Le
H
Centric heterochromatin and the efficiency of achiasmate disjunction in Drosophila female meiosis
Science
1996
, vol. 
273
 (pg. 
118
-
122
)
Keefe
DL
Marquard
K
Liu
L
The telomere theory of reproductive senescence in women
Curr Opin Obstet Gynecol
2006
, vol. 
18
 (pg. 
280
-
285
)
Keefe
DL
Liu
L
Marquard
K
Telomeres and aging-related meiotic dysfunction in women
Cell Mol Life Sci
2007
, vol. 
64
 (pg. 
139
-
143
)
Keeney
S
Neale
MJ
Initiation of meiotic recombination by formation of DNA double-strand breaks: mechanism and regulation
Biochem Soc Trans
2006
, vol. 
34
 (pg. 
523
-
525
)
Khoshnood
B
De Vigan
C
Vodovar
V
Goujard
J
Goffinet
F
A population-based evaluation of the impact of antenatal screening for Down’s syndrome in France, 1981–2000
Bjog
2004
, vol. 
a 111
 (pg. 
485
-
490
)
Khoshnood
B
Wall
S
Pryde
P
Lee
KS
Maternal education modifies the age-related increase in the birth prevalence of Down syndrome
Prenat Diagn
2004
, vol. 
b 24
 (pg. 
79
-
82
)
Kim
SH
Lin
DP
Matsumoto
S
Kitazono
A
Matsumoto
T
Fission yeast Slp1: an effector of the Mad2-dependent spindle checkpoint
Science
1998
, vol. 
279
 (pg. 
1045
-
1047
)
Kitajima
TS
Sakuno
T
Ishiguro
K
Iemura
S
Natsume
T
Kawashima
SA
Watanabe
Y
Shugoshin collaborates with protein phosphatase 2A to protect cohesin
Nature
2006
, vol. 
441
 (pg. 
46
-
52
)
Kleckner
N
Chiasma formation: chromatin/axis interplay and the role(s) of the synaptonemal complex
Chromosoma
2006
, vol. 
115
 (pg. 
175
-
194
)
Klein
F
Mahr
P
Galova
M
Buonomo
SB
Michaelis
C
Nairz
K
Nasmyth
K
A central role for cohesins in sister chromatid cohesion, formation of axial elements, and recombination during yeast meiosis
Cell
1999
, vol. 
98
 (pg. 
91
-
103
)
Knox
RV
Recruitment and selection of ovarian follicles for determination of ovulation rate in the pig
Domest Anim Endocrinol
2005
, vol. 
29
 (pg. 
385
-
397
)
Koehler
KE
Boulton
CL
Collins
HE
French
RL
Herman
KC
Lacefield
SM
Madden
LD
Schuetz
CD
Hawley
RS
Spontaneous X chromosome MI and METII nondisjunction events in Drosophila melanogaster oocytes have different recombinational histories
Nat Genet
1996
, vol. 
14
 (pg. 
406
-
414
)
Koehler
KE
Millie
EA
Cherry
JP
Burgoyne
PS
Evans
EP
Hunt
PA
Hassold
TJ
Sex-specific differences in meiotic chromosome segregation revealed by dicentric bridge resolution in mice
Genetics
2002
, vol. 
162
 (pg. 
1367
-
1379
)
Kong
A
Barnard
J
Gudbjartsson
DF
Thorleifsson
G
Jonsdottir
G
Sigurdardottir
S
Richardsson
B
Jonsdottir
J
Thorgeirsson
T
Frigge
ML
, et al. 
Recombination rate and reproductive success in humans
Nat Genet
2004
, vol. 
36
 (pg. 
1203
-
1206
)
Koubova
J
Menke
DB
Zhou
Q
Capel
B
Griswold
MD
Page
DC
Retinoic acid regulates sex-specific timing of meiotic initiation in mice
Proc Natl Acad Sci USA
2006
, vol. 
103
 (pg. 
2474
-
2479
)
Kouchi
Z
Fukami
K
Shikano
T
Oda
S
Nakamura
Y
Takenawa
T
Miyazaki
S
Recombinant phospholipase Czeta has high Ca2+ sensitivity and induces Ca2+ oscillations in mouse eggs
J Biol Chem
2004
, vol. 
279
 (pg. 
10408
-
10412
)
Kouznetsova
A
Lister
L
Nordenskjold
M
Herbert
M
Hoog
C
Bi-orientation of achiasmatic chromosomes in meiosis I oocytes contributes to aneuploidy in mice
Nat Genet
2007
, vol. 
39
 (pg. 
966
-
968
)
Kudo
NR
Wassmann
K
Anger
M
Schuh
M
Wirth
KG
Xu
H
Helmhart
W
Kudo
H
McKay
M
Maro
B
, et al. 
Resolution of chiasmata in oocytes requires separase-mediated proteolysis
Cell
2006
, vol. 
126
 (pg. 
135
-
146
)
Kuliev
A
Cieslak
J
Verlinsky
Y
Frequency and distribution of chromosome abnormalities in human oocytes
Cytogenet Genome Res
2005
, vol. 
111
 (pg. 
193
-
198
)
Kume
S
Endo
T
Nishimura
Y
Kano
K
Naito
K
Porcine SPDYA2 (RINGO A2) stimulates CDC2 Activity and accelerates meiotic maturation of porcine oocytes
Biol Reprod
2007
, vol. 
76
 (pg. 
440
-
447
)
Kurilo
LF
Oogenesis in antenatal development in man
Hum Genet
1981
, vol. 
57
 (pg. 
86
-
92
)
Lacefield
S
Murray
AW
The spindle checkpoint rescues the meiotic segregation of chromosomes whose crossovers are far from the centromere
Nat Genet
2007
, vol. 
39
 (pg. 
1273
-
1277
)
Lai
FM
Woo
BH
Tan
KH
Huang
J
Lee
ST
Yan
TB
Tan
BH
Chew
SK
Yeo
GS
Birth prevalence of Down syndrome in Singapore from 1993 to 1998
Singapore Med J
2002
, vol. 
43
 (pg. 
70
-
76
)
Lamb
NE
Freeman
SB
Savage-Austin
A
Pettay
D
Taft
L
Hersey
J
Gu
Y
Shen
J
Saker
D
May
KM
, et al. 
Susceptible chiasmate configurations of chromosome 21 predispose to non-disjunction in both maternal meiosis I and meiosis II
Nat Genet
1996
, vol. 
14
 (pg. 
400
-
405
)
Lamb
NE
Feingold
E
Savage
A
Avramopoulos
D
Freeman
S
Gu
Y
Hallberg
A
Hersey
J
Karadima
G
Pettay
D
, et al. 
Characterization of susceptible chiasma configurations that increase the risk for maternal nondisjunction of chromosome 21
Hum Mol Genet
1997
, vol. 
6
 (pg. 
1391
-
1399
)
Lawson
KA
Hage
WJ
Clonal analysis of the origin of primordial germ cells in the mouse
Ciba Found Symp
1994
, vol. 
182
 (pg. 
68
-
84
discussion 84–91
Leader
B
Lim
H
Carabatsos
MJ
Harrington
A
Ecsedy
J
Pellman
D
Maas
R
Leder
P
Formin-2, polyploidy, hypofertility and positioning of the meiotic spindle in mouse oocytes
Nat Cell Biol
2002
, vol. 
4
 (pg. 
921
-
928
)
Ledan
E
Polanski
Z
Terret
ME
Maro
B
Meiotic maturation of the mouse oocyte requires an equilibrium between cyclin B synthesis and degradation
Dev Biol
2001
, vol. 
232
 (pg. 
400
-
413
)
Lee
J
Iwai
T
Yokota
T
Yamashita
M
Temporally and spatially selective loss of Rec8 protein from meiotic chromosomes during mammalian meiosis
J Cell Sci
2003
, vol. 
116
 (pg. 
2781
-
2790
)
Lee
J
Okada
K
Ogushi
S
Miyano
T
Miyake
M
Yamashita
M
Loss of Rec8 from chromosome arm and centromere region is required for homologous chromosome separation and sister chromatid separation, respectively, in mammalian meiosis
Cell Cycle
2006
, vol. 
5
 (pg. 
1448
-
1455
)
LeMaire-Adkins
R
Radke
K
Hunt
PA
Lack of checkpoint control at the metaphase/anaphase transition: a mechanism of meiotic nondisjunction in mammalian females
J Cell Biol
1997
, vol. 
139
 (pg. 
1611
-
1619
)
Li
R
Murray
AW
Feedback control of mitosis in budding yeast
Cell
1991
, vol. 
66
 (pg. 
519
-
531
)
Li
R
Havel
C
Watson
JA
Murray
AW
The mitotic feedback control gene MAD2 encodes the alpha-subunit of a prenyltransferase
Nature
1993
, vol. 
366
 (pg. 
82
-
84
)
Li
X
Schimenti
JC
Mouse pachytene checkpoint 2 (trip13) is required for completing meiotic recombination but not synapsis
PLoS Genet
2007
, vol. 
3
 pg. 
e130
 
Lincoln
AJ
Wickramasinghe
D
Stein
P
Schultz
RM
Palko
ME
De Miguel
MP
Tessarollo
L
Donovan
PJ
Cdc25b phosphatase is required for resumption of meiosis during oocyte maturation
Nat Genet
2002
, vol. 
30
 (pg. 
446
-
449
)
Liu
J
Maller
JL
Calcium elevation at fertilization coordinates phosphorylation of XErp1/Emi2 by Plx1 and CaMK II to release metaphase arrest by cytostatic factor
Curr Biol
2005
, vol. 
15
 (pg. 
1458
-
1468
)
Losada
A
Hirano
M
Hirano
T
Cohesin release is required for sister chromatid resolution, but not for condensin-mediated compaction, at the onset of mitosis
Genes Dev
2002
, vol. 
16
 (pg. 
3004
-
3016
)
Luthardt
FW
Palmer
CG
Yu
P
Chiasma and univalent frequencies in aging female mice
Cytogenet Cell Genet
1973
, vol. 
12
 (pg. 
68
-
79
)
Ma
W
Zhang
D
Hou
Y
Li
YH
Sun
QY
Sun
XF
Wang
WH
Reduced expression of MAD2, BCL2, and MAP kinase activity in pig oocytes after in vitro aging are associated with defects in sister chromatid segregation during meiosis II and embryo fragmentation after activation
Biol Reprod
2005
, vol. 
72
 (pg. 
373
-
383
)
Madgwick
S
Jones
KT
How eggs arrest at metaphase II: MPF stabilisation plus APC/C inhibition equals Cytostatic Factor
Cell Div
2007
, vol. 
2
 pg. 
4
 
Madgwick
S
Nixon
VL
Chang
HY
Herbert
M
Levasseur
M
Jones
KT
Maintenance of sister chromatid attachment in mouse eggs through maturation-promoting factor activity
Dev Biol
2004
, vol. 
275
 (pg. 
68
-
81
)
Madgwick
S
Hansen
DV
Levasseur
M
Jackson
PK
Jones
KT
Mouse Emi2 establishes a metaphase II spindle by re-stabilizing cyclin B1 during interkinesis
J Cell Biol
2006
, vol. 
174
 (pg. 
791
-
801
)
Marangos
P
Carroll
J
The dynamics of cyclin B1 distribution during meiosis I in mouse oocytes
Reproduction
2004
, vol. 
128
 (pg. 
153
-
162
)
Marangos
P
Verschuren
EW
Chen
R
Jackson
PK
Carroll
J
Prophase I arrest and progression to metaphase I in mouse oocytes are controlled by Emi1-dependent regulation of APCCdh1
J Cell Biol
2007
, vol. 
176
 (pg. 
65
-
75
)
Masui
Y
Markert
CL
Cytoplasmic control of nuclear behavior during meiotic maturation of frog oocytes
J Exp Zool
1971
, vol. 
177
 (pg. 
129
-
145
)
McGee
EA
Hsueh
AJW
Initial and cyclic recruitment of ovarian follicles
Endocr Rev
2000
, vol. 
21
 (pg. 
200
-
214
)
McGuinness
BE
Hirota
T
Kudo
NR
Peters
JM
Nasmyth
K
Shugoshin prevents dissociation of cohesin from centromeres during mitosis in vertebrate cells
PLoS Biol
2005
, vol. 
3
 pg. 
e86
 
McLaren
A
Germ and somatic cell lineages in the developing gonad
Mol Cell Endocrinol
2000
, vol. 
163
 (pg. 
3
-
9
)
McLaren
A
Primordial germ cells in the mouse
Dev Biol
2003
, vol. 
262
 (pg. 
1
-
15
)
Mehlmann
LM
Stops and starts in mammalian oocytes: recent advances in understanding the regulation of meiotic arrest and oocyte maturation
Reproduction
2005
, vol. 
130
 (pg. 
791
-
799
)
Mehlmann
LM
Jones
TL
Jaffe
LA
Meiotic arrest in the mouse follicle maintained by a Gs protein in the oocyte
Science
2002
, vol. 
297
 (pg. 
1343
-
1345
)
Mehlmann
LM
Kalinowski
RR
Ross
LF
Parlow
AF
Hewlett
EL
Jaffe
LA
Meiotic resumption in response to luteinizing hormone is independent of a Gi family G protein or calcium in the mouse oocyte
Dev Biol
2006
, vol. 
299
 (pg. 
345
-
355
)
Mehlmann
LM
Saeki
Y
Tanaka
S
Brennan
TJ
Evsikov
AV
Pendola
FL
Knowles
BB
Eppig
JJ
Jaffe
LA
The Gs-linked receptor GPR3 maintains meiotic arrest in mammalian oocytes
Science
2004
, vol. 
306
 (pg. 
1947
-
1950
)
Menke
DB
Koubova
J
Page
DC
Sexual differentiation of germ cells in XX mouse gonads occurs in an anterior-to-posterior wave
Dev Biol
2003
, vol. 
262
 (pg. 
303
-
312
)
Meraldi
P
Draviam
VM
Sorger
PK
Timing and checkpoints in the regulation of mitotic progression
Dev Cell
2004
, vol. 
7
 (pg. 
45
-
60
)
Meredith
S
Doolin
D
Timing of activation of primordial follicles in mature rats is only slightly affected by fetal stage at meiotic arrest
Biol Reprod
1997
, vol. 
57
 (pg. 
63
-
67
)
Mitchison
TJ
Salmon
ED
Mitosis: a history of division
Nat Cell Biol
2001
, vol. 
3
 (pg. 
E17
-
E21
)
Mitra
J
Schultz
RM
Regulation of the acquisition of meiotic competence in the mouse: changes in the subcellular localization of cdc2, cyclin B1, cdc25C and wee1, and in the concentration of these proteins and their transcripts
J Cell Sci
1996
, vol. 
109
 (pg. 
2407
-
2415
)
Monje-Casas
F
Prabhu
VR
Lee
BH
Boselli
M
Amon
A
Kinetochore orientation during meiosis is controlled by aurora b and the monopolin complex
Cell
2007
, vol. 
128
 (pg. 
477
-
490
)
Monk
M
McLaren
A
X-chromosome activity in fetal germ cells of the mouse
J Embryol Exp Morphol
1981
, vol. 
63
 (pg. 
75
-
84
)
Morelli
MA
Cohen
PE
Not all germ cells are created equal: Aspects of sexual dimorphism in mammalian meiosis
Reproduction
2005
, vol. 
130
 (pg. 
761
-
781
)
Mori
T
Kurahashi
H
Shinka
T
Nakahori
Y
Taniguchi
M
Toda
T
Iwamoto
T
Candidate genes for male factor infertility—validation
Fert Steril
2006
, vol. 
86
 (pg. 
1553
-
1554
)
Morris
JK
Mutton
DE
Alberman
E
Revised estimates of the maternal age specific live birth prevalence of Down’s syndrome
J Med Screen
2002
, vol. 
9
 (pg. 
2
-
6
)
Musacchio
A
Salmon
ED
The spindle-assembly checkpoint in space and time
Nat Rev Mol Cell Biol
2007
, vol. 
8
 (pg. 
379
-
393
)
Na
J
Zernicka-Goetz
M
Asymmetric positioning and organization of the meiotic spindle of mouse oocytes requires CDC42 function
Curr Biol
2006
, vol. 
16
 (pg. 
1249
-
1254
)
Nasmyth
K
Haering
CH
The structure and function of SMC and kleisin complexes
Ann Rev Biochem
2005
, vol. 
74
 (pg. 
595
-
648
)
Nimmo
ER
Pidoux
AL
Perry
PE
Allshire
RC
Defective meiosis in telomere-silencing mutants of Schizosaccharomyces pombe
Nature
1998
, vol. 
392
 (pg. 
825
-
828
)
Nishiyama
T
Ohsumi
K
Kishimoto
T
Phosphorylation of Erp1 by p90rsk is required for cytostatic factor arrest in Xenopus laevis eggs
Nature
2007
, vol. 
446
 (pg. 
1096
-
1099
)
Nixon
VL
Levasseur
M
McDougall
A
Jones
KT
Ca2+ oscillations promote APC/C-dependent cyclin B1 degradation during metaphase arrest and completion of meiosis in fertilizing mouse eggs
Curr Biol
2002
, vol. 
12
 (pg. 
746
-
750
)
O’Keefe
SJ
Wolfes
H
Kiessling
AA
Cooper
GM
Microinjection of antisense c-mos oligonucleotides prevents meiosis II in the maturing mouse egg
Proc Natl Acad Sci USA
1989
, vol. 
86
 (pg. 
7038
-
7042
)
O’Leary
P
Breheny
N
Dickinson
JE
Bower
C
Goldblatt
J
Hewitt
B
Murch
A
Stock
R
First-trimester combined screening for Down syndrome and other fetal anomalies
Obstet Gynecol
2006
, vol. 
107
 (pg. 
869
-
876
)
Obata
Y
Kono
T
Maternal primary imprinting is established at a specific time for each gene throughout oocyte growth
J Biol Chem
2002
, vol. 
277
 (pg. 
5285
-
5289
)
Pacchierotti
F
Adler
ID
Eichenlaub-Ritter
U
Mailhes
JB
Gender effects on the incidence of aneuploidy in mammalian germ cells
Environ Res
2007
, vol. 
104
 (pg. 
46
-
69
)
Page
SL
Hawley
RS
The genetics and molecular biology of the synaptonemal complex
Annu Rev Cell Dev Biol
2004
, vol. 
20
 (pg. 
525
-
558
)
Parisi
S
McKay
MJ
Molnar
M
Thompson
MA
van der Spek
PJ
van Drunen-Schoenmaker
E
Kanaar
R
Lehmann
E
Hoeijmakers
JH
Kohli
J
Rec8p, a meiotic recombination and sister chromatid cohesion phosphoprotein of the Rad21p family conserved from fission yeast to humans
Mol Cell Biol
1999
, vol. 
19
 (pg. 
3515
-
3528
)
Pellestor
F
Andreo
B
Arnal
F
Humeau
C
Demaille
J
Maternal aging and chromosomal abnormalities: new data drawn from in vitro unfertilized human oocytes
Hum Genet
2003
, vol. 
112
 (pg. 
195
-
203
)
Pellestor
F
Anahory
T
Hamamah
S
Effect of maternal age on the frequency of cytogenetic abnormalities in human oocytes
Cytogenet Genome Res
2005
, vol. 
111
 (pg. 
206
-
212
)
Pellestor
F
Andreo
B
Anahory
T
Hamamah
S
The occurrence of aneuploidy in human: lessons from the cytogenetic studies of human oocytes
Eur J Med Genet
2006
, vol. 
49
 (pg. 
103
-
116
)
Pepling
ME
From primordial germ cell to primordial follicle: mammalian female germ cell development
Genesis
2006
, vol. 
44
 (pg. 
622
-
632
)
Pepling
ME
Spradling
AC
Female mouse germ cells form synchronously dividing cysts
Development
1998
, vol. 
125
 (pg. 
3323
-
3328
)
Pepling
ME
Spradling
AC
Mouse ovarian germ cell cysts undergo programmed breakdown to form primordial follicles
Dev Biol
2001
, vol. 
234
 (pg. 
339
-
351
)
Peters
H
Migration of gonocytes into the mammalian gonad and their differentiation
Philos Trans R Soc Lond B Biol Sci
1970
, vol. 
259
 (pg. 
91
-
101
)
Peters
JM
The anaphase-promoting complex: proteolysis in mitosis and beyond
Mol Cell
2002
, vol. 
9
 (pg. 
931
-
943
)
Peters
JM
The anaphase promoting complex/cyclosome: a machine designed to destroy
Nat Rev Mol Cell Biol
2006
, vol. 
7
 (pg. 
644
-
656
)
Pezzi
N
Prieto
I
Kremer
L
Perez Jurado
LA
Valero
C
Del Mazo
J
Martinez
AC
Barbero
JL
STAG3, a novel gene encoding a protein involved in meiotic chromosome pairing and location of STAG3-related genes flanking the Williams–Beuren syndrome deletion
Faseb J
2000
, vol. 
14
 (pg. 
581
-
592
)
Picton
HM
Danfour
MA
Harris
SE
Chambers
EL
Huntriss
J
Growth and maturation of oocytes in vitro
Reprod Suppl
2003
, vol. 
61
 (pg. 
445
-
462
)
Pines
J
Mitosis: a matter of getting rid of the right protein at the right time
Trends Cell Biol
2006
, vol. 
16
 (pg. 
55
-
63
)
Pinsky
BA
Biggins
S
The spindle checkpoint: tension versus attachment
Trends Cell Biol
2005
, vol. 
15
 (pg. 
486
-
493
)
Pittman
DL
Cobb
J
Schimenti
KJ
Wilson
LA
Cooper
DM
Brignull
E
Handel
MA
Schimenti
JC
Meiotic prophase arrest with failure of chromosome synapsis in mice deficient for dmc1, a germline-specific RecA homolog
Mol Cell
1998
, vol. 
1
 (pg. 
697
-
705
)
Polani
PE
Jagiello
GM
Chiasmata, meiotic univalents, and age in relation to aneuploid imbalance in mice
Cytogenet Cell Genet
1976
, vol. 
16
 (pg. 
505
-
529
)
Polani
PE
Crolla
JA
A test of the production line hypothesis of mammalian oogenesis
Hum Genet
1991
, vol. 
88
 (pg. 
64
-
70
)
Prieto
I
Suja
JA
Pezzi
N
Kremer
L
Martinez
AC
Rufas
JS
Barbero
JL
Mammalian STAG3 is a cohesin specific to sister chromatid arms in meiosis I
Nat Cell Biol
2001
, vol. 
3
 (pg. 
761
-
766
)
Rabitsch
KP
Toth
A
Galova
M
Schleiffer
A
Schaffner
G
Aigner
E
Rupp
C
Penkner
AM
Moreno-Borchart
AC
Primig
M
, et al. 
A screen for genes required for meiosis and spore formation based on whole-genome expression
Curr Biol
2001
, vol. 
11
 (pg. 
1001
-
1009
)
Rauh
NR
Schmidt
A
Bormann
J
Nigg
EA
Mayer
TU
Calcium triggers exit from meiosis II by targeting the APC/C inhibitor XErp1 for degradation
Nature
2005
, vol. 
437
 (pg. 
1048
-
1052
)
Reimann
JD
Freed
E
Hsu
JY
Kramer
ER
Peters
JM
Jackson
PK
Emi1 is a mitotic regulator that interacts with Cdc20 and inhibits the anaphase promoting complex
Cell
2001
, vol. 
105
 (pg. 
645
-
655
)
Reis
A
Chang
HY
Levasseur
M
Jones
KT
APCcdh1 activity in mouse oocytes prevents entry into the first meiotic division
Nat Cell Biol
2006
, vol. 
8
 (pg. 
539
-
540
)
Reis
A
Madgwick
S
Chang
HY
Nabti
I
Levasseur
M
Jones
KT
Prometaphase APCcdh1 activity prevents non-disjunction in mammalian oocytes
Nat Cell Biol
2007
, vol. 
9
 (pg. 
1192
-
1198
)
Revenkova
E
Eijpe
M
Heyting
C
Gross
B
Jessberger
R
Novel meiosis-specific isoform of mammalian SMC1
Mol Cell Biol
2001
, vol. 
21
 (pg. 
6984
-
6998
)
Revenkova
E
Eijpe
M
Heyting
C
Hodges
CA
Hunt
PA
Liebe
B
Scherthan
H
Jessberger
R
Cohesin SMC1 beta is required for meiotic chromosome dynamics, sister chromatid cohesion and DNA recombination
Nat Cell Biol
2004
, vol. 
6
 (pg. 
555
-
562
)
Reynaud
K
Fontbonne
A
Marseloo
N
Viaris de Lesegno
C
Saint-Dizier
M
Chastant-Maillard
S
In vivo canine oocyte maturation, fertilization and early embryogenesis: a review
Theriogenology
2006
, vol. 
66
 (pg. 
1685
-
1693
)
Riedel
CG
Katis
VL
Katou
Y
Mori
S
Itoh
T
Helmhart
W
Galova
M
Petronczki
M
Gregan
J
Cetin
B
, et al. 
Protein phosphatase 2A protects centromeric sister chromatid cohesion during meiosis I
Nature
2006
, vol. 
441
 (pg. 
53
-
61
)
Rieder
CL
Schultz
A
Cole
R
Sluder
G
Anaphase onset in vertebrate somatic cells is controlled by a checkpoint that monitors sister kinetochore attachment to the spindle
J Cell Biol
1994
, vol. 
127
 (pg. 
1301
-
1310
)
Roberts
R
Iatropoulou
A
Ciantar
D
Stark
J
Becker
DL
Franks
S
Hardy
K
Follicle-stimulating hormone affects metaphase I chromosome alignment and increases aneuploidy in mouse oocytes matured in vitro
Biol Reprod
2005
, vol. 
72
 (pg. 
107
-
118
)
Rockmill
B
Roeder
GS
Telomere-mediated chromosome pairing during meiosis in budding yeast
Genes Dev
1998
, vol. 
12
 (pg. 
2574
-
2586
)
Roeder
GS
Bailis
JM
The pachytene checkpoint
Trends Genet
2000
, vol. 
16
 (pg. 
395
-
403
)
Romanienko
PJ
Camerini-Otero
RD
The mouse Spo11 gene is required for meiotic chromosome synapsis
Mol Cell
2000
, vol. 
6
 (pg. 
975
-
987
)
Rosenbusch
B
The incidence of aneuploidy in human oocytes assessed by conventional cytogenetic analysis
Hereditas
2004
, vol. 
141
 (pg. 
97
-
105
)
Rosenbusch
BE
Schneider
M
Cytogenetic analysis of human oocytes remaining unfertilized after intracytoplasmic sperm injection
Fertil Steril
2006
, vol. 
85
 (pg. 
302
-
307
)
Ross
LO
Maxfield
R
Dawson
D
Exchanges are not equally able to enhance meiotic chromosome segregation in yeast
Proc Natl Acad Sci USA
1996
, vol. 
93
 (pg. 
4979
-
4983
)
Sagata
N
Watanabe
N
Vande Woude
GF
Ikawa
Y
The c-mos proto-oncogene product is a cytostatic factor responsible for meiotic arrest in vertebrate eggs
Nature
1989
, vol. 
342
 (pg. 
512
-
518
)
Salic
A
Waters
JC
Mitchison
TJ
Vertebrate shugoshin links sister centromere cohesion and kinetochore microtubule stability in mitosis
Cell
2004
, vol. 
118
 (pg. 
567
-
578
)
Saunders
CM
Larman
MG
Parrington
J
Cox
LJ
Royse
J
Blayney
LM
Swann
K
Lai
FA
PLC zeta: a sperm-specific trigger of Ca(2+) oscillations in eggs and embryo development
Development
2002
, vol. 
129
 (pg. 
3533
-
3544
)
Schmidt
A
Duncan
PI
Rauh
NR
Sauer
G
Fry
AM
Nigg
EA
Mayer
TU
Xenopus polo-like kinase Plx1 regulates XErp1, a novel inhibitor of APC/C activity
Genes Dev
2005
, vol. 
19
 (pg. 
502
-
513
)
Schmidt
A
Rauh
NR
Nigg
EA
Mayer
TU
Cytostatic factor: an activity that puts the cell cycle on hold
J Cell Sci
2006
, vol. 
119
 (pg. 
1213
-
1218
)
Shen
Y
Stalf
T
Mehnert
C
De Santis
L
Cino
I
Tinneberg
HR
Eichenlaub-Ritter
U
Light retardance by human oocyte spindle is positively related to pronuclear score after ICSI
Reprod Biomed Online
2006
, vol. 
12
 (pg. 
737
-
751
)
Sherman
SL
Freeman
SB
Allen
EG
Lamb
NE
Risk factors for nondisjunction of trisomy 21
Cytogenet Genome Res
2005
, vol. 
111
 (pg. 
273
-
280
)
Shoji
S
Yoshida
N
Amanai
M
Ohgishi
M
Fukui
T
Fujimoto
S
Nakano
Y
Kajikawa
E
Perry
AC
Mammalian Emi2 mediates cytostatic arrest and transduces the signal for meiotic exit via Cdc20
Embo J
2006
, vol. 
25
 (pg. 
834
-
845
)
Sipek
A
Gregor
V
Horacek
J
Masatova
D
Svetnicova
K
Incidence and survival in children with selected types of congenital defects in the Czech Republic from 1994 to 2000 (Part 1)
Ceska Gynekol
2004
, vol. 
69
 (pg. 
59
-
65
)
Sorensen
RA
Wassarman
PM
Relationship between growth and meiotic maturation of the mouse oocyte
Dev Biol
1976
, vol. 
50
 (pg. 
531
-
536
)
Speed
RM
Meiosis in the fetal mouse ovary
Chromosoma
1982
, vol. 
85
 (pg. 
427
-
437
)
Stemmann
O
Zou
H
Gerber
SA
Gygi
SP
Kirschner
MW
Dual inhibition of sister chromatid separation at metaphase
Cell
2001
, vol. 
107
 (pg. 
715
-
726
)
Stemmann
O
Gorr
IH
Boos
D
Anaphase topsy-turvy: Cdk1 a securin, separase a CKI
Cell Cycle
2006
, vol. 
5
 (pg. 
11
-
13
)
Steuerwald
N
Cohen
J
Herrera
RJ
Sandalinas
M
Brenner
CA
Association between spindle assembly checkpoint expression and maternal age in human oocytes
Mol Hum Reprod
2001
, vol. 
7
 (pg. 
49
-
55
)
Steuerwald
NM
Steuerwald
MD
Mailhes
JB
Post-ovulatory aging of mouse oocytes leads to decreased MAD2 transcripts and increased frequencies of premature centromere separation and anaphase
Mol Hum Reprod
2005
, vol. 
11
 (pg. 
623
-
630
)
Steuerwald
NM
Berm dez
MG
Wells
D
Munn
S
Cohen
J
Maternal age-related differential global expression profiles observed in human oocytes
Reprod BioMed Online
2007
, vol. 
14
 (pg. 
700
-
708
)
Sudakin
V
Chan
GK
Yen
TJ
Checkpoint inhibition of the APC/C in HeLa cells is mediated by a complex of BUBR1, BUB3, CDC20, and MAD2
J Cell Biol
2001
, vol. 
154
 (pg. 
925
-
936
)
Sugiura
K
Su
Y-Q
Diaz
FJ
Pangas
SA
Sharma
S
Wigglesworth
K
O’Brien
MJ
Matzuk
MM
Shimasaki
S
Eppig
JJ
Oocyte-derived BMP15 and FGFs cooperate to promote glycolysis in cumulus cells
Development
2007
, vol. 
134
 (pg. 
2593
-
2603
)
Suh
E-K
Yang
A
Kettenbach
A
Bamberger
C
Michaelis
AH
Zhu
Z
Elvin
JA
Bronson
RT
Crum
CP
McKeon
F
p63 protects the female germ line during meiotic arrest
Nature
2006
, vol. 
444
 (pg. 
624
-
628
)
Sumara
I
Vorlaufer
E
Stukenberg
PT
Kelm
O
Redemann
N
Nigg
EA
Peters
M
The dissociation of cohesin from chromosomes in prophase is regulated by polo-like kinase
Mol Cell
2002
, vol. 
9
 (pg. 
515
-
525
)
Susiarjo
M
Hassold
TJ
Freeman
E
Hunt
PA
Bisphenol A exposure in utero disrupts early oogenesis in the mouse
PLoS Genet
2007
, vol. 
3
 pg. 
e5
 
Swann
K
Saunders
CM
Rogers
NT
Lai
FA
PLCzeta(zeta): a sperm protein that triggers Ca2+ oscillations and egg activation in mammals
Semin Cell Dev Biol
2006
, vol. 
17
 (pg. 
264
-
273
)
Takahashi
T
Saito
H
Hiroi
M
Doi
K
Takahashi
E
Effects of aging on inositol 1,4,5-triphosphate-induced Ca2+release in unfertilized mouse oocytes
Mol Reprod Dev
2000
, vol. 
55
 (pg. 
299
-
306
)
Tanaka
TU
Bi-orienting chromosomes on the mitotic spindle
Curr Opin Cell Biol
2002
, vol. 
14
 (pg. 
365
-
371
)
Tarin
JJ
Perez-Albala
S
Perez-Hoyos
S
Cano
A
Postovulatory aging of oocytes decreases reproductive fitness and longevity of offspring
Biol Reprod
2002
, vol. 
66
 (pg. 
495
-
499
)
Tatone
C
Carbone
MC
Gallo
R
Delle Monache
S
Di Cola
M
Alesse
E
Amicarelli
F
Age-associated changes in mouse oocytes during postovulatory in vitro culture: possible role for meiotic kinases and survival factor BCL2
Biol Reprod
2006
, vol. 
74
 (pg. 
395
-
402
)
Terret
ME
Ferby
I
Nebreda
AR
Verlhac
MH
RINGO efficiently triggers meiosis resumption in mouse oocytes and induces cell cycle arrest in embryos
Biol Cell
2001
, vol. 
93
 (pg. 
89
-
97
)
Thomas
SE
Soltani-Bejnood
M
Roth
P
Dorn
R
Logsdon
JJM
McKee
BD
Identification of two proteins required for conjunction and regular segregation of achiasmate homologs in drosophila male meiosis
Cell
2005
, vol. 
123
 (pg. 
555
-
568
)
Tornell
J
Billig
H
Hillensjo
T
Regulation of oocyte maturation by changes in ovarian levels of cyclic nucleotides
Hum Reprod
1991
, vol. 
6
 (pg. 
411
-
422
)
Toth
A
Rabitsch
KP
Galova
M
Schleiffer
A
Buonomo
SB
Nasmyth
K
Functional genomics identifies monopolin: a kinetochore protein required for segregation of homologs during meiosis I
Cell
2000
, vol. 
103
 (pg. 
1155
-
1168
)
Tsafriri
A
Cao
X
Ashkenazi
H
Motola
S
Popliker
M
Pomerantz
SH
Resumption of oocyte meiosis in mammals: on models, meiosis activating sterols, steroids and EGF-like factors
Mol Cell Endocrinol
2005
, vol. 
234
 (pg. 
37
-
45
)
Tsurumi
C
Hoffmann
S
Geley
S
Graeser
R
Polanski
Z
The spindle assembly checkpoint is not essential for CSF arrest of mouse oocytes
J Cell Biol
2004
, vol. 
167
 (pg. 
1037
-
1050
)
Tung
JJ
Hansen
DV
Ban
KH
Loktev
AV
Summers
MK
Adler
JR
3rd
Jackson
PK
A role for the anaphase-promoting complex inhibitor Emi2/XErp1, a homolog of early mitotic inhibitor 1, in cytostatic factor arrest of Xenopus eggs
Proc Natl Acad Sci USA
2005
, vol. 
102
 (pg. 
4318
-
4323
)
Uhlmann
F
The mechanism of sister chromatid cohesion
Exp Cell Res
2004
, vol. 
296
 (pg. 
80
-
85
)
Uhlmann
F
Lottspeich
F
Nasmyth
K
Sister-chromatid separation at anaphase onset is promoted by cleavage of the cohesin subunit Scc1
Nature
1999
, vol. 
400
 (pg. 
37
-
42
)
Uhlmann
F
Wernic
D
Poupart
MA
Koonin
EV
Nasmyth
K
Cleavage of cohesin by the CD clan protease separin triggers anaphase in yeast
Cell
2000
, vol. 
103
 (pg. 
375
-
386
)
van Montfrans
JM
van Hooff
MHA
Martens
F
Lambalk
CB
Basal FSH, estradiol and inhibin B concentrations in women with a previous Down’s syndrome affected pregnancy
Hum Reprod
2002
, vol. 
17
 (pg. 
44
-
47
)
Verlhac
MH
Kubiak
JZ
Clarke
HJ
Maro
B
Microtubule and chromatin behavior follow MAP kinase activity but not MPF activity during meiosis in mouse oocytes
Development
1994
, vol. 
120
 (pg. 
1017
-
1025
)
Verlhac
MH
Kubiak
JZ
Weber
M
Geraud
G
Colledge
WH
Evans
MJ
Maro
B
Mos is required for MAP kinase activation and is involved in microtubule organization during meiotic maturation in the mouse
Development
1996
, vol. 
122
 (pg. 
815
-
822
)
Verlhac
M-H
Lefebvre
C
Guillaud
P
Rassinier
P
Maro
B
Asymmetric division in mouse oocytes: with or without Mos
Curr Biol
2000
, vol. 
10
 (pg. 
1303
-
1306
)
Vialard
F
Petit
C
Bergere
M
Gomes
DM
Martel-Petit
V
Lombroso
R
Ville
Y
Gerard
H
Selva
J
Evidence of a high proportion of premature unbalanced separation of sister chromatids in the first polar bodies of women of advanced age
Hum Reprod
2006
, vol. 
21
 (pg. 
1172
-
1178
)
Vialard
F
Lombroso
R
Bergere
M
Gomes
DM
Hammoud
I
Bailly
M
Selva
J
Oocyte aneuploidy mechanisms are different in two situations of increased chromosomal risk: older patients and patients with recurrent implantation failure after in vitro fertilization
Fertil Steril
2007
, vol. 
87
 (pg. 
1333
-
1339
)
Vinot
S
Le
T
Maro
B
Louvet-Vallee
S
Two PAR6 proteins become asymmetrically localized during establishment of polarity in mouse oocytes
Curr Biol
2004
, vol. 
14
 (pg. 
520
-
525
)
Visser
JA
Themmen
APN
Anti-Mullerian hormone and folliculogenesis
Mol Cell Endocrinol
2005
, vol. 
234
 (pg. 
81
-
86
)
Waizenegger
IC
Hauf
S
Meinke
A
Peters
JM
Two distinct pathways remove mammalian cohesin from chromosome arms in prophase and from centromeres in anaphase
Cell
2000
, vol. 
103
 (pg. 
399
-
410
)
Walker
MY
Hawley
RS
Hanging on to your homolog: the roles of pairing, synapsis and recombination in the maintenance of homolog adhesion
Chromosoma
2000
, vol. 
109
 (pg. 
3
-
9
)
Wang
JY
Lei
ZL
Nan
CL
Yin
S
Liu
J
Hou
Y
Li
YL
Chen
DY
Sun
QY
RNA interference as a tool to study the function of MAD2 in mouse oocyte meiotic maturation
Mol Reprod Dev
2007
, vol. 
74
 (pg. 
116
-
124
)
Warren
WD
Gorringe
KL
A molecular model for sporadic human aneuploidy
Trends Genet
2006
, vol. 
22
 (pg. 
218
-
224
)
Wassmann
K
Niault
T
Maro
B
Metaphase I arrest upon activation of the Mad2-dependent spindle checkpoint in mouse oocytes
Curr Biol
2003
, vol. 
13
 (pg. 
1596
-
1608
)
Watanabe
Y
Nurse
P
Cohesin Rec8 is required for reductional chromosome segregation at meiosis
Nature
1999
, vol. 
400
 (pg. 
461
-
464
)
Webb
R
Nicholas
B
Gong
JG
Campbell
BK
Gutierrez
CG
Garverick
HA
Armstrong
DG
Mechanisms regulating follicular development and selection of the dominant follicle
Reprod Suppl
2003
, vol. 
61
 (pg. 
71
-
90
)
Weiss
E
Winey
M
The Saccharomyces cerevisiae spindle pole body duplication gene MPS1 is part of a mitotic checkpoint
J Cell Biol
1996
, vol. 
132
 (pg. 
111
-
123
)
Wilhelm
D
Palmer
S
Koopman
P
Sex determination and gonadal development in mammals
Physiol Rev
2007
, vol. 
87
 (pg. 
1
-
28
)
Wolf
K
How meiotic cells deal with non-exchange chromosomes
BioEssays
1994
, vol. 
16
 (pg. 
107
-
114
)
Wolstenholme
J
Angell
RR
Maternal age and trisomy–a unifying mechanism of formation
Chromosoma
2000
, vol. 
109
 (pg. 
435
-
438
)
Wright
WE
Piatyszek
MA
Rainey
WE
Byrd
WJWS
Telomerase activity in human germline and embryonic tissues and cells
Dev Genet
1996
, vol. 
18
 (pg. 
173
-
179
)
Xu
H
Beasley
MD
Warren
WD
van der Horst
GTJ
McKay
MJ
Absence of mouse REC8 cohesin promotes synapsis of sister chromatids in meiosis
Dev Cell
2005
, vol. 
8
 (pg. 
949
-
961
)
Yin
S
Wang
Q
Liu
JH
Ai
JS
Liang
CG
Hou
Y
Chen
DY
Schatten
H
Sun
QY
Bub1 prevents chromosome misalignment and precocious anaphase during mouse oocyte meiosis
Cell Cycle
2006
, vol. 
5
 (pg. 
2130
-
2137
)
Yokobayashi
S
Watanabe
Y
The kinetochore protein Moa1 enables cohesion-mediated monopolar attachment at meiosis I
Cell
2005
, vol. 
123
 (pg. 
803
-
817
)
Yu
J
Deng
M
Medvedev
S
Yang
J
Hecht
NB
Schultz
RM
Transgenic RNAi-mediated reduction of MSY2 in mouse oocytes results in reduced fertility
Dev Biol
2004
, vol. 
268
 (pg. 
195
-
206
)
Yuan
L
Liu
J-G
Hoja
M-R
Wilbertz
J
Nordqvist
K
Hoog
C
Female germ cell aneuploidy and embryo death in mice lacking the meiosis-specific protein SCP3
Science
2002
, vol. 
296
 (pg. 
1115
-
1118
)
Zachariae
W
Schwab
M
Nasmyth
K
Seufert
W
Control of cyclin ubiquitination by CDK-regulated binding of Hct1 to the anaphase promoting complex
Science
1998
, vol. 
282
 (pg. 
1721
-
1724
)
Zhang
D
Ma
W
Li
YH
Hou
Y
Li
SW
Meng
XQ
Sun
XF
Sun
QY
Wang
WH
Intra-oocyte localization of MAD2 and its relationship with kinetochores, microtubules, and chromosomes in rat oocytes during meiosis
Biol Reprod
2004
, vol. 
71
 (pg. 
740
-
748
)
Zickler
D
Kleckner
N
The leptotene-zygotene transition of meiosis
Annu Rev Genet
1998
, vol. 
32
 (pg. 
619
-
697
)
Zou
H
McGarry
TJ
Bernal
T
Kirschner
MW
Identification of a vertebrate sister-chromatid separation inhibitor involved in transformation and tumorigenesis
Science
1999
, vol. 
285
 (pg. 
418
-
422
)
The online version of this article has been published under an open access model. Users are entitled to use, reproduce, disseminate, or display the open access version of this article for non-commercial purposes provided that: the original authorship is properly and fully attributed; the Journal and Oxford University Press are attributed as the original place of publication with the correct citation details given; if an article is subsequently reproduced or disseminated not in its entirety but only in part or as a derivative work this must be clearly indicated. For commercial re-use, please contact journals.permissions@oxfordjournals.org