Skip to main content

In vitro fertilization (IVF) in mammals: epigenetic and developmental alterations. Scientific and bioethical implications for IVF in humans

Abstract

The advent of in vitro fertilization (IVF) in animals and humans implies an extraordinary change in the environment where the beginning of a new organism takes place. In mammals fertilization occurs in the maternal oviduct, where there are unique conditions for guaranteeing the encounter of the gametes and the first stages of development of the embryo and thus its future. During this period a major epigenetic reprogramming takes place that is crucial for the normal fate of the embryo. This epigenetic reprogramming is very vulnerable to changes in environmental conditions such as the ones implied in IVF, including in vitro culture, nutrition, light, temperature, oxygen tension, embryo-maternal signaling, and the general absence of protection against foreign elements that could affect the stability of this process. The objective of this review is to update the impact of the various conditions inherent in the use of IVF on the epigenetic profile and outcomes of mammalian embryos, including superovulation, IVF technique, embryo culture and manipulation and absence of embryo-maternal signaling. It also covers the possible transgenerational inheritance of the epigenetic alterations associated with assisted reproductive technologies (ART), including its phenotypic consequences as is in the case of the large offspring syndrome (LOS). Finally, the important scientific and bioethical implications of the results found in animals are discussed in terms of the ART in humans.

Introduction

Human beings have the capacity to modify the environment and in this way to influence the development and survival of animal species and human beings. In this perspective, one topic that has had increasing importance is the impact that modifications of the environment have on early stages of mammalian development, which are particularly vulnerable to environmental changes.

In 1999 Barker et al. described the relation of maternal malnutrition during pregnancy and the threat of developing certain diseases in adulthood [1]. Greater risk of coronary disease, hypertension, type two diabetes, metabolic syndrome and others have been described. This has been called the fetal origins of adult diseases or the Barker hypothesis [2, 3]. The transcendence of this discovery has been named with the initials DOHaD (Developmental Origins of Health and Disease) [4, 5]. Soon afterwards it was hypothesized that this effect could also occur in the pre-implantation embryo [68]. This was confirmed in several studies in rats that showed that malnutrition and hypoproteic diets administered only during the pre-implantation stage resulted in altered development such as low birth weight and abnormal blood pressure [912]. These findings were very important to stimulate research on the influence that assisted reproductive technologies (ART) could have on development and epigenetic reprogramming in the pre-implantation period, as these techniques imply great changes in the environment [13, 14]. The major concerns are related to the possible effects that ART may have on normal development in humans. It has been calculated that in developed countries 1–3 % of children are conceived using these techniques [15].

Fertilization of eutherian animals occurs in the maternal oviduct. This is the natural and unique environment to achieve the necessary requirements for embryo life and its early and late development. The embryo conceived in vitro is manipulated and cultured in very different conditions [11, 16].

Conrad Waddington highlighted many decades ago the relevance of the environment in development [17]. He emphasized the importance of studying the conditions that control development that mediate the interactions between genotype and phenotype. Genetics had discovered the laws of inheritance and had explained how different characters are transmitted from parents to offspring. But, Waddington underlined, that there wasn’t much knowledge about the mechanisms of development. He named this process Epigenetics, which is now understood as the conditions that control the expression of genes that are highly influenced by environment. A more precise definition is: “The study of changes in gene function that are mitotically and/or meiotically heritable and that do not entail a change in DNA sequence” [18]. This concept has broadened biological research in order to understand how this process can be altered by the environment and thus influence normal development and impact the etiology, susceptibility and onset of adult diseases [1924]. Epigenetics has demonstrated that normal development does not depend only on a healthy genome.

There is an important epigenetic reprogramming during gametogenesis and the preimplantational period of the embryo, especially in imprinted genes, defined by their parental origin [25, 26]. This period has a sensible window to environmental changes which can alter the process of reprogramming, and thereby affect survival and the early and late development of mammal embryos [27, 28]. The discovery of the epigenetic mechanisms that control gene expression, at a molecular level, has been very useful to understand and detect when there has been an alteration at this level [20, 29]. Some of the mechanisms known today are: cytosine-adenine methylation; histone modifications and the control that different microRNA has on gene expression [24, 29, 30]. In the mid-1990s Sasaki et al., and Tremblay et al. showed that rats produced in vitro carried imprinted gene alterations, especially in the H19 paternal gene [31, 32]. Today we have a huge amount of information, mostly derived from animals, about the alterations that can occur in embryos produced in vitro because of the manipulation and artificial environmental conditions associated with these techniques [33].

It has been discovered that the following processes and techniques associated with in vitro fertilization (IVF) can alter the epigenetic reprogramming of gametes, embryos and normal mammal development [34] (Fig. 1).

Fig. 1
figure 1

Schematic representation of in vivo and in vitro steps of mammalian fertilization. I Ovarian hormonal stimulation promotes follicle maturation and then ovulation. II ART uses different fertilization techniques to achieve fertilization, while in vivo female and male gametes interact and fuse in the female tract (infundibulum). III After fertilization, the preimplantation embryo spends a time under cultured conditions that may affect its further development. IV During this period of time, different techniques such as preimplantational embryo biopsies can be applied. V Finally, the in vitro produced embryo is transfered to a recipient female. On the other hand, the developing embryo moves towards the uterus interacting with the female reproductive system in a optimal environment

  1. 1.

    Superovulation and in vitro maturation (IVM) of gametes.

  2. 2.

    Fertilization technique: ICSI vs IVF.

  3. 3.

    Embryo culture.

  4. 4.

    Embryo manipulation: pre-implantation genetic diagnosis and embryo transfer.

  5. 5.

    Maternal-embryo signaling during the pre-implantation period.

The aim of this review is to update the currently available information provided by animal studies exploring offspring alterations in the epigenetic profile, development, survival and phenotype associated with the artificial environment in which IVF is performed. We propose that this information has scientific and bioethical implications that must be considered in human IVF/ICSI.

Superovulation and IVM of gametes

Superovulation

In women the ovary normally produces one unique egg in each cycle. To count on more than one egg to perform IVF, the superovulation helps make the process more productive. This technique consists of hormonal stimulation of the ovary with exogenous gonadotropins or similar substances, to produce a larger number of eggs per cycle. The eggs obtained in this mode are usually immature and require an in vitro time of maturation.

Over the last 10 years there has been a growing body of evidence showing the relationship of superovulation and epigenetic disorders on eggs and embryos, which are associated with developmental alterations in different mammal species [22, 27, 35]. The analysis of the genomic imprinting of eggs obtained from super-ovulation shows disorders in four imprinted genes: Peg1, Kcnq1ot1, Zac and H19 in comparison with eggs from natural ovulation [3640]. Specifically, alterations in gene H19 have been demonstrated in many studies [41, 42]. The hormone dose used for this procedure seems to be important in the degree of methylation [39].

Studies on naturally conceived versus super ovulation-derived mice offspring demonstrated epigenetic alterations in the somatic tissue of the developing embryo [25]. One recent study demonstrated that epigenetic alterations associated with gonadotropins also correlate with less fetal and placental development, as well as a smaller embryo throughout life [43]. Studies in mice suggest that ovary hyper-stimulation may affect embryo implantation [44].

This significant amount of information related to the deleterious effects of superovulation in normal gene expression, which impacts various issues necessary for embryo development [45, 46] has enhanced efforts to improve the superovulation protocols in animals to diminish its negative effects [47].

IVM of gametes

In vitro maturation of eggs has been associated with epigenetic alterations in addition to the effect of superovulation [4850]. One study analyzed the DNA methylation status of the imprinted genes H19, Mest/Peg1 and Igf2R during in vitro maturation of mouse oocyte from preantral follicles [50]. The results show that when germinal vesicle oocyte after IVM are compared to those isolated from mice ovary, a loss of methylation at the Igf2R locus and Mest/Peg1 locus, and a gain of methylation at the H19 locus were found.

Similar results have been found in humans. In one of such studies, 20 metaphase II oocytes were analyzed, and 15 showed the normal unmethylated maternal pattern of H19 gene, while five originating from two different patients exhibited a methylated pattern [48].

It has been shown that the level of alteration depends on the time and the composition of culture media [39, 5153].

Sperm culture has not been associated with epigenetic alterations, probably because epididymal sperm have the epigenetic reprogramming already completed, in contrast to egg maturation. The eventual epigenetic alterations that may be found in sperm have been associated with male infertility, as shown in mice [54, 55], swine [56] and humans [46, 5759]. But a new challenge is presented when in animal and human ICSI is performed with immature sperm from the testes. Normal mice have been born with this technique [60]. After animal experiments, spermatozoa with different degrees of maturation have been used in humans with controversial results and low pregnancy rates. In vitro maturation of the sperm has also been tested. The in vitro culture of spermatids has also resulted in very poor outcome. Furthermore, if the sperm sample is maintained for an extended period of time there may be additional damage due to DNA fragmentation by nuclease release [61].

Egg epigenetic alterations associated with super-ovulation and in vitro culture can be maintained during the embryo and placental period [62, 63]. Authors recognize the difficulty in distinguishing the impact of super-ovulation versus the in vitro culture of eggs in these results.

In conclusion, there is evidence that superovulation as well as in vitro egg culture in different animal species produces epigenetic alterations in the egg and embryo, and that this could affect the outcome of the pregnancy.

Technique of fertilization: ICSI vs IVF

There is evidence showing that the kind of IVF technique used can alter the epigenetic reprogramming and eventually development [64]. Although in regular IVF there is a selection of gametes and embryos of better quality, at present the prevalent and more economically efficient technique used in humans, is ICSI.

Intra-cytoplasmatic sperm injection (ICSI)

ICSI was introduced in 1992 with great success but without previous experimental testing. Animal models were considered unsuitable. This technique bypasses several physiological events: natural selection of the fertilizing sperm, sperm capacitation, acrosomic reaction and membrane fusion [65, 66]. The use of this technique in humans is increasing, and is used even as a standard method in cases with normal sperm [67]. Data provided from 56 reporting countries shows an increase of ICSI in all initiated cycles from 60.6 % in 2004 to 66 % in 2006 (96 % in the Middle East; 81 % in Latin America; 70 % in the USA) [68, 69]. The effects of ICSI in humans have been studied for decades, including implantation percentage, live-born, incidence of malformation and developmental disorders [7073]. But since research specifically on epigenetic effects implies embryo destruction, it has not been performed in humans because of ethical reasons, underlining the importance of animal studies.

Studies on different animal species produced by ICSI, have found an asynchronous remodeling of chromatin decondensation of the male pronucleus in primates, [71, 74] mice and cattle [75, 76]. Mice produced by ICSI compared to those produced by regular IVF have long-lasting transcriptome disturbances that are maintained until the neonatal stage. But up to date these alterations have not correlated with changes in the phenotypic profile or with transgenerational effects [77]. It has been described that mouse ICSI blastocysts, compared to in vivo conceived groups, have a reduction in the inner mass cells and significant differences in gene expression related to cell function, development and metabolism [78]. However, there is one study that did not find any differences in pre-implantation development in IVF or ICSI-produced mice compared to naturally conceived mice. It is thought that the difference in the studies may be explained because of the protocol used for ICSI in different species and the different strategies used to activate the egg and guarantee embryo development [79, 80].

It has been observed in mice offspring produced by ICSI that DNA fragmentation is associated with more abnormal development, health and behavior [61]. That is why there is so much effort in selecting non-fragmented DNA sperm to have better results [81]. Finally, descendants of male mice produced by ICSI have a higher rate of apoptosis in their testicle sperm, which requires further research [82].

In conclusion, in animals, ICSI is a low efficiency technology and shows more epigenetic alterations in comparison with regular IVF. Thus its application in animals in contrast of what happens in human, is less widespread [80].

Plasticware exposure

Notably, in 2003 Hunt et al. [83] communicated that exposing female mice to the xenoestrogen Bisphenol A (BPA), meiosis was altered and the rate of aneuploidy increase in these animals. Other studies demonstrated that exposing mice to BPA at the implantation period produced alterations in uterine morphology, estrogen and progesterone receptors and a lower implantation percentage [8487]. This compound, among other xenoestrogens, is used in plastic and epoxy resin production [88]. IMV, IVF, ICSI or embryo culture implies exposition to plasticware and different studies have shown leakage of xenoestrogens, phthalates and from plastic polymers [89]. Since many in vivo and in vitro studies have shown that these compounds induce epigenetic alterations such as cytosine methylation and/or histone acetylation [90, 91], the exposure of gametes and embryos to plasticware during in vitro manipulation may induce epigenetic abnormalities which will may affect further development.

This information motivated research on whether there are significant levels of chemical compounds in IVF, especially BPA, which could affect the reproductive process. In this review we did not find any animal studies addressing this aspect. There is one study on women that did not find detectable BPA either in the culture medium or in the material used for IVF [92].

Embryo culture

Culture media and nutrition

In vitro culture (IVC) is probably the most relevant factor in the alterations of epigenetic reprogramming and development of animal embryos produced by IVF. This process is obligatory for IVF in contrast to some of the aspects mentioned above. Since the 1980s several studies have focused on investigating the effects of culture media in implantation, survival and development in different animal species, and the influence of several modifications in culture media such as the effects of proteins, the quality of the water and serum [9399]. An increasing number of studies confirm the influence of the culture medium in epigenetic preimplantation reprogramming and its impact on early embryo development [100102].

Various effects associated with in vitro embryo culture can be observed early in the period from fertilization to implantation, such as:

  • Low implantation rate.

  • Disturbances in development speed, embryo quality and low trophoblast development.

  • Abnormal preimplantation epigenetic reprogramming.

It has been shown that suboptimal culture media affect the percentage of implantation and the survival of embryos that could achieve implantation [103]. It has been shown in cattle and other species that there is an association between the early timing of the first cleavage and the probability of reaching the blastocyst stage. This fact is thought to be related with the suboptimal IVC [104106]. Gutierrez-Adan et al. also detected differences in the mRNA pattern and development speed between embryos produced in vitro and in vivo [107]. They hypothesized that this epigenetic pattern could also be a good system for the selection of better quality embryos [108, 109].

The suboptimal environment of IVC media also affects the development of the trophoblast [108]. In mice, the dysregulation of the trophoblast epigenetic profile is maintained in the placenta, which seems to be more sensitive than the embryo to IVC [110, 111].

The results of different studies clearly demonstrate that the above alterations are associated with different degrees of variation in the epigenetic profile of the embryo and placenta, and have an impact on development [112]. Of special importance are the modifications of imprinted genes such as gene H19, which has been widely studied because of the different expressions it has depending on the culture medium [32, 33, 47, 113, 114].

Later, after implantation various deleterious effects have been found in the fetal period (implantation to birth). These include:

  • Unbalanced fetal-placental development

  • Abnormal fetal growth

  • Abnormal metabolic responses

In some cases these changes have been associated with epigenetic modifications in the preimplantation period, especially of imprinted genes that play a key role in fetal and placental development as has been indicated [115]. Recent studies have confirmed these findings, showing cellular aberrations in placenta and fetus linked to changes in gene expression, and the association of epigenetic aberrations with glucose metabolism and fetal growth in mice [116, 117]. This could affect the critical role that imprinted genes have in growth and development of the fetus and adult. Placental abnormalities and a greater mortality rate have also been reported [6].

The effect of impaired nutrition by maternal diet and in vitro culture is associated with developmental and metabolic alterations [118, 119]. Glucose metabolism is found to be altered in response to suboptimal culture conditions [120]. Specific alterations in glucose metabolism have also been also found to be different between male and female embryos [116]. Abnormal development of skeletal muscle in bovine fetuses associated with a decreased expression of mRNA for myostatin has been reported in embryos produced in vitro in comparison with in vivo [121]. These anomalies are not detected in the implantation period, but in later stages of development.

Effects in the postnatal and adult period have been described in relation to IVC. For decades it has been described that ruminants born by IVF have an abnormally large birth size and visceromegaly [122, 123]. The first and most relevant alteration in phenotype in animals produced by IVF is the Large Offspring Syndrome (LOS) [124]. It is characterized by large size at birth, gross abnormalities in different organs, mainly visceromegaly, and metabolic alterations, especially in the glucose-insulin system, hypoglycemia, large tongue and umbilical hernia. All these features are similar to those found in the Beckwith–Wiedemann (BWS) syndrome in humans [125]. It is of special interest that the epigenetic alterations in the LOS are very similar to those found in the BWS [126128]. The phenotypic similarity of these syndromes has helped to understand the epigenetic alteration of the BWS. In fact one study has found in LOS the absence of methylation at the KvDMR1 on the maternal allele, which is the major molecular signature of BWS [125]. The relation between epigenetic defects in LOS and BWS is very complex and needs more study. Studies with the LOS bovine model seem to be a good approach [125].

Embryo culture itself, independent of the effects of embryo transfer and of fetal growth, is associated with higher systolic blood pressure in 21-week old mice compared to in vivo controls. This study also found elevated activity of serum angiotensin and hepatic enzymes involved in the control of gluconeogenesis [129].

The addition of serum to the embryo culture medium has been associated with abnormal skeleton and organ development [121]. In vitro culture of rat embryos results with little but significant alterations related to anxiety, psychomotor activity and special memory. Investigators propose that it is possible that a hippocampal alteration coupled to an impaired brain connection could alter memory development [130]. The type of culture medium also has a certain effect on the degree of these alterations in behavior [130]. Interestingly, a recent article reported that there is a synergistic effect of IVC with the type of diet given prenatally and postnatally in relation to its effect on anxiety and behavior. When these two parameters were evaluated separately there was no statistical difference. The authors point out that these findings highlight the importance of diet in women undergoing IVF [131].

The impact of culture medium on the outcome of ART is today undoubtedly a major constraint for these techniques in mammalian species. Thus special attention has been paid first to search for the best conditions of culture medium that can minimize its deleterious effects on epigenetic reprogramming and development [132138]. And, secondly to research how to detect major alterations in the epigenetic profile of mammalian embryos produced by IVF [139141]. In these latter aspects some authors have highlighted the importance of modifications in the expression of gene H19 that may be used as a sensor to the epigenetic embryo quality. Gene H19 has a central role in the control of imprinting genes [108]. Although there are differences between species, studies in animals may contribute to understand what happens in human [142, 143].

Oxygen concentration effects

Oxygen concentration in the oviduct where natural fertilization takes place in various mammalian species is between 1 and 9 %, which corresponds to an oxygen tension of around 11–60 mmHG [144146]. In the beginning, human embryos produced by IVF were cultured at oxygen concentrations near 20 % (the same as the atmospheric level), very different than the in vivo condition [147, 148]. Studies in mice suggested that in vitro concentrations of oxygen similar to atmosphere levels could produce oxidative stress, mediated by free radicals of O2, and affect embryogenesis [149152]. Today we have a large amount of data showing that culture of embryos at elevated O2 concentrations impairs blastocyst development, cell number and embryo metabolism in a variety of species [153156]. Consistent with this, other studies have shown that culture at low oxygen pressure, around 5 %, help to produce better quality embryos in mice and cattle [157160]. These deleterious effects were observed similarly during the cleavage and post-compaction phases [152]. Atmospheric oxygen concentration correlates with an increase of reactive oxygen species (ROS) compared to culture at 5 % oxygen [155]. ROS can alter protein synthesis and function and lipids, affecting cell membrane stability and DNA damage [161]. Experiments adding antioxidants to culture have had some effect in reducing oxidative stress [162165].

Human studies have also shown the correlation between O2 pressures and IVF results [166, 167]. The Cochrane Data Base review confirmed the results of different studies in that the IVC of human embryos under conditions of low oxygen concentration improves the outcome of IVF and ICSI [168].

Embryo manipulation: pre-implantation genetic diagnosis and embryo transfer

Pre-implantation genetic diagnosis (PGD)

This procedure has the aim to study the genetic condition of the embryo before implantation [169]. One or two blastomeres are extracted and genetically analyzed to perform this procedure.

There are studies in different species, especially cattle, to evaluate the effect of PGD [170]. PGD is performed in the embryo-transfer industry of animals for efficient sexing of the preimplantation bovine embryo in order to control the sex of the offspring [171]. The best technique to perform the biopsy with minimum damage to the embryo has been evaluated in mice and bovines [172, 173]. To assess the consequences of PGD on embryo and placental development a comparison between groups of biopsied mice versus non-biopsied control groups has been performed [174176]. It was demonstrated with time-lapse videos that biopsied embryos had a delay from one embryo stage to another. This may probably be due to later compaction and hatching from the zona pellucida [177, 178].

The effects of blastomer biopsy on steroid metabolism have also been investigated in mice. The results demonstrated that mice born after this procedure had a lower birth weight and a deregulation of steroid metabolism, which could have severe effects on posterior development [175].

Recent studies that investigated short and long term effects on mouse development found behavioral disorders with three different tests (Morris, water maze and pole climbing tests), compared to control groups. These disorders may occur because of altered epigenetic patterns in the mouse brain [179181].

Embryo transfer

Rivera et al. in a meticulous study isolated the effects of embryo transfer on mice independently of other factors [182]. To do this they studied the methylation profile of ten imprinted genes (H19, Snrpn, Igf2, Kcnq1ot1, Cdkn1c, Kcnq1, Mknr3, Ascl2, Zim1, Peg3). A control group of embryos was conceived in vivo, not cultured or transferred. The female mice were sacrificed on day 9.5 and the embryos immediately collected for epigenetic study (unmanipulated group). A second group, conceived in vivo, was extracted at the blastocyst stage and transferred after an hour and a half without passing through in vitro culture. Concepti were collected and processed as described in the control group (embryo transfer group). A third group of embryos conceived in vivo was extracted at the two blastomere stage and cultured in vitro for 2 or 3 days until blastocyst and then transferred. Concepti were collected and processed as described in the control group (embryo culture + transfer). The epigenetic profile was studied in the three groups at nine and a half days. The embryo transfer group conceived in vivo that was transferred without going through culture had an aberrant expression of imprinted genes compared to the control group. Furthermore, in the embryo cultures + transfer, the effects of transfer was increased by culture as shown by the number of genes with aberrant allelic expression in embryonic and extraembryonic tissues. Alterations in the imprinting pattern were more significant in the placenta and yolk sac. Interestingly, they found that biallelic expression of Kcnq1ot1 is related to loss of methylation on the maternal allele of the KvDMR1 locus, which has been often associated with the human syndrome Beckwith–Wiedemann (BWS). These data shows that the sole embryo manipulation induces aberrant methylation gene pattern.

Maternal-embryo signaling during the pre-implantation period

In mammals, during the pre-implantation stage there is an exchange of different types of signals between the mother and the embryo that is thought to be critical for embryo development and implantation [183].

The presence of the embryo is recognized through these signals and a cross-talk takes place in the maternal tract which prepares an appropriate environment for implantation [184186]. Thus, successful pregnancy in mammals involves synchronization between a receptive endometrium and a viable embryo [183, 187, 188]. This includes the modulation of the maternal immune system by the embryo [189, 190]. It is difficult to study these interactions in vivo; so much of what is known comes from in vitro models.

Absence of these signals in IVF raises the question of how much the technique can affect the results [191, 192]. In bovines it has been considered that the absence of these signals could be an important factor in the low efficiency of IVF [193]. Most recent reviews indicate that while there is considerable evidence of the influence of the oviduct on the quality of the developing embryo, there is little evidence of signals from the embryo to the oviduct or the endometrium in this early stage [109, 194197].

By contrast, strong evidence exists in relation to the importance of the embryo/endometrium interaction for normal implantation and pregnancy development [198, 199]. Nevertheless new research has shown that the embryo has an effect on the regulation of the epithelial cells of the oviduct [200].

All this information addresses the fact of the key role of the oviduct and maternal embryo cross-talk in embryo development and implantation [201]. New genomic technologies open the horizon for future research in this aspect of pre-implantation development [184].

Several modifications to the technique to overcome this problem have been proposed, such as the use of intermediate host oviducts for IVF embryos or adding synthetic conditions similar to the oviduct liquid [202204]. One vital factor of this exchange of signals for pre-implantation development is TGF- β [205].

The specific role of the environmental factors that we have discussed above in development and phenotype alterations are still not fully understood and difficult to isolate. The results of this artificial environment as a whole can be understood as the embryo being under different kinds of stress.

Trans-generational inheritance

A transgenerational effect occurs when the alterations pass through several generations. Evidence shows that epigenetic imprinting could be trans-generationally transmitted, despite the fact that during gametogenesis there is epigenetic reprograming; but some molecular epigenetic elements not related to DNA sequence can resist this reprograming [206]. This has been studied in animals [206]. Some facts related to this “resistance” to epigenetic reprogramming are described in the large offspring syndrome (LOS), such as organomegaly, which could pass to a second generation of individuals [16]. In cryopreserved rabbit embryos it was demonstrated that IVF could alter female reproduction in the next generation [207]. A recent study in mice showed that suboptimal IVC was associated with transgenerational alterations of glucose metabolism and hepatomegaly in the male offspring [208].

Some relevant elements in epigenetic mechanisms related to this process are the following:

  • Some specific types of transposons, which are resistant to post-fertilization demethylation, such as the intracisternal-A particles (IAPs,) long-terminal repeat retrotransposons [209].

  • De novo mutations in the DNA sequence of preimplantation embryos such as: numerical and structural chromosomal abnormalities, point mutations, copy number variant (CNV) changes and duplications/deletions of microsatellites [210]

  • Histone and chromatin modification [211]. Very recent data generated using Caenorhabditis elegans provide evidence for transmission of male gamete-mediated chromatin states through several rounds of replication, although it remains to be seen if a mechanism such as this is conserved in mammals [212].

  • There are various cytoplasmic RNAs, such as nc(non-coding) RNAs which are transmitted by germ cells and are essential for post-fertilization development [213].

Discussion

The general objective of animal reproduction research is mainly to improve breeding, reproductive and productive efficiency in harmony with ecological challenges and animal health [214]. For this purpose one of the essential topic is to study the genetic and epigenetic mechanisms that control the preimplantation embryo development in vivo and in vitro.

From another perspective it must be pointed out that animal studies are important to evaluate the safety of new drugs and treatments as a previous bioethical step before clinical trials or medical innovations are performed in humans [215].

The focus of this review was to update the scientific information concerning the epigenetic alterations produced by the techniques of in vitro manipulation related to IVF in animals and their impact in phenotype, behavior and transgenerational transmission. A substantial amount of research available in animals shows convincingly that ART particularly IVF, produces significant epigenetic modifications, altering the expression of different genes, particularly of imprinted genes that have a major role in normal animal development. The consequences of these perturbations on the outcome of the offspring raises important questions not completely resolved that need to be addressed in order to clarify them. On one hand, it is quite difficult to isolate the degree of influence of the different factors implicated in the ART technique, such as: Superovulation and IVC of gametes, ICSI, embryo culture and absence of maternal embryo-signaling. The existing information centers especially on the importance of embryo culture, which is essential to perform IVF. On the other hand, effects on the phenotype have been reported consistently in many studies and the finding of trans-generational effects in some studies is an interesting and important datum which needs more investigation. The most significant and clear fact is the LOS which mimics the BWS in humans.

These results obtained in animals rise scientific and bioethical questions of how this evidence should be considered on the increasing use of ART in humans [49]. Studies in animals, when possible and necessary are a bioethical imperative to assess the safety of new treatments and procedures; these are made before introducing the new techniques in medical praxis [216, 217]. However this process, as it is performed today, did not occur with human IVF, due to reasons that are beyond the scope of this review.

Although extrapolation of the results of animal experiments to humans has limitations due to differences among mammalian species, they frequently provide light on what may also happen in humans and are definitively valuable to orient new research.

Ideally, similar experiments as the ones performed in animals, could be done in humans and answer crucial questions. But there are major ethical restrictions to perform such epigenetic studies in human embryos. Most likely, in the near future the analysis of the epigenetic profile of children and adults born by IVF will be possible and this will allow access to information and consequences of this technique that at present are mostly unknown. We agree with several authors in terms that it is mandatory to have better human epidemiological studies and that efforts should be made in order to have a clinical follow up during the whole life of every child born by IVF, because epigenetic alterations in the early stages of development can be expressed later in adult life as different pathologies [218]. This is similar as the discoveries of Barker and others in relation to the effects environmental conditions of intrauterine life in pediatric and adulthood diseases.

Finally, we think that a major concern is how to give parents who seek IVF techniques a clear, objective and prudent information of the evolving knowledge of the possible risks involved in this procedure [219].

References

  1. Barker DJ. Fetal origins of cardiovascular disease. Ann Med. 1999;31(Suppl 1):3–6.

    PubMed  Google Scholar 

  2. Barker DJ, Thornburg KL. The obstetric origins of health for a lifetime. Clin Obstet Gynecol. 2013;56:511–9.

    Article  PubMed  Google Scholar 

  3. Eriksson JG, et al. The developmental origins of chronic rheumatic heart disease. Am J Hum Biol. 2013;25:655–8.

    Article  PubMed  Google Scholar 

  4. Gillman MW. Developmental origins of health and disease. N Engl J Med. 2005;353:1848–50.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  5. Gluckman PD, et al. Towards a new developmental synthesis: adaptive developmental plasticity and human disease. Lancet. 2009;373:1654–7.

    Article  PubMed  Google Scholar 

  6. Fernandez-Gonzalez R, et al. Suboptimal in vitro culture conditions: an epigenetic origin of long-term health effects. Mol Reprod Dev. 2007;74:1149–56.

    Article  CAS  PubMed  Google Scholar 

  7. Gluckman PD, et al. Effect of in utero and early-life conditions on adult health and disease. N Engl J Med. 2008;359:61–73.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  8. Johnson MH. The problematic in vitro embryo in the age of epigenetics. Reprod Biomed Online. 2005;10(Suppl 1):88–96.

    Article  PubMed  Google Scholar 

  9. Gluckman PD, Hanson MA. Living with the past: evolution, development, and patterns of disease. Science. 2004;305:1733–6.

    Article  CAS  PubMed  Google Scholar 

  10. Kwong WY, Osmond C, Fleming TP. Support for Barker hypothesis upheld in rat model of maternal undernutrition during the preimplantation period: application of integrated ‘random effects’ statistical model. Reprod Biomed Online. 2004;8:574–6.

    Article  CAS  PubMed  Google Scholar 

  11. Kwong WY, et al. Maternal undernutrition during the preimplantation period of rat development causes blastocyst abnormalities and programming of postnatal hypertension. Development. 2000;127:4195–202.

    CAS  PubMed  Google Scholar 

  12. Walters E, Edwards R. Further thoughts regarding evidence offered in support of the ‘Barker hypothesis. Reprod Biomed Online. 2009;9:129–31.

    Article  Google Scholar 

  13. Khosla S, et al. Culture of preimplantation mouse embryos affects fetal development and the expression of imprinted genes. Biol Reprod. 2001;64:918–26.

    Article  CAS  PubMed  Google Scholar 

  14. Schultz RM, Williams CJ. The science of ART. Science. 2002;296:2188–90.

    Article  CAS  PubMed  Google Scholar 

  15. Manipalviratn S, DeCherney A, Segars J. Imprinting disorders and assisted reproductive technology. Fertil Steril. 2009;91:305–15.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  16. Calle A, et al. Long-term and transgenerational effects of in vitro culture on mouse embryos. Theriogenology. 2012;77:785–93.

    Article  PubMed  Google Scholar 

  17. Waddington CH. Development as an epigenetic process. In: An introduction to modern genetics. 1939; London: Allen and Unwin.

  18. Wu C, Morris JR. Genes, genetics, and epigenetics: a correspondence. Science. 2001;293:1103–5.

    Article  CAS  Google Scholar 

  19. Bedregal P, et al. Contribution of epigenetics to understand human development. Rev Med Chil. 2010;138:366–72.

    PubMed  Google Scholar 

  20. Holliday R. Epigenetics comes of age in the twentyfirst century. J Genet. 2002;81:1–4.

    Article  CAS  PubMed  Google Scholar 

  21. Hurd PJ. The era of epigenetics. Brief Funct Genomics. 2010;9:425–8.

    Article  PubMed Central  PubMed  Google Scholar 

  22. Osborne-Majnik A, Fu Q, Lane RH. Epigenetic mechanisms in fetal origins of health and disease. Clin Obstet Gynecol. 2013;56:622–32.

    Article  PubMed Central  PubMed  Google Scholar 

  23. Rattner BP. In this issue of epigenetics: special focus on environmental epigenetics. Epigenetics. 2011;6:789–90.

    Article  PubMed  Google Scholar 

  24. Skinner MK. Environmental epigenomics and disease susceptibility. EMBO Rep. 2011;12:620–2.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  25. de Waal E, et al. Gonadotropin stimulation contributes to an increased incidence of epimutations in ICSI-derived mice. Hum Mol Genet. 2012;21:4460–72.

    Article  PubMed Central  PubMed  CAS  Google Scholar 

  26. Faulk C, Dolinoy DC. Timing is everything: the when and how of environmentally induced changes in the epigenome of animals. Epigenetics. 2011;6:791–7.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  27. Mak W, Weaver JR, Bartolomei MS. Is ART changing the epigenetic landscape of imprinting? Anim Reprod. 2010;7:168–76.

    Google Scholar 

  28. Macdonald WA. Epigenetic mechanisms of genomic imprinting: common themes in the regulation of imprinted regions in mammals, plants, and insects. Genet Res Int. 2012;2012:585024.

    PubMed Central  PubMed  Google Scholar 

  29. Inbar-Feigenberg M, et al. Basic concepts of epigenetics. Fertil Steril. 2013;99:607–15.

    Article  CAS  PubMed  Google Scholar 

  30. Kim JK, Samaranayake M, Pradhan S. Epigenetic mechanisms in mammals. Cell Mol Life Sci. 2009;66:596–612.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  31. Sasaki H, et al. Temporal and spatial regulation of H19 imprinting in normal and uniparental mouse embryos. Development. 1995;121:4195–202.

    CAS  PubMed  Google Scholar 

  32. Tremblay KD, et al. A paternal-specific methylation imprint marks the alleles of the mouse H19 gene. Nat Genet. 1995;9:407–13.

    Article  CAS  PubMed  Google Scholar 

  33. Lucas E. Epigenetic effects on the embryo as a result of periconceptional environment and assisted reproduction technology. Reprod Biomed Online. 2013;27:477–85.

    Article  PubMed  Google Scholar 

  34. McEvoy TG, Robinson JJ, Sinclair KD. Developmental consequences of embryo and cell manipulation in mice and farm animals. Reproduction. 2001;122:507–18.

    Article  CAS  PubMed  Google Scholar 

  35. Shi W, Haaf T. Aberrant methylation patterns at the two-cell stage as an indicator of early developmental failure. Mol Reprod Dev. 2002;63:329–34.

    Article  CAS  PubMed  Google Scholar 

  36. Arnaud P. Genomic imprinting in germ cells: imprints are under control. Reproduction. 2010;140:411–23.

    Article  CAS  PubMed  Google Scholar 

  37. Khoueiry R, et al. Dynamic CpG methylation of the KCNQ1OT1 gene during maturation of human oocytes. J Med Genet. 2008;45:583–8.

    Article  CAS  PubMed  Google Scholar 

  38. Laprise SL. Implications of epigenetics and genomic imprinting in assisted reproductive technologies. Mol Reprod Dev. 2009;76:1006–18.

    Article  CAS  PubMed  Google Scholar 

  39. Market-Velker BA, Fernandes AD, Mann MR. Side-by-side comparison of five commercial media systems in a mouse model: suboptimal in vitro culture interferes with imprint maintenance. Biol Reprod. 2010;83:938–50.

    Article  CAS  PubMed  Google Scholar 

  40. Sato A, et al. Aberrant DNA methylation of imprinted loci in superovulated oocytes. Hum Reprod. 2007;22:26–35.

    Article  CAS  PubMed  Google Scholar 

  41. Fauque P, et al. Assisted reproductive technology affects developmental kinetics, H19 imprinting control region methylation and H19 gene expression in individual mouse embryos. BMC Dev Biol. 2007;7:116.

    Article  PubMed Central  PubMed  CAS  Google Scholar 

  42. Fortier AL, et al. Superovulation alters the expression of imprinted genes in the midgestation mouse placenta. Hum Mol Genet. 2008;17:1653–65.

    Article  CAS  PubMed  Google Scholar 

  43. Grazul-Bilska AT, et al. Placental development during early pregnancy in sheep: effects of embryo origin on fetal and placental growth and global methylation. Theriogenology. 2013;79:94–102.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  44. Fossum GT, Davidson A, Paulson RJ. Ovarian hyperstimulation inhibits embryo implantation in the mouse. J In Vitro Fert Embryo Transf. 1989;6:7–10.

    Article  CAS  PubMed  Google Scholar 

  45. Barros CM, et al. Effect of superstimulatory treatments on the expression of genes related to ovulatory capacity, oocyte competence and embryo development in cattle. Reprod Fertil Dev. 2012;25:17–25.

    Article  PubMed  CAS  Google Scholar 

  46. Liang XW, et al. Superovulation induces defective methylation in line-1 retrotransposon elements in blastocyst. Reprod Biol Endocrinol. 2013;11:69.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  47. Bo GA, Mapletoft RJ. Historical perspectives and recent research on superovulation in cattle. Theriogenology. 2014;81:38–48.

    Article  PubMed  Google Scholar 

  48. Borghol N, et al. Epigenetic status of the H19 locus in human oocytes following in vitro maturation. Genomics. 2006;87:417–26.

    Article  CAS  PubMed  Google Scholar 

  49. El Hajj N, Haaf T. Epigenetic disturbances in in vitro cultured gametes and embryos: implications for human assisted reproduction. Fertil Steril. 2013;99:632–41.

    Article  PubMed  CAS  Google Scholar 

  50. Kerjean A, et al. In vitro follicular growth affects oocyte imprinting establishment in mice. Eur J Hum Genet. 2003;11:493–6.

    Article  CAS  PubMed  Google Scholar 

  51. Anckaert E, et al. Effects of low methyl donor levels in culture medium during mouse follicle culture on oocyte imprinting establishment. Biol Reprod. 2010;83:377–86.

    Article  CAS  PubMed  Google Scholar 

  52. Russell DF, et al. The impact of oocyte maturation media on early bovine embryonic development. Mol Reprod Dev. 2006;73:1255–70.

    Article  CAS  PubMed  Google Scholar 

  53. van de Sandt JJ, Schroeder AC, Eppig JJ. Culture media for mouse oocyte maturation affect subsequent embryonic development. Mol Reprod Dev. 1990;25:164–71.

    Article  PubMed  Google Scholar 

  54. Marchal R, et al. DNA methylation in mouse gametogenesis. Cytogenet Genome Res. 2004;105:316–24.

    Article  CAS  PubMed  Google Scholar 

  55. Oakes CC, et al. Developmental acquisition of genome-wide DNA methylation occurs prior to meiosis in male germ cells. Dev Biol. 2007;307:368–79.

    Article  CAS  PubMed  Google Scholar 

  56. Congras A, et al. Sperm DNA methylation analysis in swine reveals conserved and species-specific methylation patterns and highlights an altered methylation at the GNAS locus in infertile boars. Biol Reprod. 2014;91:137.

    Article  PubMed  CAS  Google Scholar 

  57. Marques CJ, et al. Abnormal methylation of imprinted genes in human sperm is associated with oligozoospermia. Mol Hum Reprod. 2008;14:67–74.

    Article  CAS  PubMed  Google Scholar 

  58. Poplinski A, et al. Idiopathic male infertility is strongly associated with aberrant methylation of MEST and IGF2/H19 ICR1. Int J Androl. 2010;33:642–9.

    CAS  PubMed  Google Scholar 

  59. Rousseaux S, et al. Establishment of male-specific epigenetic information. Gene. 2005;345:139–53.

    Article  CAS  PubMed  Google Scholar 

  60. Hasegawa H, et al. A novel culture system for mouse spermatid maturation which produces elongating spermatids capable of inducing calcium oscillation during fertilization and embryonic development. J Assist Reprod Genet. 2010;27:565–70.

    Article  PubMed Central  PubMed  Google Scholar 

  61. Fernandez-Gonzalez R, et al. Long-term effects of mouse intracytoplasmic sperm injection with DNA-fragmented sperm on health and behavior of adult offspring. Biol Reprod. 2008;78:761–72.

    Article  CAS  PubMed  Google Scholar 

  62. Ertzeid G, Storeng R. The impact of ovarian stimulation on implantation and fetal development in mice. Hum Reprod. 2001;16:221–5.

    Article  CAS  PubMed  Google Scholar 

  63. Van der Auwera I, D’Hooghe T. Superovulation of female mice delays embryonic and fetal development. Hum Reprod. 2001;16:1237–43.

    Article  PubMed  Google Scholar 

  64. Piotrowska-Nitsche K, Chan AW. Effect of sperm entry on blastocyst development after in vitro fertilization and intracytoplasmic sperm injection—mouse model. J Assist Reprod Genet. 2013;30:81–9.

    Article  PubMed Central  PubMed  Google Scholar 

  65. Ajduk A, Yamauchi Y, Ward MA. Sperm chromatin remodeling after intracytoplasmic sperm injection differs from that of in vitro fertilization. Biol Reprod. 2006;75:442–51.

    Article  CAS  PubMed  Google Scholar 

  66. Hewitson L, et al. Cellular and molecular events after in vitro fertilization and intracytoplasmic sperm injection. Theriogenology. 2000;53:95–104.

    Article  CAS  PubMed  Google Scholar 

  67. Nygren KG, Andersen AN. Assisted reproductive technology in Europe, 1999. Results generated from European registers by ESHRE. Hum Reprod. 2002;17:3260–74.

    Article  CAS  PubMed  Google Scholar 

  68. Mansour R, et al. International Committee for Monitoring Assisted Reproductive Technologies world report: assisted reproductive technology 2006. Hum Reprod. 2014;29:1536–51.

    Article  PubMed  Google Scholar 

  69. Sullivan EA, et al. International Committee for Monitoring Assisted Reproductive Technologies (ICMART) world report: assisted reproductive technology 2004. Hum Reprod. 2013;28:1375–90.

    Article  CAS  PubMed  Google Scholar 

  70. Bonduelle M, et al. Mental development of 201 ICSI children at 2 years of age. Lancet. 1998;351:1553.

    Article  CAS  PubMed  Google Scholar 

  71. Hansen M, et al. The risk of major birth defects after intracytoplasmic sperm injection and in vitro fertilization. N Engl J Med. 2002;346:725–30.

    Article  PubMed  Google Scholar 

  72. Lie RT, et al. Birth defects in children conceived by ICSI compared with children conceived by other IVF-methods; a meta-analysis. Int J Epidemiol. 2005;34:696–701.

    Article  PubMed  Google Scholar 

  73. Palermo GD, et al. Genetic and epigenetic characteristics of ICSI children. Reprod Biomed Online. 2008;17:820–33.

    Article  PubMed  Google Scholar 

  74. Hewitson L, et al. Fertilization and embryo development to blastocysts after intracytoplasmic sperm injection in the rhesus monkey. Hum Reprod. 1998;13:3449–55.

    Article  CAS  PubMed  Google Scholar 

  75. Dozortsev D, et al. Intracytoplasmic sperm injection in the rat. Zygote. 1998;6:143–7.

    Article  CAS  PubMed  Google Scholar 

  76. Goto K, et al. Fertilization by sperm injection in cattle. Theriogenology. 1990;33:238a.

    Article  Google Scholar 

  77. Kohda T, et al. Intracytoplasmic sperm injection induces transcriptome perturbation without any transgenerational effect. Biochem Biophys Res Commun. 2011;410:282–8.

    Article  CAS  PubMed  Google Scholar 

  78. Giritharan G, et al. Effect of ICSI on gene expression and development of mouse preimplantation embryos. Hum Reprod. 2010;25:3012–24.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  79. Bridges PJ, et al. Methodology matters: IVF versus ICSI and embryonic gene expression. Reprod Biomed Online. 2011;23:234–44.

    Article  PubMed Central  PubMed  Google Scholar 

  80. Garcia-Rosello E, et al. Intracytoplasmic sperm injection in livestock species: an update. Reprod Domest Anim. 2009;44:143–51.

    Article  CAS  PubMed  Google Scholar 

  81. Parmegiani L, Cognigni GE, Filicori M. Sperm selection: effect on sperm DNA quality. Adv Exp Med Biol. 2014;791:151–72.

    Article  PubMed  Google Scholar 

  82. Yu Y, et al. Microinjection manipulation resulted in the increased apoptosis of spermatocytes in testes from intracytoplasmic sperm injection (ICSI) derived mice. PLoS One. 2011;6:e22172.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  83. Hunt PA, et al. Bisphenol a exposure causes meiotic aneuploidy in the female mouse. Curr Biol. 2003;13:546–53.

    Article  CAS  PubMed  Google Scholar 

  84. Berger RG, Foster WG, deCatanzaro D. Bisphenol-A exposure during the period of blastocyst implantation alters uterine morphology and perturbs measures of estrogen and progesterone receptor expression in mice. Reprod Toxicol. 2010;30:393–400.

    Article  CAS  PubMed  Google Scholar 

  85. Berger RG, Hancock T, deCatanzaro D. Influence of oral and subcutaneous bisphenol-A on intrauterine implantation of fertilized ova in inseminated female mice. Reprod Toxicol. 2007;23:138–44.

    Article  CAS  PubMed  Google Scholar 

  86. Berger RG, Shaw J, deCatanzaro D. Impact of acute bisphenol-A exposure upon intrauterine implantation of fertilized ova and urinary levels of progesterone and 17beta-estradiol. Reprod Toxicol. 2008;26:94–9.

    Article  CAS  PubMed  Google Scholar 

  87. Varayoud J, et al. Neonatal exposure to bisphenol A alters rat uterine implantation-associated gene expression and reduces the number of implantation sites. Endocrinology. 2011;152:1101–11.

    Article  CAS  PubMed  Google Scholar 

  88. Lagos-Cabre R, Moreno RD. Contribution of environmental pollutants to male infertily: a working model of germ cell apoptosis induced by plasticizers. Biol Res. 2012;45:5–14.

    Article  PubMed  Google Scholar 

  89. Bittner GD, Yang CZ, Stoner MA. Estrogenic chemicals often leach from BPA-free plastic products that are replacements for BPA-containing polycarbonate products. Environ Health. 2014;13:41.

    Article  PubMed Central  PubMed  CAS  Google Scholar 

  90. Grady R, Sathyanarayana S. An update on phthalates and male reproductive development and function. Curr Urol Rep. 2012;13:307–10.

    Article  PubMed  Google Scholar 

  91. Manikkam M, et al. Plastics derived endocrine disruptors (BPA, DEHP and DBP) induce epigenetic transgenerational inheritance of obesity, reproductive disease and sperm epimutations. PLoS One. 2013;8:e55387.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  92. Mahalingaiah S, et al. Bisphenol A is not detectable in media or selected contact materials used in IVF. Reprod Biomed Online. 2012;25:608–11.

    Article  CAS  PubMed  Google Scholar 

  93. Allen RL, Bondioli KR, Wright RW Jr. The ability of fetal calf serum, new-born calf serum and normal steer serum to promote the in vitro development of bovine morulae. Theriogenology. 1982;18:185–9.

    Article  CAS  PubMed  Google Scholar 

  94. Betterbed B, Wright RW Jr. Development of one-cell ovine embryos in two culture media under two gas atmospheres. Theriogenology. 1985;23:547–53.

    Article  CAS  PubMed  Google Scholar 

  95. Caro CM, Trounson A. The effect of protein on preimplantation mouse embryo development in vitro. J In Vitro Fert Embryo Transf. 1984;1:183–7.

    Article  CAS  PubMed  Google Scholar 

  96. Dukelow WR, et al. Preimplantation development of the primate embryo after in vitro fertilization. J Exp Zool. 1983;228:215–21.

    Article  CAS  PubMed  Google Scholar 

  97. Fukuda A, et al. Influence of water quality on in vitro fertilization and embryo development for the mouse. J In Vitro Fert Embryo Transf. 1987;4:40–5.

    Article  CAS  PubMed  Google Scholar 

  98. Kruger TF, et al. The effect of serum supplementation on the cleavage of human embryos. J In Vitro Fert Embryo Transf. 1987;4:10–2.

    Article  CAS  PubMed  Google Scholar 

  99. Naz RK, et al. Factors influencing murine embryo bioassay: effects of proteins, aging of medium, and surgical glove coatings. Fertil Steril. 1986;46:914–9.

    CAS  PubMed  Google Scholar 

  100. Fleming TP, et al. The embryo and its future. Biol Reprod. 2004;71:1046–54.

    Article  CAS  PubMed  Google Scholar 

  101. Lonergan P, Fair T. The ART of studying early embryo development: progress and challenges in ruminant embryo culture. Theriogenology. 2014;81:49–55.

    Article  PubMed  Google Scholar 

  102. Lonergan P, et al. Effect of culture environment on gene expression and developmental characteristics in IVF-derived embryos. Theriogenology. 2006;65:137–52.

    Article  CAS  PubMed  Google Scholar 

  103. Morgan HD, et al. Epigenetic reprogramming in mammals. Hum Mol Genet. 2005;14 Spec(1):R47–58.

  104. Dinnyes A, et al. Timing of the first cleavage post-insemination affects cryosurvival of in vitro-produced bovine blastocysts. Mol Reprod Dev. 1999;53:318–24.

    Article  CAS  PubMed  Google Scholar 

  105. Lonergan P, et al. Effect of time interval from insemination to first cleavage on the developmental characteristics, sex ratio and pregnancy rate after transfer of bovine embryos. J Reprod Fertil. 1999;117:159–67.

    Article  CAS  PubMed  Google Scholar 

  106. Totey SM, et al. Differential cleavage and developmental rates and their correlation with cell numbers and sex ratios in buffalo embryos generated in vitro. Theriogenology. 1996;45:521–33.

    Article  CAS  PubMed  Google Scholar 

  107. Gutierrez-Adan A, et al. Effect of speed of development on mRNA expression pattern in early bovine embryos cultured in vivo or in vitro. Mol Reprod Dev. 2004;68:441–8.

    Article  CAS  PubMed  Google Scholar 

  108. Lonergan P, et al. Temporal sensitivity of bovine embryos to culture environment after fertilization and the implications for blastocyst quality. Reproduction. 2003;126:337–46.

    Article  CAS  PubMed  Google Scholar 

  109. Rizos D, et al. Consequences of bovine oocyte maturation, fertilization or early embryo development in vitro versus in vivo: implications for blastocyst yield and blastocyst quality. Mol Reprod Dev. 2002;61:234–48.

    Article  CAS  PubMed  Google Scholar 

  110. Fauque P, et al. In vitro fertilization and embryo culture strongly impact the placental transcriptome in the mouse model. PLoS One. 2010;5:e9218.

    Article  PubMed Central  PubMed  CAS  Google Scholar 

  111. Fowden AL, et al. The placenta and intrauterine programming. J Neuroendocrinol. 2008;20:439–50.

    Article  CAS  PubMed  Google Scholar 

  112. Beaujean N. Epigenetics, embryo quality and developmental potential. Reprod Fertil Dev. 2014;27:53–62.

    Article  PubMed  CAS  Google Scholar 

  113. Doherty AS, et al. Differential effects of culture on imprinted H19 expression in the preimplantation mouse embryo. Biol Reprod. 2000;62:1526–35.

    Article  CAS  PubMed  Google Scholar 

  114. Li T, et al. IVF results in de novo DNA methylation and histone methylation at an Igf2-H19 imprinting epigenetic switch. Mol Hum Reprod. 2005;11:631–40.

    Article  CAS  PubMed  Google Scholar 

  115. Mann MR, et al. Selective loss of imprinting in the placenta following preimplantation development in culture. Development. 2004;131:3727–35.

    Article  CAS  PubMed  Google Scholar 

  116. Donjacour A, et al. In vitro fertilization affects growth and glucose metabolism in a sex-specific manner in an outbred mouse model. Biol Reprod. 2014;90:80.

    Article  PubMed Central  PubMed  CAS  Google Scholar 

  117. Sui L, et al. Dynamic proteomic profiles of in vivo- and in vitro-produced mouse postimplantation extraembryonic tissues and placentas. Biol Reprod. 2014;91:155.

    Article  PubMed  CAS  Google Scholar 

  118. Chason RJ, et al. Environmental and epigenetic effects upon preimplantation embryo metabolism and development. Trends Endocrinol Metab. 2011;22:412–20.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  119. Fleming TP, et al. Nutrition of females during the peri-conceptional period and effects on foetal programming and health of offspring. Anim Reprod Sci. 2012;130:193–7.

    Article  CAS  PubMed  Google Scholar 

  120. Chen M, et al. Altered glucose metabolism in mouse and humans conceived by IVF. Diabetes. 2014;63:3189–98.

    Article  CAS  PubMed  Google Scholar 

  121. Crosier AE, et al. Development of skeletal muscle and expression of candidate genes in bovine fetuses from embryos produced in vivo or in vitro. Biol Reprod. 2002;67:401–8.

    Article  CAS  PubMed  Google Scholar 

  122. Behboodi E, et al. Birth of large calves that developed from in vitro-derived bovine embryos. Theriogenology. 1995;44:227–32.

    Article  CAS  PubMed  Google Scholar 

  123. Walker SK, Hartwich KM, Seamark RF. The production of unusually large offspring following embryo manipulation: concepts and challenges. Theriogenology. 1996;45:111–20.

    Article  Google Scholar 

  124. Young LE, Sinclair KD, Wilmut I. Large offspring syndrome in cattle and sheep. Rev Reprod. 1998;3:155–63.

    Article  CAS  PubMed  Google Scholar 

  125. Chen Z, et al. Large offspring syndrome: a bovine model for the human loss-of-imprinting overgrowth syndrome Beckwith–Wiedemann. Epigenetics. 2013;8:591–601.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  126. Farin PW, Piedrahita JA, Farin CE. Errors in development of fetuses and placentas from in vitro-produced bovine embryos. Theriogenology. 2006;65:178–91.

    Article  PubMed  Google Scholar 

  127. Hori N, et al. Aberrant CpG methylation of the imprinting control region KvDMR1 detected in assisted reproductive technology-produced calves and pathogenesis of large offspring syndrome. Anim Reprod Sci. 2010;122:303–12.

    Article  CAS  PubMed  Google Scholar 

  128. Young LE, et al. Epigenetic change in IGF2R is associated with fetal overgrowth after sheep embryo culture. Nat Genet. 2001;27:153–4.

    Article  CAS  PubMed  Google Scholar 

  129. Watkins AJ, et al. Mouse embryo culture induces changes in postnatal phenotype including raised systolic blood pressure. Proc Natl Acad Sci USA. 2007;104:5449–54.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  130. Ecker DJ, et al. Long-term effects of culture of preimplantation mouse embryos on behavior. Proc Natl Acad Sci USA. 2004;101:1595–600.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  131. Strata F, et al. Behavior and brain gene expression changes in mice exposed to preimplantation and prenatal stress. Reprod Sci. 2015;22:23–30.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. Baldoceda-Baldeon LM, et al. Improvement of bovine in vitro embryo production by vitamin K(2) supplementation. Reproduction. 2014;148:489–97.

    Article  PubMed  CAS  Google Scholar 

  133. Cebrian-Serrano A, et al. Effect of the bovine oviductal fluid on in vitro fertilization, development and gene expression of in vitro-produced bovine blastocysts. Reprod Domest Anim. 2013;48:331–8.

    Article  CAS  PubMed  Google Scholar 

  134. Eswari S, Sai Kumar G, Sharma GT. Expression of mRNA encoding leukaemia inhibitory factor (LIF) and its receptor (LIFRbeta) in buffalo preimplantation embryos produced in vitro: markers of successful embryo implantation. Zygote. 2013;21:203–13.

    Article  CAS  PubMed  Google Scholar 

  135. Fukuda A, et al. Effects of the supernatants of mixed lymphocyte cultures and decidual cell line cultures on mouse embryo development in vitro. J In Vitro Fert Embryo Transf. 1989;6:59–64.

    Article  CAS  PubMed  Google Scholar 

  136. Lloyd RE, et al. Effects of oviductal fluid on the development, quality, and gene expression of porcine blastocysts produced in vitro. Reproduction. 2009;137:679–87.

    Article  CAS  PubMed  Google Scholar 

  137. Moore K, Bondioli KR. Glycine and alanine supplementation of culture medium enhances development of in vitro matured and fertilized cattle embryos. Biol Reprod. 1993;48:833–40.

    Article  CAS  PubMed  Google Scholar 

  138. Wang F, et al. Melatonin improves the quality of in vitro produced (IVP) bovine embryos: implications for blastocyst development, cryotolerance, and modifications of relevant gene expression. PLoS One. 2014;9:e93641.

    Article  PubMed Central  PubMed  CAS  Google Scholar 

  139. Bauer BK, et al. Transcriptional profiling by deep sequencing identifies differences in mRNA transcript abundance in in vivo-derived versus in vitro-cultured porcine blastocyst stage embryos. Biol Reprod. 2010;83:791–8.

    Article  CAS  PubMed  Google Scholar 

  140. Boulesteix C, Beaujean N. Fluorescent immunodetection of epigenetic modifications on preimplantation mouse embryos. Methods Mol Biol. 2015;1222:113–26.

    Article  CAS  PubMed  Google Scholar 

  141. Prather RS, et al. Genomic profiling to improve embryogenesis in the pig. Anim Reprod Sci. 2014;149:39–45.

    Article  CAS  PubMed  Google Scholar 

  142. Burgess DJ. Stem cells: epigenome reprogramming—of mice and men. Nat Rev Genet. 2014;15:571.

    Article  PubMed  CAS  Google Scholar 

  143. Chavez SL, et al. Comparison of epigenetic mediator expression and function in mouse and human embryonic blastomeres. Hum Mol Genet. 2014;23:4970–84.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  144. Ar A, Mover H. Oxygen tensions in developing embryos: system inefficiency or system requirement? Isr. J. Zool. 1994;40:306–26.

    Google Scholar 

  145. Fischer B, Bavister BD. Oxygen tension in the oviduct and uterus of rhesus monkeys, hamsters and rabbits. J Reprod Fertil. 1993;99:673–9.

    Article  CAS  PubMed  Google Scholar 

  146. Ross RN, Graves CN. O2 levels in the female rabbit reproductive tract. J Anim Sci. 1979;39:94–9.

    Google Scholar 

  147. Mehta RH. Growth of human preimplantation embryos in vitro. Reprod Biomed Online. 2001;2:113–9.

    Article  CAS  PubMed  Google Scholar 

  148. Pool TB. Recent advances in the production of viable human embryos in vitro. Reprod Biomed Online. 2002;4:294–302.

    Article  PubMed  Google Scholar 

  149. Karagenc L, et al. Impact of oxygen concentration on embryonic development of mouse zygotes. Reprod Biomed Online. 2004;9:409–17.

    Article  CAS  PubMed  Google Scholar 

  150. Kwon HC, et al. Effects of low oxygen condition on the generation of reactive oxygen species and the development in mouse embryos cultured in vitro. J Obstet Gynaecol Res. 1999;25:359–66.

    Article  CAS  PubMed  Google Scholar 

  151. Orsi NM, Leese HJ. Protection against reactive oxygen species during mouse preimplantation embryo development: role of EDTA, oxygen tension, catalase, superoxide dismutase and pyruvate. Mol Reprod Dev. 2001;59:44–53.

    Article  CAS  PubMed  Google Scholar 

  152. Umaoka Y, et al. Effects of oxygen toxicity on early development of mouse embryos. Mol Reprod Dev. 1992;31:28–33.

    Article  CAS  PubMed  Google Scholar 

  153. Adam AA, et al. Effects of oxygen tension in the gas atmosphere during in vitro maturation, in vitro fertilization and in vitro culture on the efficiency of in vitro production of mouse embryos. Jpn J Vet Res. 2004;52:77–84.

    PubMed  Google Scholar 

  154. Bing YZ, et al. Effects of thioredoxin on the preimplantation development of bovine embryos. Theriogenology. 2003;59:863–73.

    Article  CAS  PubMed  Google Scholar 

  155. Kitagawa Y, et al. Effects of oxygen concentration and antioxidants on the in vitro developmental ability, production of reactive oxygen species (ROS), and DNA fragmentation in porcine embryos. Theriogenology. 2004;62:1186–97.

    Article  CAS  PubMed  Google Scholar 

  156. Wale PL, Gardner DK. Oxygen regulates amino acid turnover and carbohydrate uptake during the preimplantation period of mouse embryo development. Biol Reprod. 2012;87(24):1–8.

    Google Scholar 

  157. Arias ME, Sanchez R, Felmer R. Evaluation of different culture systems with low oxygen tension on the development, quality and oxidative stress-related genes of bovine embryos produced in vitro. Zygote. 2012;20:209–17.

    Article  CAS  PubMed  Google Scholar 

  158. Mantikou E, et al. Low oxygen concentrations for embryo culture in assisted reproductive technologies. Hum Reprod Update. 2013;19:209.

    Article  CAS  PubMed  Google Scholar 

  159. Wale PL, Gardner DK. Time-lapse analysis of mouse embryo development in oxygen gradients. Reprod Biomed Online. 2010;21:402–10.

    Article  CAS  PubMed  Google Scholar 

  160. Yuan YQ, et al. Influence of oxygen tension on apoptosis and hatching in bovine embryos cultured in vitro. Theriogenology. 2003;59:1585–96.

    Article  CAS  PubMed  Google Scholar 

  161. Liu L, Keefe DL. Cytoplasm mediates both development and oxidation-induced apoptotic cell death in mouse zygotes. Biol Reprod. 2000;62:1828–34.

    Article  CAS  PubMed  Google Scholar 

  162. Elamaran G, et al. Oxygen concentration and cysteamine supplementation during in vitro production of buffalo (Bubalus bubalis) embryos affect mRNA expression of BCL-2, BCL-XL, MCL-1, BAX and BID. Reprod Domest Anim. 2012;47:1027–36.

    Article  CAS  PubMed  Google Scholar 

  163. Kawamura Y, et al. Sirt3 protects in vitro-fertilized mouse preimplantation embryos against oxidative stress-induced p53-mediated developmental arrest. J Clin Invest. 2010;120:2817–28.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  164. Rocha-Frigoni NA, et al. Effects of gaseous atmosphere and antioxidants on the development and cryotolerance of bovine embryos at different periods of in vitro culture. Zygote. 2015;23:159–68.

    Article  CAS  PubMed  Google Scholar 

  165. Yoon SB, et al. Developmental competence of bovine early embryos depends on the coupled response between oxidative and endoplasmic reticulum stress. Biol Reprod. 2014;90:104.

    Article  PubMed  CAS  Google Scholar 

  166. Catt JW, Henman M. Toxic effects of oxygen on human embryo development. Hum Reprod. 2000;15(Suppl 2):199–206.

    Article  PubMed  Google Scholar 

  167. Kovacic B, Vlaisavljevic V. Influence of atmospheric versus reduced oxygen concentration on development of human blastocysts in vitro: a prospective study on sibling oocytes. Reprod Biomed Online. 2008;17:229–36.

    Article  CAS  PubMed  Google Scholar 

  168. Bontekoe S, et al. Low oxygen concentrations for embryo culture in assisted reproductive technologies. Cochrane Database Syst Rev. 2012;7:CD008950.

  169. Harper JC, Sengupta SB. Preimplantation genetic diagnosis: state of the art 2011. Hum Genet. 2012;131:175–86.

    Article  PubMed  Google Scholar 

  170. Polisseni J, et al. Post-biopsy bovine embryo viability and whole genome amplification in preimplantation genetic diagnosis. Fertil Steril. 2010;93:783–8.

    Article  CAS  PubMed  Google Scholar 

  171. Park JH, et al. Rapid sexing of preimplantation bovine embryo using consecutive and multiplex polymerase chain reaction (PCR) with biopsied single blastomere. Theriogenology. 2001;55:1843–53.

    Article  CAS  PubMed  Google Scholar 

  172. Cenariu M, et al. Evaluation of bovine embryo biopsy techniques according to their ability to preserve embryo viability. J Biomed Biotechnol. 2012;2012:541384.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  173. Takeuchi K, et al. Preclinical models for human pre-embryo biopsy and genetic diagnosis. I. Efficiency and normalcy of mouse pre-embryo development after different biopsy techniques. Fertil Steril. 1992;57:425–30.

    CAS  PubMed  Google Scholar 

  174. Duncan FE, et al. The effect of blastomere biopsy on preimplantation mouse embryo development and global gene expression. Fertil Steril. 2009;91:1462–5.

    Article  PubMed Central  PubMed  Google Scholar 

  175. Sugawara A, et al. Blastomere removal from cleavage-stage mouse embryos alters steroid metabolism during pregnancy. Biol Reprod. 2012;87(4):1–9.

    Google Scholar 

  176. Ugajin T, et al. Aberrant behavior of mouse embryo development after blastomere biopsy as observed through time-lapse cinematography. Fertil Steril. 2010;93:2723–8.

    Article  PubMed  Google Scholar 

  177. Kirkegaard K, Hindkjaer JJ, Ingerslev HJ. Human embryonic development after blastomere removal: a time-lapse analysis. Hum Reprod. 2012;27:97–105.

    Article  PubMed  Google Scholar 

  178. Terada Y, et al. Different embryonic development after blastomere biopsy for preimplantation genetic diagnosis, observed by time-lapse imaging. Fertil Steril. 2009;92:1470–1.

    Article  PubMed  Google Scholar 

  179. Sampino S, et al. Effects of blastomere biopsy on post-natal growth and behavior in mice. Hum Reprod. 2014;29:1875–83.

    Article  CAS  PubMed  Google Scholar 

  180. Wu Y, et al. Blastomere biopsy influences epigenetic reprogramming during early embryo development, which impacts neural development and function in resulting mice. Cell Mol Life Sci. 2014;71:1761–74.

    Article  CAS  PubMed  Google Scholar 

  181. Zhao HC, et al. Aberrant epigenetic modification in murine brain tissues of offspring from preimplantation genetic diagnosis blastomere biopsies. Biol Reprod. 2013;89:117.

    Article  PubMed  CAS  Google Scholar 

  182. Rivera RM, et al. Manipulations of mouse embryos prior to implantation result in aberrant expression of imprinted genes on day 9.5 of development. Hum Mol Genet. 2008;17:1–14.

    Article  CAS  PubMed  Google Scholar 

  183. Fazeli A. Maternal communication with gametes and embryos. Theriogenology. 2008;70:1182–7.

    Article  CAS  PubMed  Google Scholar 

  184. Alminana C, Fazeli A. Exploring the application of high-throughput genomics technologies in the field of maternal-embryo communication. Theriogenology. 2012;77:717–37.

    Article  CAS  PubMed  Google Scholar 

  185. Ulbrich SE, et al. In vitro systems for intercepting early embryo-maternal cross-talk in the bovine oviduct. Theriogenology. 2010;73:802–16.

    Article  CAS  PubMed  Google Scholar 

  186. Wolf E, et al. Embryo-maternal communication in bovine—strategies for deciphering a complex cross-talk. Reprod Domest Anim. 2003;38:276–89.

    Article  CAS  PubMed  Google Scholar 

  187. Robertson SA, et al. Peri-conceptual cytokines—setting the trajectory for embryo implantation, pregnancy and beyond. Am J Reprod Immunol. 2011;66(Suppl 1):2–10.

    Article  PubMed  Google Scholar 

  188. Rottmayer R, et al. A bovine oviduct epithelial cell suspension culture system suitable for studying embryo-maternal interactions: morphological and functional characterization. Reproduction. 2006;132:637–48.

    Article  CAS  PubMed  Google Scholar 

  189. Bauersachs S, Wolf E. Immune aspects of embryo-maternal cross-talk in the bovine uterus. J Reprod Immunol. 2013;97:20–6.

    Article  CAS  PubMed  Google Scholar 

  190. Walker CG, et al. Modulation of the maternal immune system by the pre-implantation embryo. BMC Genomics. 2010;11:474.

    Article  PubMed Central  PubMed  CAS  Google Scholar 

  191. O’Neill C, Li Y, Jin XL. Survival signaling in the preimplantation embryo. Theriogenology. 2012;77:773–84.

    Article  PubMed  CAS  Google Scholar 

  192. Ostrup E, Hyttel P, Ostrup O. Embryo-maternal communication: signalling before and during placentation in cattle and pig. Reprod Fertil Dev. 2011;23:964–75.

    Article  PubMed  Google Scholar 

  193. Thatcher WW, et al. Uterine-conceptus interactions and reproductive failure in cattle. Theriogenology. 2001;56:1435–50.

    Article  CAS  PubMed  Google Scholar 

  194. Enright BP, et al. Culture of in vitro produced bovine zygotes in vitro vs in vivo: implications for early embryo development and quality. Theriogenology. 2000;54:659–73.

    Article  CAS  PubMed  Google Scholar 

  195. Lonergan P, Forde N. Maternal-embryo interaction leading up to the initiation of implantation of pregnancy in cattle. Animal. 2014;8(Suppl 1):64–9.

    Article  CAS  PubMed  Google Scholar 

  196. Maillo V, et al. Oviduct-embryo interactions in cattle: two-way traffic or a one-way street? Biol Reprod. 2015;92:144.

    Article  PubMed  CAS  Google Scholar 

  197. Rizos D, et al. Culture of bovine embryos in intermediate host oviducts with emphasis on the isolated mouse oviduct. Theriogenology. 2010;73:777–85.

    Article  CAS  PubMed  Google Scholar 

  198. Dorniak P, et al. Endometrial HSD11B1 and cortisol regeneration in the ovine uterus: effects of pregnancy, interferon tau, and prostaglandins. Biol Reprod. 2012;86:124.

    Article  PubMed  CAS  Google Scholar 

  199. Spencer TE, et al. Conceptus-derived prostaglandins regulate gene expression in the endometrium prior to pregnancy recognition in ruminants. Reproduction. 2013;146:377–87.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  200. Schmaltz-Panneau B, et al. Early bovine embryos regulate oviduct epithelial cell gene expression during in vitro co-culture. Anim Reprod Sci. 2014;149:103–16.

    Article  CAS  PubMed  Google Scholar 

  201. Besenfelder U, Havlicek V, Brem G. Role of the oviduct in early embryo development. Reprod Domest Anim. 2012;47(Suppl 4):156–63.

    Article  PubMed  Google Scholar 

  202. Block J, et al. Use of insulin-like growth factor-I during embryo culture and treatment of recipients with gonadotropin-releasing hormone to increase pregnancy rates following the transfer of in vitro-produced embryos to heat-stressed, lactating cows. J Anim Sci. 2003;81:1590–602.

    CAS  PubMed  Google Scholar 

  203. Block J, et al. Improving post-transfer survival of bovine embryos produced in vitro: actions of insulin-like growth factor-1, colony stimulating factor-2 and hyaluronan. Theriogenology. 2011;76:1602–9.

    Article  CAS  PubMed  Google Scholar 

  204. Sirisathien S, Hernandez-Fonseca HJ, Brackett BG. Influences of epidermal growth factor and insulin-like growth factor-I on bovine blastocyst development in vitro. Anim Reprod Sci. 2003;77:21–32.

    Article  CAS  PubMed  Google Scholar 

  205. Li G, et al. Genes of the transforming growth factor-beta signalling pathway are associated with pre-implantation embryonic development in cattle. J Dairy Res. 2012;79:310–7.

    Article  CAS  PubMed  Google Scholar 

  206. Mahsoudi B, Li A, O’Neill C. Assessment of the long-term and transgenerational consequences of perturbing preimplantation embryo development in mice. Biol Reprod. 2007;77:889–96.

    Article  CAS  PubMed  Google Scholar 

  207. Lavara R, et al. Long-term and transgenerational effects of cryopreservation on rabbit embryos. Theriogenology. 2014;81:988–92.

    Article  CAS  PubMed  Google Scholar 

  208. Calle A, et al. Male mice produced by in vitro culture have reduced fertility and transmit organomegaly and glucose intolerance to their male offspring. Biol Reprod. 2012;87:34.

    Article  PubMed  CAS  Google Scholar 

  209. Fernandez-Gonzalez R, et al. Histone modifications at the blastocyst Axin1(Fu) locus mark the heritability of in vitro culture-induced epigenetic alterations in mice. Biol Reprod. 2010;83:720–7.

    Article  CAS  PubMed  Google Scholar 

  210. Delbes G, Hales BF, Robaire B. Toxicants and human sperm chromatin integrity. Mol Hum Reprod. 2010;16:14–22.

    Article  CAS  PubMed  Google Scholar 

  211. Szyf M. Nongenetic inheritance and transgenerational epigenetics. Trends Mol Med. 2015;21:134–44.

    Article  PubMed  Google Scholar 

  212. Gaydos LJ, Wang W, Strome S. Gene repression. H3K27me and PRC2 transmit a memory of repression across generations and during development. Science. 2014;345:1515–8.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  213. Rassoulzadegan M, et al. RNA-mediated non-mendelian inheritance of an epigenetic change in the mouse. Nature. 2006;441:469–74.

    Article  CAS  PubMed  Google Scholar 

  214. Rowson LE. The second Hammond Memorial Lecture. The role of reproductive research in animal production. J Reprod Fertil. 1971;26:113–26.

    Article  CAS  PubMed  Google Scholar 

  215. Hackam DG, Redelmeier DA. Translation of research evidence from animals to humans. JAMA. 2006;296:1731–2.

    CAS  PubMed  Google Scholar 

  216. Hyun I, et al. New ISSCR guidelines underscore major principles for responsible translational stem cell research. Cell Stem Cell. 2008;3:607–9.

    Article  CAS  PubMed  Google Scholar 

  217. Santos MJ, Ventura-Junca P. Bioethical aspects of basic research and medical applications of human stem cells. Biol Res. 2012;45:317–26.

    Article  PubMed  Google Scholar 

  218. Niemitz EL, Feinberg AP. Epigenetics and assisted reproductive technology: a call for investigation. Am J Hum Genet. 2004;74:599–609.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  219. Hansen M, et al. Assisted reproductive technology and birth defects: a systematic review and meta-analysis. Hum Reprod Update. 2013;19:330–53.

    Article  PubMed  Google Scholar 

Download references

Authors’ contributions

PVJ conceived and design the study. Coordination of the group. Writing the first draft. II, AR participated in the search and resume of articles and evaluation of its quality and helped to draft the manuscript. JG participated in the search and resume of articles and evaluation of its quality. RM, MJS participated in the design of the study, in evaluating the molecular aspects involve in the revision and in reviewing the draft. All authors participated in the bioethical discussion. All authors read and approved the final manuscript.

Acknowledgements

This work was partly financed by Grants VRI-PUC 1573/DPCC 2012; 4467/DPCC 2015; and 1150352 (to RDM).

Competing interests

The authors declare that they have no competing interests.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Patricio Ventura-Juncá.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ventura-Juncá, P., Irarrázaval, I., Rolle, A.J. et al. In vitro fertilization (IVF) in mammals: epigenetic and developmental alterations. Scientific and bioethical implications for IVF in humans. Biol Res 48, 68 (2015). https://doi.org/10.1186/s40659-015-0059-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40659-015-0059-y

Keywords