Next Article in Journal
Nerve Structure-Function: Unusual Structural Details and Unmasking of Sulfhydryl Groups by Electrical Stimulation or Asphyxia in Axon Membranes and Gap Junctions
Previous Article in Journal
Metabolomic Analysis of Pediatric Patients with Idiosyncratic Drug-Induced Liver Injury According to the Updated RUCAM
Previous Article in Special Issue
The Next Frontier in ART: Harnessing the Uterine Immune Profile for Improved Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Long-Term Effects of ART on the Health of the Offspring

by
Hamid Ahmadi
1,
Leili Aghebati-Maleki
2,3,
Shima Rashidiani
4,
Timea Csabai
1,5,6,7,
Obodo Basil Nnaemeka
8 and
Julia Szekeres-Bartho
1,5,6,7,9,*
1
Department of Medical Biology and Central Electron Microscope Laboratory, Medical School, Pécs University, 7624 Pécs, Hungary
2
Department of Immunology, School of Medicine, Tabriz University of Medical Sciences, Tabriz 5165665931, Iran
3
Immunology Research Center, Tabriz University of Medical Sciences, Tabriz 5165665931, Iran
4
Department of Medical Biochemistry, Medical School, Pécs University, 7624 Pécs, Hungary
5
János Szentágothai Research Centre, Pécs University, 7624 Pécs, Hungary
6
Endocrine Studies, Centre of Excellence, Pécs University, 7624 Pécs, Hungary
7
National Laboratory of Human Reproduction, 7624 Pécs, Hungary
8
Department of Laboratory Diagnostics, Faculty of Health Sciences, Pécs University, 7621 Pécs, Hungary
9
MTA—PTE Human Reproduction Research Group, 7624 Pecs, Hungary
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(17), 13564; https://doi.org/10.3390/ijms241713564
Submission received: 17 August 2023 / Revised: 29 August 2023 / Accepted: 29 August 2023 / Published: 1 September 2023

Abstract

:
Assisted reproductive technologies (ART) significantly increase the chance of successful pregnancy and live birth in infertile couples. The different procedures for ART, including in vitro fertilization (IVF), intracytoplasmic sperm injection (ICSI), intrauterine insemination (IUI), and gamete intrafallopian tube transfer (GIFT), are widely used to overcome infertility-related problems. In spite of its inarguable usefulness, concerns about the health consequences of ART-conceived babies have been raised. There are reports about the association of ART with birth defects and health complications, e.g., malignancies, high blood pressure, generalized vascular functional disorders, asthma and metabolic disorders in later life. It has been suggested that hormonal treatment of the mother, and the artificial environment during the manipulation of gametes and embryos may cause genomic and epigenetic alterations and subsequent complications in the health status of ART-conceived babies. In the current study, we aimed to review the possible long-term consequences of different ART procedures on the subsequent health status of ART-conceived offspring, considering the confounding factors that might account for/contribute to the long-term consequences.

1. Introduction

Forty years after the birth of the first in vitro fertilization (IVF)- conceived baby and more than 27 years after the first intracytoplasmic sperm injection (ICSI), these techniques are widely used to overcome infertility-related problems [1]. It has been estimated that around 200,000 babies are annually born using assisted reproductive technologies (ART) worldwide, accounting for 5% of all conceived babies in developed countries [2]. In parallel with higher demands for ART, concerns about the health status of ART-conceived babies have also increased [3].
The concluding results from a careful analysis of epidemiological data confirmed that ART is associated with a slight increase in birth defects, with an anomaly rate of 3–4% at birth versus 2–3% at natural conception [4]. Although some studies found no association between ART and the health of the offspring, others suggested that ART-conceived babies exhibited a higher prevalence rate of cancer [5,6], high blood pressure, generalized vascular functional disorders, asthma, as well as metabolic and epigenetic disorders [7,8,9].
According to the theory of the developmental origin of health and disease, a suboptimal early-life environment increases the risk for non-communicable diseases throughout life. Pre-pregnancy, pre-implantation, and intrauterine environments might affect gene expression, adult phenotype, and susceptibility to disease in the offspring’s later life [10]. During ART processes, altered environmental conditions, including hormone administration, gametes, and embryo manipulation, and interfering in the natural selection of gametes in the preimplantation period lead to genomic and epigenetic alterations, which might result in enhanced incidence risk of malformations and disorders throughout the later life of ART-conceived babies [11]. Preimplantation embryos are highly sensitive to environmental factors. Some of the potential stress factors result in impaired embryo development and implantation failure, while others might have long-term effects.
However, the population that requires ART to conceive is different from fertile patients in many respects. Furthermore, infertility in itself might affect the offspring; the mothers are considerably older at the time of conception than the fertile population. Finally, yet importantly, many of the ART-conceived mothers had multiple pregnancies. Therefore, the available evidence must be handled carefully before reaching a conclusion [12]. The current study is aimed to review the data concerning the association between infertility treatment and offspring’s health status in later life.

2. Assisted Reproductive Technologies (ART)

The infertility treatment protocol should particularly target the diagnosed cause of infertility, applying proper methods that guarantee the highest chance for successful pregnancy and delivery of viable offspring. Infertility treatments include three main therapeutic strategies: pharmacologic therapy, surgical therapy, and ART. Recent decades have witnessed great development in ART, resulting in the treatment of previously untreatable and challenging cases [8]. Since the birth of Louis Brown in 1978, over 8 million babies have been born using different ART procedures [13].
Over the past four decades, several important new techniques have been introduced. Adaptation of ultrasonography to evaluate ovarian follicles, the introduction of controlled ovarian hyperstimulation with aromatase inhibitors, clomiphene citrate and menopausal gonadotrophins, administration of gonadotropin-releasing hormone (GnRH) agonist and antagonist in stimulation protocols, performing oocyte retrieval under ultrasonographic guidance, embryo freezing, and ultra-rapid embryo freezing, have resulted in increased efficiency of ART [14,15,16,17].
The first successful pregnancy and live birth after microsurgical epididymal sperm aspiration (MESA) was reported in 1988, followed by live birth after intracytoplasmic sperm injection (ICSI) in 1992. These procedures are highly effective in overcoming male factor infertility and are considered significant progress in the ART world [18,19].
IVF, ICSI, and, in the past, gamete intrafallopian tube transfer (GIFT) and zygote intrafallopian tube transfer (ZIFT) are/were the procedures to treat infertility in humans. In IVF, fertilization occurs as a result of adding a defined amount of sperm to the extracted oocyte in a culture medium, followed by culture and transfer of the embryo into the uterus, whereas in ICSI a single spermatozoon is injected into the cytoplasm of a mature oocyte [18]. ICSI is associated with high fertilization and pregnancy rates regardless of sperm concentration, morphology, and motility in male-related infertility; on the other hand, the competition among the spermatozoa is eliminated. GIFT could be used as an alternative to IVF in women with at least one patent tube. Intrauterine insemination (IUI) is often the first therapeutic intervention that is offered to treat subfertility. In this procedure, the sperm are brought closer to the oocyte for fertilization at the appropriate time. Economic evidence favors an IUI trial of three to four cycles prior to IVF in many cases and provides for in vivo embryo development [18,19].
In vitro maturation (IVM) of immature oocytes was another important step in ART development [20]. In IVM, immature follicles are minimally exposed to hormonal stimulation, therefore this is an alternative approach for IVF in women with polycystic ovary syndrome (PCOS), who are at risk of ovarian hyperstimulation syndrome.
Recombinant gonadotrophins have been used in hyperstimulation protocols since 1992 [21]. Preimplantation genetic testing (PGT) was introduced in 1999 and the first live birth after blastocyst biopsy was reported three years later [22,23]. Figure 1 shows the different ART procedures.

3. Preimplantation Stress and Development

David Barker was the first who clearly articulated the developmental origin of health and disease theory [24]. During development, there is a specific period when the organism is plastic and sensitive to its surrounding environment. While plasticity provides adaptation to the environment and functions as a survival mechanism for the developing embryo, environmental stimuli during early development generate permanent effects that persist throughout the lifespan [24,25] and predispose the resultant offspring to diseases in adulthood [26,27].
During the fertilization and preimplantation period, organisms are highly vulnerable to environmental conditions [28]. Preimplantation development is characterized by highly coordinated physiological and epigenetic alterations as the zygote develops into a blastocyst. The diverse energy requirement for this transitional phase are supplied by the spatio-temporal availability of nutrients, oxygen, and growth factors as the developing embryo travels from the oviduct to the uterus.
It is accepted that the intrauterine environment affects gene expression, adult phenotypes, and disease susceptibility. Animal experiments confirmed the effect of varying pre-implantation and prenatal conditions on the offspring phenotype. Feeding rats with a low-protein diet for 4 days prior to implantation resulted in lower blastocyst cell numbers lower perinatal and postnatal growth rates, and increased incidence of cardiovascular disease, hypertension, and mild hyperglycemia in adult offspring [26,29]. Maternal malnutrition has similar effects on the offspring in humans [25] In another study, the low-protein diet of the mother caused adult hypertension, as well as hyperactive behavior and altered postnatal growth in mice [27]. Adamiak et al. showed a correlation between undernutrition during the prenatal period and increased vasoconstriction, as well as impaired glucose tolerance in offspring in cattle [30].
Preimplantation development encompasses the fertilization of the oocyte by the sperm through the implantation of the hatched blastocyst into the uterus lining. Following the fertilization of the oocyte, the single-cell zygote is formed, which undergoes several rounds of cleavage without increasing the whole embryo volume [31]. As the embryo reaches the eight-cell stage, polarization allows compaction, which provides a foundation for establishing distinct cell lineages: the trophectoderm (TE) and inner cell mass (ICM). These cell lineages evolve over subsequent asymmetric cleavage divisions, continuing through the formation of the blastocoel cavity. Cavitation results in the full expansion of the blastocyst, which hatches from its surrounding zona pellucid layer before implanting into the endometrium [32].
These stages are characterized by different nutritional requirements. During the progression from the zygote to the two-cell stage, there is an absolute requirement for pyruvate, while further developmental stages are highly supported by lactate [33].
Glucose is not required until the late four-cell/early eight-cell stage but becomes the main carbohydrate metabolized during embryo compaction [34]. The transition in metabolic and energy substrate requirements is affected by the availability of amino acids, carbohydrates, and extracellular pH, which influences oxidative phosphorylation, glycolytic activity, and membrane transport, with potential effects on proliferation and growth [35,36]. The development of the preimplantation embryo occurs over a dynamic spectrum of conditions, indicating strong metabolic plasticity and the capacity of the embryo to compensate for metabolic fluctuations at this time [37,38].
During mammalian preimplantation development, coordinated reprogramming of the genome occurs, which maintains after birth [39]. Since preimplantation development is characterized by epigenetic rearrangement, the embryos may be particularly vulnerable to insults. Environmental changes during the preimplantation period could cause errors or epimutations or alter the programming of the cell states and response pathways [40].
Evidence from animal studies suggest that stress limited to this period can exert both short and long-term effects [27,30]. In addition to metabolic conditions, toxicologic agents can affect early pregnancy [41]. Developmental exposure to diethylstilbestrol or bisphenol A activates estrogen-responsive genes, affects proliferation, and increases susceptibility to cancer in adulthood [41]. Higher doses of bisphenol A disrupted embryo growth and development in mice [42]. Smoking before pregnancy impairs embryonic development. Tobacco exposure is associated with decreased ovulation rate, altered tubal function, and perturbated embryo growth in female mice [43]. In mice, tobacco exposure reduced both sperm count and preimplantation embryo development [44].
During the laboratory manipulation of ART, the gametes and embryos are exposed to different kinds of stress, e.g., in vitro culture of the embryos, light, etc. During natural conception, mammalian gametes and embryos are not exposed to direct light and therefore they may not have developed the protective mechanisms [45]. The laboratory manipulation of light-related damage to oocytes and embryos can be seen at various levels, including mitochondrial degradation [46], DNA fragmentation [47] and ROS production in the cytoplasm [48]. Bognar et al. showed that exposure to white light impairs the implantation potential of in vitro cultured mouse embryos. The harmful effects, which were related to the wavelength rather than to the brightness of the light, could be partly corrected by using a red filter [49]. Based on these results, in a human IVF facility, red filters were applied on laboratory lamps and UV or infrared filters in microscopes in order to eliminate white light exposure of the cells throughout all work stages. These precautions resulted in significantly higher fertilization, blastocyst development, and better clinical pregnancy rates in light-protected than in non-protected ICSI cycles [50].
Different types of preimplantation stresses and their effects on embryo development and long-term effects are summarized in Table 1.
Animal models are essential to evaluate the effect of preimplantation stress on subsequent development and health status and to understand the mechanisms underlying the embryo stress response and potential outcomes later in life [40]. Since animal studies remove infertility as a potentially confounding factor, they allow the assessment of a particular controlled stress and extend investigations with available molecular tools. Such studies may provide useful information that could help to authenticate human ART findings [69,70]. However, animal studies are not completely immune to errors and bias [71].
It is documented that in vitro culture postpones embryo development by 18 to 24 h [72]. Variables such as the composition of culture media, fertilization procedures, oxygen tension, temperature fluctuations, and culture dish stiffness can impact the ratio of the embryo developing to blastocyst, TE and ICM cell number and lineage ratio, embryo morphology and gene expression [11,56]. Schwarzer and colleagues investigated the effect of 13 different culture protocols on mouse embryo development and found that different culture conditions exerted distinct effects on developmental competence, lineage composition, and global gene expression [73]. Each of these differences could significantly impact offspring growth and health in later life. IVF and ICSI as well as in vitro culture reduce differential gene expression between inner cell mass and trophectoderm, downregulate the expression of genes involved in transportation of micronutrients and placentation as well as altered expression of various imprinted genes [52,54,56]. This is particularly relevant, since imprinted genes take part in regulating placentation, placenta efficiency, and nutrient transport capacity [74].
Fetuses from IVF have smaller sizes throughout the first half of gestation, but exhibit rapid catch-up growth to controls during the second half of pregnancy, probably due to enlarged placentae. At birth, there is no statistically significant difference in weight between IVF and control animals, suggesting that birth weight cannot be a strong indicator of fetal growth and nutrition [75].
Ecker and colleagues were the first who reported that stress during the preimplantation period can produce clear definable postnatal phenotypes. They demonstrated that culturing embryos from the two-cell to blastocyst stage results in adult male offspring with decreased anxiety and impaired spatial memory. The results also demonstrated that the culture or embryo transfer has no significant effects on embryo development to term [28]. Another study by Fernandez-Gonzalez et al. observed similar behavioral changes in mice transferred at the two-cell stage following ICSI [76].
Increased body weight after preimplantation culture, ICSI, or IVF has been described by several research groups [12,76,77]. IVF and ICSI-conceived mice exhibit signs of glucose intolerance and insulin resistance [77]. It has been indicated that in vitro stress affects organ size and integrity: in a study by Gonzalez et al., it was demonstrated that mice embryo culture is associated with kidney inflammation, pneumonia, and testicular atrophy, as well as possible liver steatosis, reflecting disrupted glycogen metabolism and subsequent lipid mobilization as a compensatory energy source [12]. It has also been suggested that the stress during embryo culture results in altered expression of several metabolic and cardiovascular genes, as well as increased systolic blood pressure in adult mice [78].

4. ART and Imprinting Disorders (IDs)

Epigenetics is defined as the study of inheritable alterations in gene function that do not entail a change in DNA sequence. Epigenetic modifications include adjustments to DNA bases, histone modification, post-translational regulations, and genomic imprinting [79]. It has been confirmed by human and animal studies that reprogramming of specific genes is important for proper embryo development [80,81].
Several studies suggest an association of ART with imprinting disorders. Rare genetic diseases, e.g., Beckwith–Weidemann syndrome, Angelman syndrome, Silver–Russell syndrome, and Prader–Willi syndrome seem to be more frequent among ART-conceived babies [82,83,84,85,86], but when comparing 75,000 ART conceived children to 2,775,239 naturally conceived children, only the incidence of Beckwith–Wiedemann syndrome was higher in the former group [81].
Though it is plausible that ART possibly interferes with the establishment and/or maintenance of imprints, the data are controversial and are not clear enough to allow conclusions. The patients’ data are incomplete and lack molecular characterizations of the diagnosis; furthermore, variables such as maternal age, infertility, underlying causes, and specific ART methods are often not included in the analyses [87].
ART procedures occur during a critically coordinated period of epigenetic reprogramming, the preimplantation period, which is highly vulnerable to epigenetic aberrations [88].
Because of the heterogeneous nature of ART procedures, differences between imprinted regions, and disparity in tissue samples and detection techniques, our understanding of the effects of ART on imprinting and DNA methylation is limited [89]. ART involves several processes, such as hormonal ovarian hyperstimulation, in vitro oocyte maturation, male gamete processing, methods of fertilization, cryopreservation of gametes and embryos, PGT, and embryo transfer, all of which can potentially disturb normal genomic imprinting [88].
Although it is not clear which aspects of the ART procedures are responsible for disturbed imprinting in humans, information has been obtained from studying mouse models. Several studies demonstrated DNA methylation errors in oocytes, embryos, and placentas following induced superovulation in mice [90,91,92,93]. Chen et al. reported that superovulation (but not IVF or IVM) had a significant effect on the mRNA expression and methylation level of paternally imprinted growth factor receptor-bound protein 10 (Grb10) and H19 (the gene regulating the production of noncoding RNAs) [91]. Grb10 regulates fetal and placental development, while after birth its expression is associated with the pathogenesis of Type 2 diabetes and Alzheimer’s disease [94,95].
Other studies confirmed the negative effects of hyperstimulation on imprinted genes H19, insulin-like growth factor 2 (Igf2), and small nuclear ribonucleoprotein polypeptide N (Snrpn) in mouse blastocysts [96,97]. The Igf2-H19 locus encodes critical paternally imprinted genes that regulate normal embryonic development. The removal of hypomethylation of the Igf2-H19 locus plays a significant role in orchestrating quiescence of pluripotent stem cells in adult organisms, and is probably involved in the regulation of lifespan [98].
Other components of ART procedures have been investigated in human and animal models. Altered methylation at KvDMRI (CpG island) was confirmed in the placenta and fetuses from cryopreserved mouse embryos [99].

4.1. Effects of Cryopreservation and Vitrification on Epigenetic Alteration

Gamete and embryo cryopreservation have become important techniques in ART facilities. In this process, embryos in a high concentration of cryoprotectant are plunged into liquid nitrogen. Despite the significant progress in vitrification techniques, there are still concerns regarding their short and long-term effects on embryos and the resulting offspring [100,101]. It has been demonstrated that cryopreservation is associated with increased DNA fragmentation and apoptotic gene expression [102], as well as a decreased rate of blastocyst hatching in mouse embryos [103]. Epigenetic disorders have also been shown in human and animal models as a result of using cryopreservation techniques. Chen et al. showed that oocyte vitrification causes epigenetic instabilities in bovine embryos. DNA methylation and H3K9me3 levels in oocytes and early cleavage embryos were lower, while the levels of acH3K9 were higher in vitrified embryos than in the controls [104]. While mouse embryo vitrification from the two-cell or eight-cell to blastocyst stages may result in reduced levels of global DNA methylation, vitrification seems to have no significant effects on the DNA methylation status of H19/IGF2 differentially methylation region (DMR) in human blastocysts [105]. Studies on human placentas from frozen-thawed embryo transfers (FET) have identified imprinting disorders that may be associated with the positive regulation of gene expression, growth, development, cell migration, and type II diabetes mellitus [106].

4.2. Effects of Culture Media on Epigenetic Alteration

The culture medium provides an environment for the developing embryo after fertilization, when much of the genome is demethylated. The methylation loss of imprinted genes is believed to result in fetal abnormalities [107,108]. Several mechanisms have been postulated for the aberrant imprinting that occurred in cultured embryos, such as alterations in the expression and subcellular localization of DNA methyltransferases (DNMTs) that are essential for imprint maintenance [84].
In vitro culture is probably the most relevant factor in the alterations of epigenetic reprogramming and development of embryos produced by ART. The importance of culture media in epigenetic preimplantation reprogramming and its impact on early embryo development has been confirmed by an increasing number of studies [109,110,111]. The suboptimal conditions of IVC exert effects on trophoblast development [112]. In mice, the disrupted epigenetic profile of trophoblast is maintained in the placenta, which seems to be more sensitive than the embryo to the IVC environment [113].
Epigenetic modifications in the preimplantation period have been linked to unbalanced fetal–placental development, abnormal fetal growth, and altered metabolic responses. Some studies have confirmed these findings, suggesting cellular aberrations in the placenta and fetus are connected to alterations in gene expression and the association of epigenetic disruption with glucose metabolism and fetal growth in mice [114,115].
Culture medium supplemented by serum has been linked to abnormal skeleton and organ development [116]. Mouse offspring from IVC embryos exhibit anxiety, special memory, and psychomotor activity alterations. The type of culture medium also has a certain impact on the extent of these behavioral changes [28].

4.3. Effects of In Vitro Maturation (IVM) on Epigenetic Alteration

IVM of oocytes is associated with genetic alterations [117]. In vitro matured germline vesicle oocytes show methylation gain at the H19 locus and a methylation loss at Igf2R and Mest/Peg1 locus compared to those obtained from the ovary [118]. In pigs, IVM oocytes have a decreased epigenetic competence and a reduced ability to transform sperm chromatin into the male pronucleus. Reduced active demethylation and histone H4 acetylation epigenetic competence in the male pronucleus (PN) in IVM zygotes has been demonstrated through the analysis of global DNA methylation in the late PN stage [119]. An altered methylation pattern of the H19 differentially methylated regions, which are normally unmethylated in maternal alleles, was detected in 5 out of 20 (25%) human IVM oocytes [120]. In another human study, one paternally-imprinted (GTL2) and three maternally-methylated (LIT1, SNRPN, and PEG3) imprinted genes were analyzed in 71 in vitro- and 38 in vivo matured oocytes. In vitro matured oocytes were retrieved from subjects with PCOS in minimally stimulated cycles without priming with human chorionic gonadotropin (hCG). Single-cell methylation analysis (using limiting dilution bisulfite sequencing) demonstrated no significant increase in imprinting mutations at GTL2 and LIT1, SNRPN, and PEG3 in IVM oocytes [121]. Pluisch and colleagues examined the possible transmission of epigenetic defects to the next generation. They measured some developmentally important genes and interspersed repeats in 11 human IVM offspring and 19 conventional ART-conceived counterparts as control. The results showed that IVM did not exert any significant effects on the chorionic villus and cord blood DNA methylation [122].

4.4. Effects of ICSI on Epigenetic Alteration

Studies on different animal models conceived by ICSI have shown an asynchronous remodeling of chromatin decondensation of the male pronucleus in primates, cattle, and mice [123,124,125]. ICSI-conceived mice compared to conventional IVF-conceived counterparts present long-lasting transcriptome disturbances that are maintained until the neonatal stage. However, these alterations have not been associated with changes in the phenotypic profile or with transgenerational effects [126]. Mouse blastocysts from ICSI have a reduced number of inner mass cells and significant differences in gene expression related to cell function, development, and metabolism [52]. These discrepancies are a result of the different strategies used to activate the oocyte and guarantee embryo development and the protocols used for ICSI in different species [127].

5. ART and Genomic Imprinting in the Placenta

The placenta regulates the growth and development of the embryo during gestation by establishing a bilateral connection, to control the exchange of nutrients and gas between the mother and fetus [128]. A series of imprinted genes are highly expressed in the placenta and play critical roles in its development. These genes take part in the growth, morphology, and nutrient transport capacity of the placenta, which control the nutrient supply for fetal growth [129]. Studies on transgenic mouse models have confirmed roles for imprinted genes in placental function and fetal development. It has been demonstrated that Ascl2, Phlda2, and Peg10 are important for proper placental morphology and function [130]. Igf2 has been indicated to take part in nutrient control, placental size, and morphology [130].
Studies evaluating the effects of ART on placenta development in humans have demonstrated that alterations in DNA methylation may result in dysregulated gene expression. Sakian and colleagues examined human ART-derived placentas and reported alterations in H19 and Igf2 gene expression, due to a loss of imprinting on the paternal allele [131]. It has also been indicated that, compared to naturally conceived human pregnancies, ART-derived placentas show an increased expression of MEST and SERPINF1 genes, a decreased gene expression of COPG2 and NNT genes, a loss of methylation in GRB10, MEST, PEG3, SERPINF1, SCL22A2, and no alterations in the methylation or expression of DLK1, GNAS, H19, and KCNQ10T1. Since these genes play roles in the development and differentiation of adipocytes, insulin signaling, and obesity, altered expression of these genes in the placenta could suggest an increased risk of metabolic syndrome in adulthood [132]. Shi et. al. reported an altered H19/Igf2 ICR methylation pattern in placentas from three IVF-conceived pregnancies. Nonetheless, these children appeared healthy. It was hypothesized that the methylation disruption was due to imprinting errors in the gametes or errors that occur during embryo culture [133]. Nelissen and colleagues observed a reduced DNA methylation in H19 and PEG1 and enhanced expression of H19 and PHLDA2 in human placentas from IVF/ICSI pregnancies, but they did not observe significant differences in body weight in the studied children [134].
Although some of these changes may be due to ART procedures, it is possible that these epigenetic alterations may be a result of the underlying infertility of the couples. Gamete’s methylation pattern can be altered prior to fertilization, which can influence gene expression, exerting effects on offspring phenotype [135,136]. Since human ART studies are affected by confounding factors, conducting studies on animal models has proven to be critical in determining the effects of ART. The results from animal studies have demonstrated that embryo culture and ART procedures can alter epigenetic gene regulation and cause placental and fetal abnormalities. In large animal studies, cattle and sheep, derived from in vitro embryo culture, often present with overgrowth syndrome and sheep with large offspring syndrome had abnormal methylation and reduced expression of the Igf2r gene [137].
Mouse models used to examine the effect of ART on placentation and development revealed that superovulation and IVF as well as oxygen exposure during embryo culture, can induce morphological and epigenetic alterations in the placenta by disturbing the methylation of the ICR of the H19/Igf2, SNRPN, PEG3, and KCNQ10T1 loci [138,139]. Taken together, optimizing ART procedures may decrease adverse effects, by minimizing the occurrence of placental abnormalities and epigenetic alterations.

6. ART and Underlying Parental Characteristics

Adverse perinatal outcomes associated with ART are related to the subfertility of couples, multiple pregnancies, and ART procedures [140]. It is well-documented that multiple pregnancies following multiple embryo transfers are the major cause of preterm delivery and low birth weight, with long-term health risks [141]. In addition, compared to spontaneous conception, singleton pregnancies from ART are more often associated with adverse obstetric and perinatal outcomes [142,143]. These include antepartum hemorrhage, low birth weight, preterm birth, stillbirth, small for gestational age babies, perinatal mortality, need for neonatal intensive care, pregnancy-related hypertensive disorders, preterm rupture of membranes, gestational diabetes, labor induction and Cesarean section [142,143,144,145,146,147].
It is not clear whether the ART procedures or the underlying parental characteristics or genetics are the main cause of the worse obstetric and perinatal outcomes. A large study by Seggers and colleagues using siblingship analysis showed that maternal characteristics, such as age and subfertility, but not IVF treatment, are linked with lower birth weight in IVF-conceived babies [148]. A systematic review and meta-analysis suggested that the risk of preterm birth and small for gestational age babies after IVF is higher in women with endometriosis compared to women without endometriosis [149]. A cohort study reported a higher risk for preterm birth and large for gestational age babies among IVF-conceived newborns from women with PCOS [150]. However, other studies indicated that women with unexplained infertility did not show a higher risk of adverse birth outcomes after IVF [146,151,152].
Huiting Yu et al. demonstrated that 9480 IVF-conceived singletons were at higher risk for adverse birth outcomes compared to singletons in the general population of Shanghai. Underlying factors of infertility, such as endometriosis, PCOS, semen abnormalities, and uterine factors, were linked to preterm birth and abnormal birth weight [153]. Another study demonstrated that male-factor infertility is correlated with low birth weight (LBW) at full term in ART-conceived singletons [154]. Uterine factors, such as uterine inflammation, anatomical defects, impaired receptivity, and cervical insufficiency, have been shown to be associated with an increased risk of preterm birth and low birthweight [155,156]. Some studies have reported that tubal factor infertility is associated with a higher risk of miscarriage, perinatal outcome preterm birth (PTB), and low birth weight (LBW) for singletons conceived by ART; the etiological reasons are mainly attributed to infections and inflammation [157,158,159].
The association between male factor infertility and increased risk of birth defects has not been completely determined. It has been demonstrated that semen abnormality is associated with a higher incidence of LBW, but not PTB. It has also been suggested that sperm abnormality, asthenozoospermia, and oligozoospermia are associated with an increased risk of LBW [153]. On the contrary, most other studies found no significant association between infertility and a higher risk of PTB and LBW compared to unexplained infertility [157]. Moreover, semen parameters did not affect embryo quality or live birth outcomes [160]. These results may partly be interpreted by ART-related aspects, such as improved endometrial receptivity, embryos with higher quality surviving the freezing–thawing process, and the effect of cryoprotectants [137,161]. ART procedures, such as in vitro culture, gametes and embryo manipulation, and freezing and thawing procedures during the preimplantation period can cause epigenetic alteration resulting in birth defects. Currently, the underlying mechanism is not fully understood, and epigenetic effects in ART-conceived offspring require further investigation. While perinatal outcomes, such as LGA and macrosomia, may not be a serious threat to infant survival, they are correlated with an increased risk of cardiovascular disease in adulthood [162,163]. Thus, the association between ART and perinatal outcome should be carefully investigated and longer-term follow-up of children’s development is also important.

7. ART and Congenital Malformations

IVF-conceived infants have an increased risk for neural tube defects, omphalocele, hypospadias, and alimentary tract artesia [164]. The increased risk of birth defects after ART may be associated with ovarian hyperstimulation [165] IVF or ICSI, in vitro culture of gametes and embryos, possible damage during preimplantation genetic testing (PGT), freezing–thawing, luteal support, and embryo transfer [166,167,168]. Some studies have shown a higher rate of malformations after ICSI compared to IVF [169,170], whereas others did not [171]. Luke et al., reported that the rate of major chromosomal birth defects in ART-conceived infants is 18% higher than in naturally conceived counterparts [172].
Some studies have reported a significantly increased risk of birth defects among ART-conceived babies [146,173,174,175,176].
The maternal age of ART patients is considerably higher than that of fertile women, and consequently subfertility in itself might account for the increased risk of birth defects [177]. Therefore, it cannot be ruled out that the underlying infertility, and not the ART procedures, are responsible for the higher rates of birth defects [151,178].

8. ART and Neurological Disorders

Different ART procedures could have different effects on neurodevelopment. La Rovere at al. reviewed the possible effects of epigenetic alterations (due to hormone exposure, gametes preparation, gametes, and embryo freezing, use of culture media, growth conditions for embryos, etc.) on the neurological development of the fetus [179].
Animal studies suggest the possibility of long-term and transgenerational effects of ART on the behaviour and neurodevelopment of adult mice [28]. Adult mice derived from in vitro cultured embryos showed hippocampus and behavioral alterations [28]. H19 mRNA expression was slightly decreased in blastocysts cultured in fetal calf serum supplemented medium, and the mice from these embryos showed anxiety and memory deficiencies at adulthood [12]. Blastomere biopsy results in impaired spatial learning, enhanced neuron degeneration, and altered expression of proteins involved in neural degeneration in aged mice in comparison to aged control mice [180]. These studies suggest that ART might affect the development of the nervous system, but the data from these studies are often inconsistent. A follow up to 12 years demonstrated an increased risk for cerebral palsy (CP) in ART-conceived children [181]. These results were confirmed by Lindegaard et. al., reporting an 80% increased risk of CP in children from IVF [182]. Subsequently, other authors, while confirming the increased risk of CP in children from ART, demonstrated that, following adjustment for confounding factors such as maternal age, multiple pregnancies, and preterm birth, this association became less evident [183,184]. Zhu et. al. demonstrated a higher risk of CP in ART-conceived children even after adjustment for multiple pregnancies and preterm delivery [185]. A more careful analysis of potential factors associated with both IVF and CP and following adjustments for maternal age, year of birth, parity, and smoking, concluded that ART is associated with only a moderately increased risk for CP, and even this might result from increased neonatal morbidity associated with multiple births [183].
Several studies have addressed the possible risk of intellectual disability in ART-conceived babies. Minimally increased risk for intellectual disability was reported, when gestational age, birth weight, socioeconomic and parental educational status were all taken in account [186,187,188,189].
As for the association between ART and increased risk of autism or autism spectrum disorder (ASD), the results are conflicting. Maimburg et. al. demonstrated a lower risk for infantile autism among ART-conceived children, even after adjusting risk factors related to ART and infantile autism [190]. Another study reported a slightly increased rate of ASD among ART-conceived children, which disappeared following the adjustment for confounding factors such as maternal age, parity, educational level, smoking, multiplicity, and birth weight [191]. Kissin and colleagues found a higher incidence of ASD during the first five years of life in ICSI-conceived children compared to IVF-conceived counterparts [192]. A meta-analysis including three cohort studies and eight case-control studies did not show an increased risk of ASD in ART-conceived children [193].

9. ART and Cancer Risk

A systematic meta-analysis by Hargreaves et al. revealed an increased overall cancer risk in children born after infertility treatment. The results showed that the offspring of women with fertility problems had a significantly increased risk for leukemia in childhood and for cancer of the endocrine glands in young adulthood [194]. In a meta-analysis by Chiavarini et al. the possible association between ART and childhood cancer was investigated, and an increased risk of all cancers in ART-conceived children was found [195] In particular, an increased risk of hematological cancers, leukemias, sarcomas, and hepatic cancers was observed in ART-conceived children. In contrast, no significant association between neuroblastoma, retinoblastoma, and solid tumors was reported [195]. Similarly, another meta-analysis reported an increased risk of all cancers, leukemias, hematological cancers, and neural cancers for ART-conceived children [194]. The results by Wang et. al. demonstrated that children conceived by ART had a significantly increased risk of all cancers, leukemia, hematological malignancies, and hepatic tumors [196].
The results indicating increased risk for specific cancer types are inconsistent across studies [197,198,199]. These studies had limitations, such as short follow-up duration, small sample size, restriction to a general population comparison group, and lack of adjustment for confounding factors. It has been shown by a large-scale study with 21 years of follow-up that overall cancer risk in ART-conceived children is not higher, neither when compared with naturally conceived babies born to sub-fertile women, nor when compared with the general population [200]. In a British study among 12,137 ART-conceived children with 7.9 years of follow-up, the overall cancer risk was not higher compared with the general population [197]. In a Swedish cohort study by Kallen et. al., no increased risk for cancer was observed among ICSI-conceived babies [201].
Studies that investigated the association between ART and cancer risks have several limitations. Since cancer in children and young adults is rare, the number was rather low among subgroups, despite the large sample size and long follow-up of the cohorts. Therefore, the reported not significantly increased risks must be interpreted carefully. Confounding factors, such as parental subfertility, child’s birth year, medical record information about ovulation induction, intrauterine insemination, and using expert knowledge about the clinical practice at the time, should be considered when the association of ART with cancer risk among ART-conceived children is investigated.

10. ART and Cardio Metabolic Disorders

Early epigenetic alterations resulting from ART may lead to adverse perinatal outcomes and chronic cardiometabolic disease in adulthood [202,203,204]. Endothelial dysfunction and increased stiffness of vasculature [205] and higher systolic blood pressure [206], as well as disrupted functions of fatty-acid metabolism-related enzymes [77] and altered glucose parameters [114,207], have been reported in ART-conceived mice. Human studies have also demonstrated elevated blood pressure and impaired vascular function [208,209], abnormal retinal vessel morphology, congenital heart defects and altered protein expression profile in the umbilical veins of ART-conceived children [210,211]. Furthermore, ART pregnancies are often associated with preterm birth and low birth weight babies, and it has been shown that small for gestational age babies were at increased risk of ischemic heart, and infants who are born preterm or with very low birth weight have modestly higher systolic blood pressure later in life [212,213]. A systematic review and meta-analysis by Gou et. al. included data from 19 studies (both singleton and twin pregnancies included) that investigated blood pressure, cardiovascular function, adiposity, and metabolism during childhood, puberty, and early adulthood [214]. The results showed a slightly significantly higher blood pressure, increased vessel thickness, and suboptimal cardiac diastolic function in ART-conceived offspring [214].
It has been demonstrated that children conceived by ART showed differences from controls in systemic circulation, artery structure, and systolic pulmonary artery pressure in a hypoxic situation [215]. In line with this study, Chen et. al. showed that IVF-conceived offspring develop increased blood pressure when fed with a high-caloric diet [216]. Ceelen et. al. reported increased glucose levels in pubertal IVF-conceived offspring independently of parental characteristics (subfertility, body weight, and maternal age) or early life factors (gestational age and birth weight) [217]. Chen and colleagues demonstrated only reduced peripheral insulin sensitivity in young adults from IVF. When the offspring were challenged with a high–caloric diet, they showed high fasting blood sugar, glucose intolerance, and insulin resistance [218]. However, other studies found no significant alterations in glucose metabolism after IVF [217,218]. Moreover, it has been shown that the ICSI procedure amplifies the sexual dimorphism in body fat accumulation and distribution at puberty (increased central adipose tissue in female offspring, and an increase in peripheral adiposity at pubertal age in males) [219]. Increased levels of triglycerides have been reported in ART-conceived offspring [220]. These studies show that ART procedures could impact lipid and glucose metabolism in the offspring, resulting in a high risk of type 2 diabetes and metabolic syndrome. Furthermore, the alterations appear to be more evident at pubertal age rather than in early childhood, indicating that even if ART-conceived offspring seems healthy at a young age, long-term follow-up studies are necessary, since metabolic alterations could arise later in life and impact the development of adult-onset disorders [221].

11. ART and Altered Immune Functions

Several studies reported altered immune functions and higher rates of immune-related diseases in ART-conceived offspring [222,223]. ART-conceived mice responded less efficiently to vaccines and skewed toward T helper (Th) 2-dominated responses [2,224]. Children born after fresh embryo transfer exhibit an increased risk of immune dysfunction in childhood, manifested by elevated Interleukin-4 (IL-4) serum levels and decreased IFN-ɣ/IL-4 ratio [225]. It has also been shown that the expression of three genes, the endoplasmic reticulum aminopeptidase 2 (ERAP2), kynureninase (KYNU), and signal transducer and activator of transcription 4 (STAT4), all of which are involved in immune responses, was altered in placentae of ART-treated women [226,227]. The increased cancer risk in ART conceived individuals might be an indicator of enhanced immune tolerance to tumor antigens [228].
Although the exact etiology is not fully understood, gene–environment interactions appear to play an important role in the development of asthma and allergy [229,230]. ART may enhance the risk of atopic diseases in the offspring through direct epigenetic alterations [231,232]. At the same time, it is important to keep in mind, that ART pregnancies are associated with preterm birth, low birth weight babies, and Cesarean section, all of which are risk factors for asthma [233,234].
Some studies have investigated the risk of atopic disorders in ART-conceived children. A study by Hart et. al. showed that there is no significant increase in asthma and allergies in ART-conceived offspring [235]. Another review by Kettner et. al. concluded that the data on the risk of asthma and allergies among ART conceived are inconsistent [236]. Otherwise, a number of population-based studies as well as systematic reviews and meta-analyses have demonstrated an increased risk of atopy among ART-conceived children [237,238]. A subgroup analysis on singletons and multiple births showed a statistically significant linkage with the risk of asthma only for the former group [238]. However, only four studies were conducted on multiple births, and the results from the larger one were significant [239].
Infertility itself and ART treatment are stressful for couples [240]. Maternal stress or anxiety could increase blood cortisol levels during pregnancy, and subsequent transplacental passage of cortisol could exert a programming effect on the development of the hypothalamic–pituitary–adrenal (HPA) axis in the fetus, resulting in altered stress response later in life [241,242]. In support of these findings, in our previous study, we showed that the prooxidant/antioxidant ratios were slightly higher in IVF-conceived mice than in naturally conceived counterparts, but the difference was not statistically significant [243]. It has been demonstrated that prenatal maternal stress affects asthma development through incomplete development of respiratory tract and offspring immune dysregulation [244]. In addition, evidence suggests that maternal stress may alter the composition of offspring bacterial microbiota. These alterations may adjust the immune system development and enhance the risk of asthma and allergic disease in offspring [245]. ART-associated risks and complications are summarized in Figure 2.

12. Conclusions and Future Perspectives

The early stages of embryo development are sensitive to the microenvironment, and even slight changes in the microenvironment could result in long-term consequences for fetal, postnatal, and adult health. ART procedures occur during a critically coordinated period of epigenetic reprogramming, thus the preimplantation period is highly vulnerable to epigenetic aberrations, which could result in cardiometabolic alterations, neurological disorders, increased cancer risk, altered immune responses, asthma, and allergy, as well as other chronic health problems developing in ART-conceived children, adolescents, or adults.
Although there is still a debate on the health status of ART-conceived children, this possibility has to be considered. Investigations on the potential side effects of ART on offspring health status have a long way to go. Long-term follow-up of ART-conceived individuals and transgenerational studies are required. Confounding factors, e.g., infertility itself, the higher maternal age of ART patients, and consequences of multiple embryo transfers, need to be eliminated, given that minor alterations can result in a high risk later in life.
Over time, ART technologies have considerably improved. Using less invasive protocols to induce superovulation should lead to better epigenomes in the oocytes [246], and decreased aneuploidy rates of the resulting embryos [247]. Improved superovulation regimens have been developed for individuals with specific conditions, such as PCOS and cancers [248,249].
The perinatal outcomes of ART have become better, mainly as a result of single embryo transfer and frozen-thawed embryo transfer [250,251,252]. Multiple pregnancy is a risk for prematurity and preterm birth. The goal is to achieve a single pregnancy, by transferring a single embryo. This, however, requires better techniques to identify the competent embryo. Up to now, the selection of gametes and embryos was based on morphological and morphogenetic criteria. Selection based on the morphological features of the embryo is highly prone to subjectivity. Morpho-kinetic measurements provide more objective data. Invasive methods, such as pre-implantation genetic testing for aneuploidy, involve certain risks, since biopsy might negatively influence further development of the embryo. Analysis of spent embryo culture media detects changes that would reflect the metabolism and functional state of the embryo [253,254,255,256]. Further methods include analysis of the cumulus cells using polymerase chain reaction (PCR) and/or DNA microarrays to measure the oocyst’s competency [257,258]. New emerging technologies, such as microfluidic and microfabricated devices, may increase the safety and efficiency of ART procedures [259].
It has been demonstrated that any alteration in either the contact materials (glassware, metalware, plastic ware, etc.) or micro-environment conditions (culture media, freezing media, oxygen, CO2, incubator, ambient air, etc.) may exert deleterious effects on the oocyte and resultant embryo quality, consequently enhancing the rate of genetic defects. Therefore, quality assurance and quality control must be considered important steps in decreasing the potential risks associated with ART. Utilization of microfluidic technologies in ART procedures may lead to at least four predictable advantages: (1) exactly controlled gamete or embryo manipulation; (2) establishing a biomimetic environment for culture; (3) promoting microscale genetic and molecular bioassays; and (4) allowing automatization and miniaturization [260].
In spite of the continuously developing technology, the debate on the long-term consequences of ART remains open. Well-controlled epidemiological studies with large sample sizes are necessary. More investigations are also necessary to indicate whether the increased obstetric, perinatal, and health problems observed in ART-conceived children are the direct results of the ART procedure itself or a result of the underlying infertility of the parents. As many parental characteristics cannot be changed, further investigations to identify the optimal ART procedures that improve both perinatal and long-term health status are needed.

Author Contributions

Conception and design of the study, H.A. and J.S.-B.; write the manuscript, L.A.-M. and O.B.N.; literature survey, S.R. and T.C.; Figure and Table designing, H.A.; final editing of the manuscript, J.S.-B. and L.A.-M. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by National Laboratory on Human Reproduction (RRF-2.3.1-21-2022-00012) and the National Research, Development and Innovation Fund of Hungary, financed under the TKP2021-EGA funding scheme (TKP2021-EGA-10).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data availability is not applicable to this article as no new data were created or analysed in this study.

Acknowledgments

We thank Yalda Jahanbani (Student research Committee, Tabriz University of Medical Sciences, Tabriz, Iran & Faculty of Pharmacy, Tabriz University of Medical Sciences, Tabriz, Iran) for his help in designing the figures and the table.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fauser, B.C.; Devroey, P.; Diedrich, K.; Balaban, B.; Bonduelle, M.; Delemarre-van de Waal, H.A.; Estella, C.; Ezcurra, D.; Geraedts, J.P.; Howles, C.M.; et al. Health outcomes of children born after IVF/ICSI: A review of current expert opinion and literature. Reprod. Biomed. Online 2014, 28, 162–182. [Google Scholar] [CrossRef]
  2. Ahmadi, H.; Fathi, F.; Karimi, H.; Amidi, F.; Mehdinejadiani, S.; Moeini, A.; Bahram Rezai, M.; Hoseini, S.; Sobhani, A. Altered TH1, TH2, TH17 balance in assisted reproductive technology conceived mice. J. Reprod. Immunol. 2020, 139, 103117. [Google Scholar] [CrossRef]
  3. Hoorsan, H.; Mirmiran, P.; Chaichian, S.; Moradi, Y.; Hoorsan, R.; Jesmi, F. Congenital Malformations in Infants of Mothers Undergoing Assisted Reproductive Technologies: A Systematic Review and Meta-analysis Study. J. Prev. Med. Public Health 2017, 50, 347–360. [Google Scholar] [CrossRef]
  4. Simpson, J.L. Birth defects and assisted reproductive technologies. Semin. Fetal Neonatal Med. 2014, 19, 177–182. [Google Scholar] [CrossRef]
  5. Moll, A.C.; Imhof, S.M.; Cruysberg, J.R.; Schouten-van Meeteren, A.Y.; Boers, M.; van Leeuwen, F.E. Incidence of retinoblastoma in children born after in-vitro fertilisation. Lancet 2003, 361, 309–310. [Google Scholar] [CrossRef]
  6. Petridou, E.T.; Sergentanis, T.N.; Panagopoulou, P.; Moschovi, M.; Polychronopoulou, S.; Baka, M.; Pourtsidis, A.; Athanassiadou, F.; Kalmanti, M.; Sidi, V.; et al. In vitro fertilization and risk of childhood leukemia in Greece and Sweden. Pediatr. Blood Cancer 2012, 58, 930–936. [Google Scholar] [CrossRef]
  7. Massaro, P.A.; MacLellan, D.L.; Anderson, P.A.; Romao, R.L. Does intracytoplasmic sperm injection pose an increased risk of genitourinary congenital malformations in offspring compared to in vitro fertilization? A systematic review and meta-analysis. J. Urol. 2015, 193, 1837–1842. [Google Scholar] [CrossRef]
  8. Kamel, R.M. Assisted reproductive technology after the birth of louise brown. J. Reprod. Infertil. 2013, 14, 96–109. [Google Scholar] [CrossRef]
  9. Wang, L.Y.; Wang, N.; Le, F.; Li, L.; Lou, H.Y.; Liu, X.Z.; Zheng, Y.M.; Qian, Y.Q.; Chen, Y.L.; Jiang, X.H.; et al. Superovulation Induced Changes of Lipid Metabolism in Ovaries and Embryos and Its Probable Mechanism. PLoS ONE 2015, 10, e0132638. [Google Scholar] [CrossRef]
  10. Carpinello, O.J.; DeCherney, A.H.; Hill, M.J. Developmental Origins of Health and Disease: The History of the Barker Hypothesis and Assisted Reproductive Technology. Semin. Reprod. Med. 2018, 36, 177–182. [Google Scholar] [CrossRef]
  11. Rinaudo, P.; Schultz, R.M. Effects of embryo culture on global pattern of gene expression in preimplantation mouse embryos. Reproduction 2004, 128, 301–311. [Google Scholar] [CrossRef]
  12. Fernández-Gonzalez, R.; Moreira, P.; Bilbao, A.; Jiménez, A.; Pérez-Crespo, M.; Ramírez, M.A.; Rodríguez De Fonseca, F.; Pintado, B.; Gutiérrez-Adán, A. Long-term effect of in vitro culture of mouse embryos with serum on mRNA expression of imprinting genes, development, and behavior. Proc. Natl. Acad. Sci. USA 2004, 101, 5880–5885. [Google Scholar] [CrossRef] [PubMed]
  13. Zegers-Hochschild, F.; Adamson, G.D.; Dyer, S.; Racowsky, C.; de Mouzon, J.; Sokol, R.; Rienzi, L.; Sunde, A.; Schmidt, L.; Cooke, I.D.; et al. The International Glossary on Infertility and Fertility Care, 2017. Hum. Reprod. 2017, 32, 1786–1801. [Google Scholar] [CrossRef]
  14. Cohen, J.; Trounson, A.; Dawson, K.; Jones, H.; Hazekamp, J.; Nygren, K.G.; Hamberger, L. The early days of IVF outside the UK. Hum. Reprod. Update 2005, 11, 439–459. [Google Scholar] [CrossRef]
  15. Lenz, S.; Lauritsen, J.G. Ultrasonically guided percutaneous aspiration of human follicles under local anesthesia: A new method of collecting oocytes for in vitro fertilization. Fertil. Steril. 1982, 38, 673–677. [Google Scholar] [CrossRef] [PubMed]
  16. Trounson, A.; Mohr, L. Human pregnancy following cryopreservation, thawing and transfer of an eight-cell embryo. Nature 1983, 305, 707–709. [Google Scholar] [CrossRef]
  17. Trounson, A.; Peura, A.; Kirby, C. Ultrarapid freezing: A new low-cost and effective method of embryo cryopreservation. Fertil. Steril. 1987, 48, 843–850. [Google Scholar] [CrossRef]
  18. Patrizio, P.; Silber, S.; Ord, T.; Balmaceda, J.P.; Asch, R.H. Two births after microsurgical sperm aspiration in congenital absence of vas deferens. Lancet 1988, 2, 1364. [Google Scholar] [CrossRef]
  19. Palermo, G.; Joris, H.; Devroey, P.; Van Steirteghem, A.C. Pregnancies after intracytoplasmic injection of single spermatozoon into an oocyte. Lancet 1992, 340, 17–18. [Google Scholar] [CrossRef] [PubMed]
  20. Cha, K.Y.; Koo, J.J.; Ko, J.J.; Choi, D.H.; Han, S.Y.; Yoon, T.K. Pregnancy after in vitro fertilization of human follicular oocytes collected from nonstimulated cycles, their culture in vitro and their transfer in a donor oocyte program. Fertil. Steril. 1991, 55, 109–113. [Google Scholar] [CrossRef]
  21. Germond, M.; Dessole, S.; Senn, A.; Loumaye, E.; Howles, C.; Beltrami, V. Successful in-vitro fertilisation and embryo transfer after treatment with recombinant human FSH. Lancet 1992, 339, 1170. [Google Scholar]
  22. Xu, K.; Shi, Z.M.; Veeck, L.L.; Hughes, M.R.; Rosenwaks, Z. First unaffected pregnancy using preimplantation genetic diagnosis for sickle cell anemia. JAMA 1999, 281, 1701–1706. [Google Scholar] [CrossRef]
  23. Boer, K.A.; McArthur, S.; Murray, C.; Jansen, R.P.S. O-54. First live birth following blastocyst biopsy and PGD analysis. Reprod. BioMed. Online 2002, 4, 35. [Google Scholar] [CrossRef]
  24. Barker, D.J. The developmental origins of adult disease. J. Am. Coll. Nutr. 2004, 23, 588s–595s. [Google Scholar] [CrossRef]
  25. Stanner, S.A.; Bulmer, K.; Andrès, C.; Lantseva, O.E.; Borodina, V.; Poteen, V.V.; Yudkin, J.S. Does malnutrition in utero determine diabetes and coronary heart disease in adulthood? Results from the Leningrad siege study, a cross sectional study. BMJ 1997, 315, 1342–1348. [Google Scholar] [CrossRef] [PubMed]
  26. Fleming, T.P.; Lucas, E.S.; Watkins, A.J.; Eckert, J.J. Adaptive responses of the embryo to maternal diet and consequences for post-implantation development. Reprod. Fertil. Dev. 2011, 24, 35–44. [Google Scholar] [CrossRef] [PubMed]
  27. Watkins, A.J.; Ursell, E.; Panton, R.; Papenbrock, T.; Hollis, L.; Cunningham, C.; Wilkins, A.; Perry, V.H.; Sheth, B.; Kwong, W.Y.; et al. Adaptive responses by mouse early embryos to maternal diet protect fetal growth but predispose to adult onset disease. Biol. Reprod. 2008, 78, 299–306. [Google Scholar] [CrossRef] [PubMed]
  28. Ecker, D.J.; Stein, P.; Xu, Z.; Williams, C.J.; Kopf, G.S.; Bilker, W.B.; Abel, T.; Schultz, R.M. Long-term effects of culture of preimplantation mouse embryos on behavior. Proc. Natl. Acad. Sci. USA 2004, 101, 1595–1600. [Google Scholar] [CrossRef]
  29. Kwong, W.Y.; Wild, A.E.; Roberts, P.; Willis, A.C.; Fleming, T.P. Maternal undernutrition during the preimplantation period of rat development causes blastocyst abnormalities and programming of postnatal hypertension. Development 2000, 127, 4195–4202. [Google Scholar] [CrossRef]
  30. Adamiak, S.J.; Mackie, K.; Watt, R.G.; Webb, R.; Sinclair, K.D. Impact of nutrition on oocyte quality: Cumulative effects of body composition and diet leading to hyperinsulinemia in cattle. Biol. Reprod. 2005, 73, 918–926. [Google Scholar] [CrossRef]
  31. Cockburn, K.; Rossant, J. Making the blastocyst: Lessons from the mouse. J. Clin. Investig. 2010, 120, 995–1003. [Google Scholar] [CrossRef]
  32. Dey, S.K.; Lim, H.; Das, S.K.; Reese, J.; Paria, B.C.; Daikoku, T.; Wang, H. Molecular cues to implantation. Endocr. Rev. 2004, 25, 341–373. [Google Scholar] [CrossRef]
  33. Biggers, J.D.; Whittingham, D.G.; Donahue, R.P. The pattern of energy metabolism in the mouse oöcyte and zygote. Proc. Natl. Acad. Sci. USA 1967, 58, 560–567. [Google Scholar] [CrossRef]
  34. Brinster, R.L.; Thomson, J.L. Development of eight-cell mouse embryos in vitro. Exp. Cell Res. 1966, 42, 308–315. [Google Scholar] [CrossRef]
  35. Lane, M.; Gardner, D.K. Lactate regulates pyruvate uptake and metabolism in the preimplantation mouse embryo. Biol. Reprod. 2000, 62, 16–22. [Google Scholar] [CrossRef]
  36. Lane, M.; Gardner, D.K. Ammonium induces aberrant blastocyst differentiation, metabolism, pH regulation, gene expression and subsequently alters fetal development in the mouse. Biol. Reprod. 2003, 69, 1109–1117. [Google Scholar] [CrossRef] [PubMed]
  37. Edwards, L.J.; Williams, D.A.; Gardner, D.K. Intracellular pH of the mouse preimplantation embryo: Amino acids act as buffers of intracellular pH. Hum. Reprod. 1998, 13, 3441–3448. [Google Scholar] [CrossRef] [PubMed]
  38. Gardner, D.K.; Leese, H.J. The role of glucose and pyruvate transport in regulating nutrient utilization by preimplantation mouse embryos. Development 1988, 104, 423–429. [Google Scholar] [CrossRef] [PubMed]
  39. Bermejo-Alvarez, P.; Rizos, D.; Lonergan, P.; Gutierrez-Adan, A. Transcriptional sexual dimorphism during preimplantation embryo development and its consequences for developmental competence and adult health and disease. Reproduction 2011, 141, 563–570. [Google Scholar] [CrossRef]
  40. Feuer, S.; Rinaudo, P. Preimplantation stress and development. Birth Defects Res. C Embryo Today 2012, 96, 299–314. [Google Scholar] [CrossRef]
  41. Nadal, A.; Ropero, A.B.; Laribi, O.; Maillet, M.; Fuentes, E.; Soria, B. Nongenomic actions of estrogens and xenoestrogens by binding at a plasma membrane receptor unrelated to estrogen receptor alpha and estrogen receptor beta. Proc. Natl. Acad. Sci. USA 2000, 97, 11603–11608. [Google Scholar] [CrossRef]
  42. Takai, Y.; Tsutsumi, O.; Ikezuki, Y.; Kamei, Y.; Osuga, Y.; Yano, T.; Taketan, Y. Preimplantation exposure to bisphenol A advances postnatal development. Reprod. Toxicol. 2001, 15, 71–74. [Google Scholar] [CrossRef]
  43. Hassa, H.; Gurer, F.; Tanir, H.M.; Kaya, M.; Gunduz, N.B.; Sariboyaci, A.E.; Bal, C. Effect of cigarette smoke and alpha-tocopherol (vitamin E) on fertilization, cleavage, and embryo development rates in mice: An experimental in vitro fertilization mice model study. Eur. J. Obstet. Gynecol. Reprod. Biol. 2007, 135, 177–182. [Google Scholar] [CrossRef]
  44. Polyzos, A.; Schmid, T.E.; Piña-Guzmán, B.; Quintanilla-Vega, B.; Marchetti, F. Differential sensitivity of male germ cells to mainstream and sidestream tobacco smoke in the mouse. Toxicol. Appl. Pharmacol. 2009, 237, 298–305. [Google Scholar] [CrossRef] [PubMed]
  45. Takenaka, M.; Horiuchi, T.; Yanagimachi, R. Effects of light on development of mammalian zygotes. Proc. Natl. Acad. Sci. USA 2007, 104, 14289–14293. [Google Scholar] [CrossRef]
  46. Gil, M.A.; Maside, C.; Cuello, C.; Parrilla, I.; Vazquez, J.M.; Roca, J.; Martinez, E.A. Effects of Hoechst 33,342 staining and ultraviolet irradiation on mitochondrial distribution and DNA copy number in porcine oocytes and preimplantation embryos. Mol. Reprod. Dev. 2012, 79, 651–663. [Google Scholar] [CrossRef] [PubMed]
  47. Takahashi, M.; Saka, N.; Takahashi, H.; Kanai, Y.; Schultz, R.M.; Okano, A. Assessment of DNA damage in individual hamster embryos by comet assay. Mol. Reprod. Dev. 1999, 54, 1–7. [Google Scholar] [CrossRef]
  48. Oh, S.J.; Gong, S.P.; Lee, S.T.; Lee, E.J.; Lim, J.M. Light intensity and wavelength during embryo manipulation are important factors for maintaining viability of preimplantation embryos in vitro. Fertil. Steril. 2007, 88, 1150–1157. [Google Scholar] [CrossRef]
  49. Bognar, Z.; Csabai, T.J.; Pallinger, E.; Balassa, T.; Farkas, N.; Schmidt, J.; Görgey, E.; Berta, G.; Szekeres-Bartho, J.; Bodis, J. The effect of light exposure on the cleavage rate and implantation capacity of preimplantation murine embryos. J. Reprod. Immunol. 2019, 132, 21–28. [Google Scholar] [CrossRef]
  50. Bódis, J.; Gödöny, K.; Várnagy, Á.; Kovács, K.; Koppán, M.; Nagy, B.; Erostyák, J.; Herczeg, R.; Szekeres-Barthó, J.; Gyenesei, A.; et al. How to Reduce the Potential Harmful Effects of Light on Blastocyst Development during IVF. Med. Princ. Pract. 2020, 29, 558–564. [Google Scholar] [CrossRef] [PubMed]
  51. Delle Piane, L.; Lin, W.; Liu, X.; Donjacour, A.; Minasi, P.; Revelli, A.; Maltepe, E.; Rinaudo, P.F. Effect of the method of conception and embryo transfer procedure on mid-gestation placenta and fetal development in an IVF mouse model. Hum. Reprod. 2010, 25, 2039–2046. [Google Scholar] [CrossRef]
  52. Giritharan, G.; Li, M.W.; Di Sebastiano, F.; Esteban, F.J.; Horcajadas, J.A.; Lloyd, K.C.; Donjacour, A.; Maltepe, E.; Rinaudo, P.F. Effect of ICSI on gene expression and development of mouse preimplantation embryos. Hum. Reprod. 2010, 25, 3012–3024. [Google Scholar] [CrossRef]
  53. Valenzuela, O.A.; Couturier-Tarrade, A.; Choi, Y.H.; Aubrière, M.C.; Ritthaler, J.; Chavatte-Palmer, P.; Hinrichs, K. Impact of equine assisted reproductive technologies (standard embryo transfer or intracytoplasmic sperm injection (ICSI) with in vitro culture and embryo transfer) on placenta and foal morphometry and placental gene expression. Reprod. Fertil. Dev. 2018, 30, 371–379. [Google Scholar] [CrossRef]
  54. Giritharan, G.; Delle Piane, L.; Donjacour, A.; Esteban, F.J.; Horcajadas, J.A.; Maltepe, E.; Rinaudo, P. In vitro culture of mouse embryos reduces differential gene expression between inner cell mass and trophectoderm. Reprod. Sci. 2012, 19, 243–252. [Google Scholar] [CrossRef]
  55. Kleijkers, S.H.; Eijssen, L.M.; Coonen, E.; Derhaag, J.G.; Mantikou, E.; Jonker, M.J.; Mastenbroek, S.; Repping, S.; Evers, J.L.; Dumoulin, J.C.; et al. Differences in gene expression profiles between human preimplantation embryos cultured in two different IVF culture media. Hum. Reprod. 2015, 30, 2303–2311. [Google Scholar] [CrossRef]
  56. Rinaudo, P.F.; Giritharan, G.; Talbi, S.; Dobson, A.T.; Schultz, R.M. Effects of oxygen tension on gene expression in preimplantation mouse embryos. Fertil. Steril. 2006, 86, 1265.e1–1265.e36. [Google Scholar] [CrossRef]
  57. Leese, H.J.; Baumann, C.G.; Brison, D.R.; McEvoy, T.G.; Sturmey, R.G. Metabolism of the viable mammalian embryo: Quietness revisited. Mol. Hum. Reprod. 2008, 14, 667–672. [Google Scholar] [CrossRef]
  58. Consensus Group, C. ‘There is only one thing that is truly important in an IVF laboratory: Everything’ Cairo Consensus Guidelines on IVF Culture Conditions. Reprod. Biomed. Online 2020, 40, 33–60. [Google Scholar] [CrossRef]
  59. Khodavirdilou, R.; Pournaghi, M.; Oghbaei, H.; Rastgar Rezaei, Y.; Javid, F.; Khodavirdilou, L.; Shakibfar, F.; Latifi, Z.; Hakimi, P.; Nouri, M.; et al. Toxic effect of light on oocyte and pre-implantation embryo: A systematic review. Arch. Toxicol. 2021, 95, 3161–3169. [Google Scholar] [CrossRef]
  60. Kolahi, K.S.; Donjacour, A.; Liu, X.; Lin, W.; Simbulan, R.K.; Bloise, E.; Maltepe, E.; Rinaudo, P. Effect of substrate stiffness on early mouse embryo development. PLoS ONE 2012, 7, e41717. [Google Scholar] [CrossRef]
  61. Argyle, C.E.; Harper, J.C.; Davies, M.C. Oocyte cryopreservation: Where are we now? Hum. Reprod. Update 2016, 22, 440–449. [Google Scholar] [CrossRef] [PubMed]
  62. Bielanski, A.; Vajta, G. Risk of contamination of germplasm during cryopreservation and cryobanking in IVF units. Hum. Reprod. 2009, 24, 2457–2467. [Google Scholar] [CrossRef] [PubMed]
  63. Lavara, R.; Baselga, M.; Marco-Jiménez, F.; Vicente, J.S. Long-term and transgenerational effects of cryopreservation on rabbit embryos. Theriogenology 2014, 81, 988–992. [Google Scholar] [CrossRef] [PubMed]
  64. Grissom, N.M.; George, R.; Reyes, T.M. Suboptimal nutrition in early life affects the inflammatory gene expression profile and behavioral responses to stressors. Brain Behav. Immun. 2017, 63, 115–126. [Google Scholar] [CrossRef]
  65. Kawwass, J.F.; Badell, M.L. Maternal and Fetal Risk Associated with Assisted Reproductive Technology. Obstet. Gynecol. 2018, 132, 763–772. [Google Scholar] [CrossRef] [PubMed]
  66. Davies, M.J.; Rumbold, A.R.; Marino, J.L.; Willson, K.; Giles, L.C.; Whitrow, M.J.; Scheil, W.; Moran, L.J.; Thompson, J.G.; Lane, M.; et al. Maternal factors and the risk of birth defects after IVF and ICSI: A whole of population cohort study. BJOG 2017, 124, 1537–1544. [Google Scholar] [CrossRef]
  67. Cabry, R.; Merviel, P.; Madkour, A.; Lefranc, E.; Scheffler, F.; Desailloud, R.; Bach, V.; Benkhalifa, M. The impact of endocrine disruptor chemicals on oocyte/embryo and clinical outcomes in IVF. Endocr. Connect. 2020, 9, R134–R142. [Google Scholar] [CrossRef]
  68. Krisher, R.L. In vivo and in vitro environmental effects on mammalian oocyte quality. Annu. Rev. Anim. Biosci. 2013, 1, 393–417. [Google Scholar] [CrossRef]
  69. Walters, E.; Edwards, R.G. On a fallacious invocation of the Barker hypothesis of anomalies in newborn rats due to mothers’ food restriction in preimplantation phases. Reprod. Biomed. Online 2003, 7, 580–582. [Google Scholar] [CrossRef]
  70. Neitzke, U.; Harder, T.; Schellong, K.; Melchior, K.; Ziska, T.; Rodekamp, E.; Dudenhausen, J.W.; Plagemann, A. Intrauterine growth restriction in a rodent model and developmental programming of the metabolic syndrome: A critical appraisal of the experimental evidence. Placenta 2008, 29, 246–254. [Google Scholar] [CrossRef]
  71. Jansson, T.; Lambert, G.W. Effect of intrauterine growth restriction on blood pressure, glucose tolerance and sympathetic nervous system activity in the rat at 3-4 months of age. J. Hypertens. 1999, 17, 1239–1248. [Google Scholar] [CrossRef]
  72. Harlow, G.M.; Quinn, P. Development of preimplantation mouse embryos in vivo and in vitro. Aust. J. Biol. Sci. 1982, 35, 187–193. [Google Scholar] [CrossRef]
  73. Schwarzer, C.; Esteves, T.C.; Araúzo-Bravo, M.J.; Le Gac, S.; Nordhoff, V.; Schlatt, S.; Boiani, M. ART culture conditions change the probability of mouse embryo gestation through defined cellular and molecular responses. Hum. Reprod. 2012, 27, 2627–2640. [Google Scholar] [CrossRef]
  74. Angiolini, E.; Fowden, A.; Coan, P.; Sandovici, I.; Smith, P.; Dean, W.; Burton, G.; Tycko, B.; Reik, W.; Sibley, C.; et al. Regulation of placental efficiency for nutrient transport by imprinted genes. Placenta 2006, 27 (Suppl. A), S98–S102. [Google Scholar] [CrossRef]
  75. Bloise, E.; Lin, W.; Liu, X.; Simbulan, R.; Kolahi, K.S.; Petraglia, F.; Maltepe, E.; Donjacour, A.; Rinaudo, P. Impaired placental nutrient transport in mice generated by in vitro fertilization. Endocrinology 2012, 153, 3457–3467. [Google Scholar] [CrossRef]
  76. Fernández-Gonzalez, R.; Moreira, P.N.; Pérez-Crespo, M.; Sánchez-Martín, M.; Ramirez, M.A.; Pericuesta, E.; Bilbao, A.; Bermejo-Alvarez, P.; de Dios Hourcade, J.; de Fonseca, F.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, 761–772. [Google Scholar] [CrossRef]
  77. Scott, K.A.; Yamazaki, Y.; Yamamoto, M.; Lin, Y.; Melhorn, S.J.; Krause, E.G.; Woods, S.C.; Yanagimachi, R.; Sakai, R.R.; Tamashiro, K.L. Glucose parameters are altered in mouse offspring produced by assisted reproductive technologies and somatic cell nuclear transfer. Biol. Reprod. 2010, 83, 220–227. [Google Scholar] [CrossRef]
  78. Watkins, A.J.; Platt, D.; Papenbrock, T.; Wilkins, A.; Eckert, J.J.; Kwong, W.Y.; Osmond, C.; Hanson, M.; Fleming, T.P. Mouse embryo culture induces changes in postnatal phenotype including raised systolic blood pressure. Proc. Natl. Acad. Sci. USA 2007, 104, 5449–5454. [Google Scholar] [CrossRef]
  79. Wu, C.; Morris, J.R. Genes, genetics, and epigenetics: A correspondence. Science 2001, 293, 1103–1105. [Google Scholar] [CrossRef]
  80. Santos, F.; Dean, W. Epigenetic reprogramming during early development in mammals. Reproduction 2004, 127, 643–651. [Google Scholar] [CrossRef]
  81. Kiefer, J.C. Epigenetics in development. Dev. Dyn. 2007, 236, 1144–1156. [Google Scholar] [CrossRef] [PubMed]
  82. Henningsen, A.A.; Gissler, M.; Rasmussen, S.; Opdahl, S.; Wennerholm, U.B.; Spangsmose, A.L.; Tiitinen, A.; Bergh, C.; Romundstad, L.B.; Laivuori, H.; et al. Imprinting disorders in children born after ART: A Nordic study from the CoNARTaS group. Hum. Reprod. 2020, 35, 1178–1184. [Google Scholar] [CrossRef]
  83. Eroglu, A.; Layman, L.C. Role of ART in imprinting disorders. Semin. Reprod. Med. 2012, 30, 92–104. [Google Scholar] [CrossRef] [PubMed]
  84. Owen, C.M.; Segars, J.H., Jr. Imprinting disorders and assisted reproductive technology. Semin. Reprod. Med. 2009, 27, 417–428. [Google Scholar] [CrossRef]
  85. Cortessis, V.K.; Azadian, M.; Buxbaum, J.; Sanogo, F.; Song, A.Y.; Sriprasert, I.; Wei, P.C.; Yu, J.; Chung, K.; Siegmund, K.D. Comprehensive meta-analysis reveals association between multiple imprinting disorders and conception by assisted reproductive technology. J. Assist. Reprod. Genet. 2018, 35, 943–952. [Google Scholar] [CrossRef]
  86. Amor, D.J.; Halliday, J. A review of known imprinting syndromes and their association with assisted reproduction technologies. Hum. Reprod. 2008, 23, 2826–2834. [Google Scholar] [CrossRef] [PubMed]
  87. Horánszky, A.; Becker, J.L.; Zana, M.; Ferguson-Smith, A.C.; Dinnyés, A. Epigenetic Mechanisms of ART-Related Imprinting Disorders: Lessons From iPSC and Mouse Models. Genes 2021, 12, 1704. [Google Scholar] [CrossRef] [PubMed]
  88. Chi, F.; Zhao, M.; Li, K.; Lin, A.Q.; Li, Y.; Teng, X. DNA methylation status of imprinted H19 and KvDMR1 genes in human placentas after conception using assisted reproductive technology. Ann. Transl. Med. 2020, 8, 854. [Google Scholar] [CrossRef]
  89. Lazaraviciute, G.; Kauser, M.; Bhattacharya, S.; Haggarty, P.; Bhattacharya, S. A systematic review and meta-analysis of DNA methylation levels and imprinting disorders in children conceived by IVF/ICSI compared with children conceived spontaneously. Hum. Reprod. Update 2014, 20, 840–852. [Google Scholar] [CrossRef]
  90. Yu, B.; Smith, T.H.; Battle, S.L.; Ferrell, S.; Hawkins, R.D. Superovulation alters global DNA methylation in early mouse embryo development. Epigenetics 2019, 14, 780–790. [Google Scholar] [CrossRef]
  91. Chen, X.; Huang, Y.; Huang, H.; Guan, Y.; Li, M.; Jiang, X.; Yu, M.; Yang, X. Effects of superovulation, in vitro fertilization, and oocyte in vitro maturation on imprinted gene Grb10 in mouse blastocysts. Arch. Gynecol. Obstet. 2018, 298, 1219–1227. [Google Scholar] [CrossRef]
  92. Sato, A.; Otsu, E.; Negishi, H.; Utsunomiya, T.; Arima, T. Aberrant DNA methylation of imprinted loci in superovulated oocytes. Hum. Reprod. 2007, 22, 26–35. [Google Scholar] [CrossRef] [PubMed]
  93. Fortier, A.L.; Lopes, F.L.; Darricarrère, N.; Martel, J.; Trasler, J.M. Superovulation alters the expression of imprinted genes in the midgestation mouse placenta. Hum. Mol. Genet. 2008, 17, 1653–1665. [Google Scholar] [CrossRef] [PubMed]
  94. Xie, J.; Wei, Q.; Deng, H.; Li, G.; Ma, L.; Zeng, H. Negative regulation of Grb10 Interacting GYF Protein 2 on insulin-like growth factor-1 receptor signaling pathway caused diabetic mice cognitive impairment. PLoS ONE 2014, 9, e108559. [Google Scholar] [CrossRef]
  95. Chaudhry, M.; Wang, X.; Bamne, M.N.; Hasnain, S.; Demirci, F.Y.; Lopez, O.L.; Kamboh, M.I. Genetic variation in imprinted genes is associated with risk of late-onset Alzheimer’s disease. J. Alzheimers Dis. 2015, 44, 989–994. [Google Scholar] [CrossRef] [PubMed]
  96. Bonakdar, E.; Edriss, M.A.; Bakhtari, A.; Jafarpour, F.; Asgari, V.; Hosseini, S.M.; Boroujeni, N.S.; Hajian, M.; Rahmani, H.R.; Nasr-Esfahani, M.H. A physiological, rather than a superovulated, post-implantation environment can attenuate the compromising effect of assisted reproductive techniques on gene expression in developing mice embryos. Mol. Reprod. Dev. 2015, 82, 191–206. [Google Scholar] [CrossRef]
  97. Saenz-de-Juano, M.D.; Billooye, K.; Smitz, J.; Anckaert, E. The loss of imprinted DNA methylation in mouse blastocysts is inflicted to a similar extent by in vitro follicle culture and ovulation induction. Mol. Hum. Reprod. 2016, 22, 427–441. [Google Scholar] [CrossRef]
  98. Ratajczak, M.Z. Igf2-H19, an imprinted tandem gene, is an important regulator of embryonic development, a guardian of proliferation of adult pluripotent stem cells, a regulator of longevity, and a ‘passkey’ to cancerogenesis. Folia Histochem. Cytobiol. 2012, 50, 171–179. [Google Scholar] [CrossRef]
  99. Ma, Y.; Ma, Y.; Wen, L.; Lei, H.; Chen, S.; Wang, X. Changes in DNA methylation and imprinting disorders in E9.5 mouse fetuses and placentas derived from vitrified eight-cell embryos. Mol. Reprod. Dev. 2019, 86, 404–415. [Google Scholar] [CrossRef]
  100. Bhattacharya, S. Maternal and perinatal outcomes after fresh versus frozen embryo transfer-what is the risk-benefit ratio? Fertil. Steril. 2016, 106, 241–243. [Google Scholar] [CrossRef]
  101. Toraño, E.G.; García, M.G.; Fernández-Morera, J.L.; Niño-García, P.; Fernández, A.F. The Impact of External Factors on the Epigenome: In Utero and over Lifetime. Biomed. Res. Int. 2016, 2016, 2568635. [Google Scholar] [CrossRef]
  102. Park, S.Y.; Kim, E.Y.; Cui, X.S.; Tae, J.C.; Lee, W.D.; Kim, N.H.; Park, S.P.; Lim, J.H. Increase in DNA fragmentation and apoptosis-related gene expression in frozen-thawed bovine blastocysts. Zygote 2006, 14, 125–131. [Google Scholar] [CrossRef] [PubMed]
  103. Yao, J.; Geng, L.; Huang, R.; Peng, W.; Chen, X.; Jiang, X.; Yu, M.; Li, M.; Huang, Y.; Yang, X. Effect of vitrification on in vitro development and imprinted gene Grb10 in mouse embryos. Reproduction 2017, 154, 97–105. [Google Scholar] [CrossRef] [PubMed]
  104. Chen, H.; Zhang, L.; Deng, T.; Zou, P.; Wang, Y.; Quan, F.; Zhang, Y. Effects of oocyte vitrification on epigenetic status in early bovine embryos. Theriogenology 2016, 86, 868–878. [Google Scholar] [CrossRef]
  105. Derakhshan-Horeh, M.; Abolhassani, F.; Jafarpour, F.; Moini, A.; Karbalaie, K.; Hosseini, S.M.; Nasr-Esfahani, M.H. Vitrification at Day3 stage appears not to affect the methylation status of H19/IGF2 differentially methylated region of in vitro produced human blastocysts. Cryobiology 2016, 73, 168–174. [Google Scholar] [CrossRef]
  106. Hiura, H.; Hattori, H.; Kobayashi, N.; Okae, H.; Chiba, H.; Miyauchi, N.; Kitamura, A.; Kikuchi, H.; Yoshida, H.; Arima, T. Genome-wide microRNA expression profiling in placentae from frozen-thawed blastocyst transfer. Clin. Epigenetics 2017, 9, 79. [Google Scholar] [CrossRef]
  107. Velker, B.A.M.; Denomme, M.M.; Krafty, R.T.; Mann, M.R.W. Maintenance of Mest imprinted methylation in blastocyst-stage mouse embryos is less stable than other imprinted loci following superovulation or embryo culture. Environ. Epigenet 2017, 3, dvx015. [Google Scholar] [CrossRef] [PubMed]
  108. Market-Velker, B.A.; Fernandes, A.D.; Mann, M.R. 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, 938–950. [Google Scholar] [CrossRef]
  109. Fleming, T.P.; Kwong, W.Y.; Porter, R.; Ursell, E.; Fesenko, I.; Wilkins, A.; Miller, D.J.; Watkins, A.J.; Eckert, J.J. The embryo and its future. Biol. Reprod. 2004, 71, 1046–1054. [Google Scholar] [CrossRef] [PubMed]
  110. Lonergan, P.; Fair, T. The ART of studying early embryo development: Progress and challenges in ruminant embryo culture. Theriogenology 2014, 81, 49–55. [Google Scholar] [CrossRef]
  111. Lonergan, P.; Fair, T.; Corcoran, D.; Evans, A.C. Effect of culture environment on gene expression and developmental characteristics in IVF-derived embryos. Theriogenology 2006, 65, 137–152. [Google Scholar] [CrossRef] [PubMed]
  112. Lonergan, P.; Rizos, D.; Kanka, J.; Nemcova, L.; Mbaye, A.M.; Kingston, M.; Wade, M.; Duffy, P.; Boland, M.P. Temporal sensitivity of bovine embryos to culture environment after fertilization and the implications for blastocyst quality. Reproduction 2003, 126, 337–346. [Google Scholar] [CrossRef]
  113. Fauque, P.; Mondon, F.; Letourneur, F.; Ripoche, M.A.; Journot, L.; Barbaux, S.; Dandolo, L.; Patrat, C.; Wolf, J.P.; Jouannet, P.; et al. In vitro fertilization and embryo culture strongly impact the placental transcriptome in the mouse model. PLoS ONE 2010, 5, e9218. [Google Scholar] [CrossRef] [PubMed]
  114. Donjacour, A.; Liu, X.; Lin, W.; Simbulan, R.; Rinaudo, P.F. In Vitro Fertilization Affects Growth and Glucose Metabolism in a Sex-Specific Manner in an Outbred Mouse Model. Biol. Reprod. 2014, 90, 80. [Google Scholar] [CrossRef] [PubMed]
  115. Sui, L.; An, L.; Tan, K.; Wang, Z.; Wang, S.; Miao, K.; Ren, L.; Tao, L.; He, S.; Yu, Y.; et al. Dynamic proteomic profiles of in vivo- and in vitro-produced mouse postimplantation extraembryonic tissues and placentas. Biol. Reprod. 2014, 91, 155. [Google Scholar] [CrossRef]
  116. Crosier, A.E.; Farin, C.E.; Rodriguez, K.F.; Blondin, P.; Alexander, J.E.; Farin, P.W. Development of skeletal muscle and expression of candidate genes in bovine fetuses from embryos produced in vivo or in vitro. Biol. Reprod. 2002, 67, 401–408. [Google Scholar] [CrossRef]
  117. Barros, C.M.; Satrapa, R.A.; Castilho, A.C.; Fontes, P.K.; Razza, E.M.; Ereno, R.L.; Nogueira, M.F. 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. [Google Scholar] [CrossRef] [PubMed]
  118. Kerjean, A.; Couvert, P.; Heams, T.; Chalas, C.; Poirier, K.; Chelly, J.; Jouannet, P.; Paldi, A.; Poirot, C. In vitro follicular growth affects oocyte imprinting establishment in mice. Eur. J. Hum. Genet. 2003, 11, 493–496. [Google Scholar] [CrossRef] [PubMed]
  119. Gioia, L.; Barboni, B.; Turriani, M.; Capacchietti, G.; Pistilli, M.G.; Berardinelli, P.; Mattioli, M. The capability of reprogramming the male chromatin after fertilization is dependent on the quality of oocyte maturation. Reproduction 2005, 130, 29–39. [Google Scholar] [CrossRef] [PubMed]
  120. Borghol, N.; Lornage, J.; Blachère, T.; Sophie Garret, A.; Lefèvre, A. Epigenetic status of the H19 locus in human oocytes following in vitro maturation. Genomics 2006, 87, 417–426. [Google Scholar] [CrossRef] [PubMed]
  121. Kuhtz, J.; Romero, S.; De Vos, M.; Smitz, J.; Haaf, T.; Anckaert, E. Human in vitro oocyte maturation is not associated with increased imprinting error rates at LIT1, SNRPN, PEG3 and GTL2. Hum. Reprod. 2014, 29, 1995–2005. [Google Scholar] [CrossRef] [PubMed]
  122. Pliushch, G.; Schneider, E.; Schneider, T.; El Hajj, N.; Rösner, S.; Strowitzki, T.; Haaf, T. In vitro maturation of oocytes is not associated with altered deoxyribonucleic acid methylation patterns in children from in vitro fertilization or intracytoplasmic sperm injection. Fertil. Steril. 2015, 103, 720–727. [Google Scholar] [CrossRef]
  123. Hewitson, L.; Takahashi, D.; Dominko, T.; Simerly, C.; Schatten, G. Fertilization and embryo development to blastocysts after intracytoplasmic sperm injection in the rhesus monkey. Hum. Reprod. 1998, 13, 3449–3455. [Google Scholar] [CrossRef]
  124. Dozortsev, D.; Wakaiama, T.; Ermilov, A.; Yanagimachi, R. Intracytoplasmic sperm injection in the rat. Zygote 1998, 6, 143–147. [Google Scholar] [CrossRef] [PubMed]
  125. Ferré, L.B.; Alvarez-Gallardo, H.; Romo, S.; Fresno, C.; Stroud, T.; Stroud, B.; Lindsey, B.; Kjelland, M.E. Transvaginal ultrasound-guided oocyte retrieval in cattle: State-of-the-art and its impact on the in vitro fertilization embryo production outcome. Reprod. Domest. Anim. 2023, 58, 363–378. [Google Scholar] [CrossRef] [PubMed]
  126. Kohda, T.; Ogonuki, N.; Inoue, K.; Furuse, T.; Kaneda, H.; Suzuki, T.; Kaneko-Ishino, T.; Wakayama, T.; Wakana, S.; Ogura, A.; et al. Intracytoplasmic sperm injection induces transcriptome perturbation without any transgenerational effect. Biochem. Biophys. Res. Commun. 2011, 410, 282–288. [Google Scholar] [CrossRef] [PubMed]
  127. Bridges, P.J.; Jeoung, M.; Kim, H.; Kim, J.H.; Lee, D.R.; Ko, C.; Baker, D.J. Methodology matters: IVF versus ICSI and embryonic gene expression. Reprod. Biomed. Online 2011, 23, 234–244. [Google Scholar] [CrossRef]
  128. Wildman, D.E. IFPA award in placentology lecture: Phylogenomic origins and evolution of the mammalian placenta. Placenta 2016, 48 (Suppl. 1), S31–S39. [Google Scholar] [CrossRef]
  129. Okae, H.; Hiura, H.; Nishida, Y.; Funayama, R.; Tanaka, S.; Chiba, H.; Yaegashi, N.; Nakayama, K.; Sasaki, H.; Arima, T. Re-investigation and RNA sequencing-based identification of genes with placenta-specific imprinted expression. Hum. Mol. Genet. 2012, 21, 548–558. [Google Scholar] [CrossRef]
  130. Nelissen, E.C.; van Montfoort, A.P.; Dumoulin, J.C.; Evers, J.L. Epigenetics and the placenta. Hum. Reprod. Update 2011, 17, 397–417. [Google Scholar] [CrossRef] [PubMed]
  131. Sakian, S.; Louie, K.; Wong, E.C.; Havelock, J.; Kashyap, S.; Rowe, T.; Taylor, B.; Ma, S. Altered gene expression of H19 and IGF2 in placentas from ART pregnancies. Placenta 2015, 36, 1100–1105. [Google Scholar] [CrossRef]
  132. Katari, S.; Turan, N.; Bibikova, M.; Erinle, O.; Chalian, R.; Foster, M.; Gaughan, J.P.; Coutifaris, C.; Sapienza, C. DNA methylation and gene expression differences in children conceived in vitro or in vivo. Hum. Mol. Genet. 2009, 18, 3769–3778. [Google Scholar] [CrossRef]
  133. Shi, X.; Ni, Y.; Zheng, H.; Chen, S.; Zhong, M.; Wu, F.; Xia, R.; Luo, Y. Abnormal methylation patterns at the IGF2/H19 imprinting control region in phenotypically normal babies conceived by assisted reproductive technologies. Eur. J. Obstet. Gynecol. Reprod. Biol. 2011, 158, 52–55. [Google Scholar] [CrossRef] [PubMed]
  134. Nelissen, E.C.; Dumoulin, J.C.; Daunay, A.; Evers, J.L.; Tost, J.; van Montfoort, A.P. Placentas from pregnancies conceived by IVF/ICSI have a reduced DNA methylation level at the H19 and MEST differentially methylated regions. Hum. Reprod. 2013, 28, 1117–1126. [Google Scholar] [CrossRef]
  135. Hanna, C.W.; Demond, H.; Kelsey, G. Epigenetic regulation in development: Is the mouse a good model for the human? Hum. Reprod. Update 2018, 24, 556–576. [Google Scholar] [CrossRef] [PubMed]
  136. Kobayashi, H.; Sato, A.; Otsu, E.; Hiura, H.; Tomatsu, C.; Utsunomiya, T.; Sasaki, H.; Yaegashi, N.; Arima, T. Aberrant DNA methylation of imprinted loci in sperm from oligospermic patients. Hum. Mol. Genet. 2007, 16, 2542–2551. [Google Scholar] [CrossRef]
  137. Young, L.E.; Fernandes, K.; McEvoy, T.G.; Butterwith, S.C.; Gutierrez, C.G.; Carolan, C.; Broadbent, P.J.; Robinson, J.J.; Wilmut, I.; Sinclair, K.D. Epigenetic change in IGF2R is associated with fetal overgrowth after sheep embryo culture. Nat. Genet. 2001, 27, 153–154. [Google Scholar] [CrossRef] [PubMed]
  138. de Waal, E.; Vrooman, L.A.; Fischer, E.; Ord, T.; Mainigi, M.A.; Coutifaris, C.; Schultz, R.M.; Bartolomei, M.S. The cumulative effect of assisted reproduction procedures on placental development and epigenetic perturbations in a mouse model. Hum. Mol. Genet. 2015, 24, 6975–6985. [Google Scholar] [CrossRef] [PubMed]
  139. de Waal, E.; Mak, W.; Calhoun, S.; Stein, P.; Ord, T.; Krapp, C.; Coutifaris, C.; Schultz, R.M.; Bartolomei, M.S. In vitro culture increases the frequency of stochastic epigenetic errors at imprinted genes in placental tissues from mouse concepti produced through assisted reproductive technologies. Biol. Reprod. 2014, 90, 22. [Google Scholar] [CrossRef]
  140. Wisborg, K.; Ingerslev, H.J.; Henriksen, T.B. IVF and stillbirth: A prospective follow-up study. Hum. Reprod. 2010, 25, 1312–1316. [Google Scholar] [CrossRef]
  141. Hart, R.; Norman, R.J. The longer-term health outcomes for children born as a result of IVF treatment: Part I--General health outcomes. Hum. Reprod. Update 2013, 19, 232–243. [Google Scholar] [CrossRef]
  142. Pinborg, A.; Wennerholm, U.B.; Romundstad, L.B.; Loft, A.; Aittomaki, K.; Söderström-Anttila, V.; Nygren, K.G.; Hazekamp, J.; Bergh, C. Why do singletons conceived after assisted reproduction technology have adverse perinatal outcome? Systematic review and meta-analysis. Hum. Reprod. Update 2013, 19, 87–104. [Google Scholar] [CrossRef]
  143. Henningsen, A.K.; Pinborg, A.; Lidegaard, Ø.; Vestergaard, C.; Forman, J.L.; Andersen, A.N. Perinatal outcome of singleton siblings born after assisted reproductive technology and spontaneous conception: Danish national sibling-cohort study. Fertil. Steril. 2011, 95, 959–963. [Google Scholar] [CrossRef] [PubMed]
  144. Pandey, S.; Shetty, A.; Hamilton, M.; Bhattacharya, S.; Maheshwari, A. Obstetric and perinatal outcomes in singleton pregnancies resulting from IVF/ICSI: A systematic review and meta-analysis. Hum. Reprod. Update 2012, 18, 485–503. [Google Scholar] [CrossRef]
  145. Helmerhorst, F.M.; Perquin, D.A.; Donker, D.; Keirse, M.J. Perinatal outcome of singletons and twins after assisted conception: A systematic review of controlled studies. BMJ 2004, 328, 261. [Google Scholar] [CrossRef]
  146. Davies, M.J.; Moore, V.M.; Willson, K.J.; Van Essen, P.; Priest, K.; Scott, H.; Haan, E.A.; Chan, A. Reproductive technologies and the risk of birth defects. N. Engl. J. Med. 2012, 366, 1803–1813. [Google Scholar] [CrossRef]
  147. Buckett, W.M.; Chian, R.C.; Holzer, H.; Dean, N.; Usher, R.; Tan, S.L. Obstetric outcomes and congenital abnormalities after in vitro maturation, in vitro fertilization, and intracytoplasmic sperm injection. Obstet. Gynecol. 2007, 110, 885–891. [Google Scholar] [CrossRef]
  148. Seggers, J.; Pontesilli, M.; Ravelli, A.C.J.; Painter, R.C.; Hadders-Algra, M.; Heineman, M.J.; Repping, S.; Mol, B.W.J.; Roseboom, T.J.; Ensing, S. Effects of in vitro fertilization and maternal characteristics on perinatal outcomes: A population-based study using siblings. Fertil. Steril. 2016, 105, 590–598.e2. [Google Scholar] [CrossRef]
  149. Horton, J.; Sterrenburg, M.; Lane, S.; Maheshwari, A.; Li, T.C.; Cheong, Y. Reproductive, obstetric, and perinatal outcomes of women with adenomyosis and endometriosis: A systematic review and meta-analysis. Hum. Reprod. Update 2019, 25, 592–632. [Google Scholar] [CrossRef] [PubMed]
  150. Sterling, L.; Liu, J.; Okun, N.; Sakhuja, A.; Sierra, S.; Greenblatt, E. Pregnancy outcomes in women with polycystic ovary syndrome undergoing in vitro fertilization. Fertil. Steril. 2016, 105, 791–797.e2. [Google Scholar] [CrossRef] [PubMed]
  151. Isaksson, R.; Gissler, M.; Tiitinen, A. Obstetric outcome among women with unexplained infertility after IVF: A matched case-control study. Hum. Reprod. 2002, 17, 1755–1761. [Google Scholar] [CrossRef] [PubMed]
  152. Luke, B.; Gopal, D.; Cabral, H.; Stern, J.E.; Diop, H. Pregnancy, birth, and infant outcomes by maternal fertility status: The Massachusetts Outcomes Study of Assisted Reproductive Technology. Am. J. Obstet. Gynecol. 2017, 217, 327.e1–327.e14. [Google Scholar] [CrossRef] [PubMed]
  153. Yu, H.; Liang, Z.; Cai, R.; Jin, S.; Xia, T.; Wang, C.; Kuang, Y. Association of adverse birth outcomes with in vitro fertilization after controlling infertility factors based on a singleton live birth cohort. Sci. Rep. 2022, 12, 4528. [Google Scholar] [CrossRef] [PubMed]
  154. Schieve, L.A.; Meikle, S.F.; Ferre, C.; Peterson, H.B.; Jeng, G.; Wilcox, L.S. Low and very low birth weight in infants conceived with use of assisted reproductive technology. N. Engl. J. Med. 2002, 346, 731–737. [Google Scholar] [CrossRef] [PubMed]
  155. Żyła, M.M.; Wilczyński, J.; Nowakowska-Głąb, A.; Maniecka-Bryła, I.; Nowakowska, D. Pregnancy and Delivery in Women with Uterine Malformations. Adv. Clin. Exp. Med. 2015, 24, 873–879. [Google Scholar] [CrossRef]
  156. Boyle, A.K.; Rinaldi, S.F.; Norman, J.E.; Stock, S.J. Preterm birth: Inflammation, fetal injury and treatment strategies. J. Reprod. Immunol. 2017, 119, 62–66. [Google Scholar] [CrossRef] [PubMed]
  157. Sunkara, S.K.; Antonisamy, B.; Redla, A.C.; Kamath, M.S. Female causes of infertility are associated with higher risk of preterm birth and low birth weight: Analysis of 117 401 singleton live births following IVF. Hum. Reprod. 2021, 36, 676–682. [Google Scholar] [CrossRef]
  158. Dunietz, G.L.; Holzman, C.; Zhang, Y.; Li, C.; Todem, D.; Boulet, S.L.; McKane, P.; Kissin, D.M.; Copeland, G.; Bernson, D.; et al. Assisted reproduction and risk of preterm birth in singletons by infertility diagnoses and treatment modalities: A population-based study. J. Assist. Reprod. Genet. 2017, 34, 1529–1535. [Google Scholar] [CrossRef]
  159. Kawwass, J.F.; Crawford, S.; Kissin, D.M.; Session, D.R.; Boulet, S.; Jamieson, D.J. Tubal factor infertility and perinatal risk after assisted reproductive technology. Obstet. Gynecol. 2013, 121, 1263–1271. [Google Scholar] [CrossRef] [PubMed]
  160. Mariappen, U.; Keane, K.N.; Hinchliffe, P.M.; Dhaliwal, S.S.; Yovich, J.L. Neither male age nor semen parameters influence clinical pregnancy or live birth outcomes from IVF. Reprod. Biol. 2018, 18, 324–329. [Google Scholar] [CrossRef]
  161. Pinborg, A.; Loft, A.; Aaris Henningsen, A.K.; Rasmussen, S.; Andersen, A.N. Infant outcome of 957 singletons born after frozen embryo replacement: The Danish National Cohort Study 1995–2006. Fertil Steril 2010, 94, 1320–1327. [Google Scholar] [CrossRef]
  162. Mohseni, R.; Mohammed, S.H.; Safabakhsh, M.; Mohseni, F.; Monfared, Z.S.; Seyyedi, J.; Mejareh, Z.N.; Alizadeh, S. Birth Weight and Risk of Cardiovascular Disease Incidence in Adulthood: A Dose-Response Meta-analysis. Curr. Atheroscler. Rep. 2020, 22, 12. [Google Scholar] [CrossRef]
  163. Skilton, M.R.; Siitonen, N.; Würtz, P.; Viikari, J.S.; Juonala, M.; Seppälä, I.; Laitinen, T.; Lehtimäki, T.; Taittonen, L.; Kähönen, M.; et al. High birth weight is associated with obesity and increased carotid wall thickness in young adults: The cardiovascular risk in young Finns study. Arterioscler. Thromb. Vasc. Biol. 2014, 34, 1064–1068. [Google Scholar] [CrossRef] [PubMed]
  164. Hansen, M.; Kurinczuk, J.J.; Bower, C.; Webb, S. The risk of major birth defects after intracytoplasmic sperm injection and in vitro fertilization. N. Engl. J. Med. 2002, 346, 725–730. [Google Scholar] [CrossRef]
  165. Klemetti, R.; Gissler, M.; Sevón, T.; Koivurova, S.; Ritvanen, A.; Hemminki, E. Children born after assisted fertilization have an increased rate of major congenital anomalies. Fertil. Steril. 2005, 84, 1300–1307. [Google Scholar] [CrossRef] [PubMed]
  166. Smithers, P.R.; Halliday, J.; Hale, L.; Talbot, J.M.; Breheny, S.; Healy, D. High frequency of cesarean section, antepartum hemorrhage, placenta previa, and preterm delivery in in-vitro fertilization twin pregnancies. Fertil. Steril. 2003, 80, 666–668. [Google Scholar] [CrossRef] [PubMed]
  167. Manoura, A.; Korakaki, E.; Hatzidaki, E.; Bikouvarakis, S.; Papageorgiou, M.; Giannakopoulou, C. Perinatal outcome of twin pregnancies after in vitro fertilization. Acta Obstet. Gynecol. Scand. 2004, 83, 1079–1084. [Google Scholar] [CrossRef]
  168. Bosch, B.; Sutcliffe, A. Congenital Anomalies Following Assisted Reproductive Technology. In Complications and Outcomes of Assisted Reproduction; Rizk, B., Gerris, J., Eds.; Cambridge University Press: Cambridge, UK, 2017; pp. 15–23. [Google Scholar] [CrossRef]
  169. Katalinic, A.; Rösch, C.; Ludwig, M. Pregnancy course and outcome after intracytoplasmic sperm injection: A controlled, prospective cohort study. Fertil. Steril. 2004, 81, 1604–1616. [Google Scholar] [CrossRef]
  170. El Hajj, N.; Haaf, T. Epigenetic disturbances in in vitro cultured gametes and embryos: Implications for human assisted reproduction. Fertil. Steril. 2013, 99, 632–641. [Google Scholar] [CrossRef] [PubMed]
  171. Wang, J.X.; Norman, R.J.; Kristiansson, P. The effect of various infertility treatments on the risk of preterm birth. Hum. Reprod. 2002, 17, 945–949. [Google Scholar] [CrossRef]
  172. Luke, B.; Brown, M.B.; Wantman, E.; Forestieri, N.E.; Browne, M.L.; Fisher, S.C.; Yazdy, M.M.; Ethen, M.K.; Canfield, M.A.; Watkins, S.; et al. The risk of birth defects with conception by ART. Hum. Reprod. 2021, 36, 116–129. [Google Scholar] [CrossRef]
  173. Hansen, M.; Kurinczuk, J.J.; Milne, E.; de Klerk, N.; Bower, C. Assisted reproductive technology and birth defects: A systematic review and meta-analysis. Hum. Reprod. Update 2013, 19, 330–353. [Google Scholar] [CrossRef] [PubMed]
  174. Olson, C.K.; Keppler-Noreuil, K.M.; Romitti, P.A.; Budelier, W.T.; Ryan, G.; Sparks, A.E.; Van Voorhis, B.J. In vitro fertilization is associated with an increase in major birth defects. Fertil. Steril. 2005, 84, 1308–1315. [Google Scholar] [CrossRef]
  175. Kanyó, K.; Konc, J. A follow-up study of children born after diode laser assisted hatching. Eur. J. Obstet. Gynecol. Reprod. Biol. 2003, 110, 176–180. [Google Scholar] [CrossRef]
  176. Nassar, A.H.; Usta, I.M.; Rechdan, J.B.; Harb, T.S.; Adra, A.M.; Abu-Musa, A.A. Pregnancy outcome in spontaneous twins versus twins who were conceived through in vitro fertilization. Am. J. Obstet. Gynecol. 2003, 189, 513–518. [Google Scholar] [CrossRef]
  177. Hansen, M.; Kurinczuk, J.J.; de Klerk, N.; Burton, P.; Bower, C. Assisted reproductive technology and major birth defects in Western Australia. Obstet. Gynecol. 2012, 120, 852–863. [Google Scholar] [CrossRef]
  178. Zhu, J.L.; Basso, O.; Obel, C.; Bille, C.; Olsen, J. Infertility, infertility treatment, and congenital malformations: Danish national birth cohort. BMJ 2006, 333, 679. [Google Scholar] [CrossRef]
  179. La Rovere, M.; Franzago, M.; Stuppia, L. Epigenetics and Neurological Disorders in ART. Int. J. Mol. Sci. 2019, 20, 4169. [Google Scholar] [CrossRef] [PubMed]
  180. Wu, Y.; Lv, Z.; Yang, Y.; Dong, G.; Yu, Y.; Cui, Y.; Tong, M.; Wang, L.; Zhou, Z.; Zhu, H.; 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, 1761–1774. [Google Scholar] [CrossRef]
  181. Strömberg, B.; Dahlquist, G.; Ericson, A.; Finnström, O.; Köster, M.; Stjernqvist, K. Neurological sequelae in children born after in-vitro fertilisation: A population-based study. Lancet 2002, 359, 461–465. [Google Scholar] [CrossRef] [PubMed]
  182. Lidegaard, O.; Pinborg, A.; Andersen, A.N. Imprinting diseases and IVF: Danish National IVF cohort study. Hum. Reprod. 2005, 20, 950–954. [Google Scholar] [CrossRef] [PubMed]
  183. Källén, A.J.; Finnström, O.O.; Lindam, A.P.; Nilsson, E.M.; Nygren, K.G.; Olausson, P.M. Cerebral palsy in children born after in vitro fertilization. Is the risk decreasing? Eur. J. Paediatr. Neurol. 2010, 14, 526–530. [Google Scholar] [CrossRef]
  184. Hvidtjørn, D.; Grove, J.; Schendel, D.; Svaerke, C.; Schieve, L.A.; Uldall, P.; Ernst, E.; Jacobsson, B.; Thorsen, P. Multiplicity and early gestational age contribute to an increased risk of cerebral palsy from assisted conception: A population-based cohort study. Hum. Reprod. 2010, 25, 2115–2123. [Google Scholar] [CrossRef] [PubMed]
  185. Zhu, J.L.; Hvidtjørn, D.; Basso, O.; Obel, C.; Thorsen, P.; Uldall, P.; Olsen, J. Parental infertility and cerebral palsy in children. Hum. Reprod. 2010, 25, 3142–3145. [Google Scholar] [CrossRef]
  186. Bay, B.; Mortensen, E.L.; Kesmodel, U.S. Assisted reproduction and child neurodevelopmental outcomes: A systematic review. Fertil. Steril. 2013, 100, 844–853. [Google Scholar] [CrossRef] [PubMed]
  187. Abdel-Mannan, O.; Sutcliffe, A. I was born following ART: How will I get on at school? Semin. Fetal Neonatal Med. 2014, 19, 245–249. [Google Scholar] [CrossRef]
  188. Punamäki, R.L.; Tiitinen, A.; Lindblom, J.; Unkila-Kallio, L.; Flykt, M.; Vänskä, M.; Poikkeus, P.; Tulppala, M. Mental health and developmental outcomes for children born after ART: A comparative prospective study on child gender and treatment type. Hum. Reprod. 2016, 31, 100–107. [Google Scholar] [CrossRef]
  189. Rumbold, A.R.; Moore, V.M.; Whitrow, M.J.; Oswald, T.K.; Moran, L.J.; Fernandez, R.C.; Barnhart, K.T.; Davies, M.J. The impact of specific fertility treatments on cognitive development in childhood and adolescence: A systematic review. Hum. Reprod. 2017, 32, 1489–1507. [Google Scholar] [CrossRef]
  190. Maimburg, R.D.; Vaeth, M. Do children born after assisted conception have less risk of developing infantile autism? Hum. Reprod. 2007, 22, 1841–1843. [Google Scholar] [CrossRef]
  191. Hvidtjørn, D.; Grove, J.; Schendel, D.; Schieve, L.A.; Sværke, C.; Ernst, E.; Thorsen, P. Risk of autism spectrum disorders in children born after assisted conception: A population-based follow-up study. J. Epidemiol. Community Health 2011, 65, 497–502. [Google Scholar] [CrossRef] [PubMed]
  192. Kissin, D.M.; Zhang, Y.; Boulet, S.L.; Fountain, C.; Bearman, P.; Schieve, L.; Yeargin-Allsopp, M.; Jamieson, D.J. Association of assisted reproductive technology (ART) treatment and parental infertility diagnosis with autism in ART-conceived children. Hum. Reprod. 2015, 30, 454–465. [Google Scholar] [CrossRef] [PubMed]
  193. Liu, L.; Gao, J.; He, X.; Cai, Y.; Wang, L.; Fan, X. Association between assisted reproductive technology and the risk of autism spectrum disorders in the offspring: A meta-analysis. Sci. Rep. 2017, 7, 46207. [Google Scholar] [CrossRef]
  194. Hargreave, M.; Jensen, A.; Toender, A.; Andersen, K.K.; Kjaer, S.K. Fertility treatment and childhood cancer risk: A systematic meta-analysis. Fertil. Steril. 2013, 100, 150–161. [Google Scholar] [CrossRef] [PubMed]
  195. Chiavarini, M.; Ostorero, A.; Naldini, G.; Fabiani, R. Cancer Risk in Children and Young Adults (Offspring) Born after Medically Assisted Reproduction: A Systematic Review and Meta-Analysis. J 2019, 2, 430–448. [Google Scholar] [CrossRef]
  196. Wang, T.; Chen, L.; Yang, T.; Wang, L.; Zhao, L.; Zhang, S.; Ye, Z.; Chen, L.; Zheng, Z.; Qin, J. Cancer risk among children conceived by fertility treatment. Int. J. Cancer 2019, 144, 3001–3013. [Google Scholar] [CrossRef]
  197. Williams, C.L.; Bunch, K.J.; Murphy, M.F.G.; Stiller, C.A.; Botting, B.J.; Wallace, W.H.; Davies, M.C.; Sutcliffe, A.G. Cancer risk in children born after donor ART. Hum. Reprod. 2018, 33, 140–146. [Google Scholar] [CrossRef]
  198. Sundh, K.J.; Henningsen, A.K.; Källen, K.; Bergh, C.; Romundstad, L.B.; Gissler, M.; Pinborg, A.; Skjaerven, R.; Tiitinen, A.; Vassard, D.; et al. Cancer in children and young adults born after assisted reproductive technology: A Nordic cohort study from the Committee of Nordic ART and Safety (CoNARTaS). Hum. Reprod. 2014, 29, 2050–2057. [Google Scholar] [CrossRef]
  199. Lerner-Geva, L.; Boyko, V.; Ehrlich, S.; Mashiach, S.; Hourvitz, A.; Haas, J.; Margalioth, E.; Levran, D.; Calderon, I.; Orvieto, R.; et al. Possible risk for cancer among children born following assisted reproductive technology in Israel. Pediatr. Blood Cancer 2017, 64, e26292. [Google Scholar] [CrossRef] [PubMed]
  200. Spaan, M.; van den Belt-Dusebout, A.W.; van den Heuvel-Eibrink, M.M.; Hauptmann, M.; Lambalk, C.B.; Burger, C.W.; van Leeuwen, F.E. Risk of cancer in children and young adults conceived by assisted reproductive technology. Hum. Reprod. 2019, 34, 740–750. [Google Scholar] [CrossRef]
  201. Källén, B.; Finnström, O.; Lindam, A.; Nilsson, E.; Nygren, K.G.; Olausson, P.O. Cancer risk in children and young adults conceived by in vitro fertilization. Pediatrics 2010, 126, 270–276. [Google Scholar] [CrossRef]
  202. Barker, D.J.; Osmond, C.; Golding, J.; Kuh, D.; Wadsworth, M.E. Growth in utero, blood pressure in childhood and adult life, and mortality from cardiovascular disease. BMJ 1989, 298, 564–567. [Google Scholar] [CrossRef]
  203. Barker, D.J. The origins of the developmental origins theory. J. Intern. Med. 2007, 261, 412–417. [Google Scholar] [CrossRef]
  204. Morton, J.S.; Cooke, C.L.; Davidge, S.T. In Utero Origins of Hypertension: Mechanisms and Targets for Therapy. Physiol. Rev. 2016, 96, 549–603. [Google Scholar] [CrossRef] [PubMed]
  205. Schenewerk, A.L.; Ramírez, F.; Foote, C.; Ji, T.; Martínez-Lemus, L.A.; Rivera, R.M. Effects of the use of assisted reproduction and high-caloric diet consumption on body weight and cardiovascular health of juvenile mouse offspring. Reproduction 2014, 147, 111–123. [Google Scholar] [CrossRef]
  206. Rexhaj, E.; Pireva, A.; Paoloni-Giacobino, A.; Allemann, Y.; Cerny, D.; Dessen, P.; Sartori, C.; Scherrer, U.; Rimoldi, S.F. Prevention of vascular dysfunction and arterial hypertension in mice generated by assisted reproductive technologies by addition of melatonin to culture media. Am. J. Physiol. Heart Circ. Physiol. 2015, 309, H1151–H1156. [Google Scholar] [CrossRef]
  207. Wang, L.Y.; Le, F.; Wang, N.; Li, L.; Liu, X.Z.; Zheng, Y.M.; Lou, H.Y.; Xu, X.R.; Chen, Y.L.; Zhu, X.M.; et al. Alteration of fatty acid metabolism in the liver, adipose tissue, and testis of male mice conceived through assisted reproductive technologies: Fatty acid metabolism in ART mice. Lipids Health Dis. 2013, 12, 5. [Google Scholar] [CrossRef]
  208. Yeung, E.H.; Druschel, C. Cardiometabolic health of children conceived by assisted reproductive technologies. Fertil. Steril. 2013, 99, 318–326. [Google Scholar] [CrossRef]
  209. Scherrer, U.; Rexhaj, E.; Allemann, Y.; Sartori, C.; Rimoldi, S.F. Cardiovascular dysfunction in children conceived by assisted reproductive technologies. Eur. Heart J. 2015, 36, 1583–1589. [Google Scholar] [CrossRef]
  210. Wikstrand, M.H.; Niklasson, A.; Strömland, K.; Hellström, A. Abnormal vessel morphology in boys born after intracytoplasmic sperm injection. Acta Paediatr. 2008, 97, 1512–1517. [Google Scholar] [CrossRef] [PubMed]
  211. Gao, Q.; Pan, H.T.; Lin, X.H.; Zhang, J.Y.; Jiang, Y.; Tian, S.; Chen, L.T.; Liu, M.E.; Xiong, Y.M.; Huang, H.F.; et al. Altered protein expression profiles in umbilical veins: Insights into vascular dysfunctions of the children born after in vitro fertilization. Biol. Reprod. 2014, 91, 71. [Google Scholar] [CrossRef] [PubMed]
  212. de Jong, F.; Monuteaux, M.C.; van Elburg, R.M.; Gillman, M.W.; Belfort, M.B. Systematic review and meta-analysis of preterm birth and later systolic blood pressure. Hypertension 2012, 59, 226–234. [Google Scholar] [CrossRef]
  213. Kaijser, M.; Bonamy, A.K.; Akre, O.; Cnattingius, S.; Granath, F.; Norman, M.; Ekbom, A. Perinatal risk factors for ischemic heart disease: Disentangling the roles of birth weight and preterm birth. Circulation 2008, 117, 405–410. [Google Scholar] [CrossRef] [PubMed]
  214. Guo, X.Y.; Liu, X.M.; Jin, L.; Wang, T.T.; Ullah, K.; Sheng, J.Z.; Huang, H.F. Cardiovascular and metabolic profiles of offspring conceived by assisted reproductive technologies: A systematic review and meta-analysis. Fertil. Steril. 2017, 107, 622–631.e5. [Google Scholar] [CrossRef] [PubMed]
  215. Scherrer, U.; Rimoldi, S.F.; Rexhaj, E.; Stuber, T.; Duplain, H.; Garcin, S.; de Marchi, S.F.; Nicod, P.; Germond, M.; Allemann, Y.; et al. Systemic and pulmonary vascular dysfunction in children conceived by assisted reproductive technologies. Circulation 2012, 125, 1890–1896. [Google Scholar] [CrossRef] [PubMed]
  216. Chen, M.; Wu, L.; Zhao, J.; Wu, F.; Davies, M.J.; Wittert, G.A.; Norman, R.J.; Robker, R.L.; Heilbronn, L.K. Altered glucose metabolism in mouse and humans conceived by IVF. Diabetes 2014, 63, 3189–3198. [Google Scholar] [CrossRef]
  217. Ceelen, M.; van Weissenbruch, M.M.; Vermeiden, J.P.; van Leeuwen, F.E.; Delemarre-van de Waal, H.A. Cardiometabolic differences in children born after in vitro fertilization: Follow-up study. J. Clin. Endocrinol. Metab. 2008, 93, 1682–1688. [Google Scholar] [CrossRef]
  218. Belva, F.; Henriet, S.; Liebaers, I.; Van Steirteghem, A.; Celestin-Westreich, S.; Bonduelle, M. Medical outcome of 8-year-old singleton ICSI children (born >or=32 weeks’ gestation) and a spontaneously conceived comparison group. Hum. Reprod. 2007, 22, 506–515. [Google Scholar] [CrossRef]
  219. Belva, F.; Painter, R.; Bonduelle, M.; Roelants, M.; Devroey, P.; De Schepper, J. Are ICSI adolescents at risk for increased adiposity? Hum. Reprod. 2012, 27, 257–264. [Google Scholar] [CrossRef]
  220. Sakka, S.D.; Loutradis, D.; Kanaka-Gantenbein, C.; Margeli, A.; Papastamataki, M.; Papassotiriou, I.; Chrousos, G.P. Absence of insulin resistance and low-grade inflammation despite early metabolic syndrome manifestations in children born after in vitro fertilization. Fertil. Steril. 2010, 94, 1693–1699. [Google Scholar] [CrossRef]
  221. Heber, M.F.; Ptak, G.E. The effects of assisted reproduction technologies on metabolic health and disease†. Biol. Reprod. 2021, 104, 734–744. [Google Scholar] [CrossRef]
  222. Carson, C.; Sacker, A.; Kelly, Y.; Redshaw, M.; Kurinczuk, J.J.; Quigley, M.A. Asthma in children born after infertility treatment: Findings from the UK Millennium Cohort Study. Hum. Reprod. 2013, 28, 471–479. [Google Scholar] [CrossRef]
  223. Cui, L.; Zhou, W.; Xi, B.; Ma, J.; Hu, J.; Fang, M.; Hu, K.; Qin, Y.; You, L.; Cao, Y.; et al. Increased risk of metabolic dysfunction in children conceived by assisted reproductive technology. Diabetologia 2020, 63, 2150–2157. [Google Scholar] [CrossRef]
  224. Karimi, H.; Mahdavi, P.; Fakhari, S.; Faryabi, M.R.; Esmaeili, P.; Banafshi, O.; Mohammadi, E.; Fathi, F.; Mokarizadeh, A. Altered helper T cell-mediated immune responses in male mice conceived through in vitro fertilization. Reprod. Toxicol. 2017, 69, 196–203. [Google Scholar] [CrossRef]
  225. Xu, X.; Wu, H.; Bian, Y.; Cui, L.; Man, Y.; Wang, Z.; Zhang, X.; Zhang, C.; Geng, L. The altered immunological status of children conceived by assisted reproductive technology. Reprod. Biol. Endocrinol. 2021, 19, 171. [Google Scholar] [CrossRef]
  226. Zhang, Y.; Cui, Y.; Zhou, Z.; Sha, J.; Li, Y.; Liu, J. Altered global gene expressions of human placentae subjected to assisted reproductive technology treatments. Placenta 2010, 31, 251–258. [Google Scholar] [CrossRef]
  227. Paolino, M.; Koglgruber, R.; Cronin, S.J.F.; Uribesalgo, I.; Rauscher, E.; Harreiter, J.; Schuster, M.; Bancher-Todesca, D.; Pranjic, B.; Novatchkova, M.; et al. RANK links thymic regulatory T cells to fetal loss and gestational diabetes in pregnancy. Nature 2021, 589, 442–447. [Google Scholar] [CrossRef]
  228. Hargreave, M.; Jensen, A.; Hansen, M.K.; Dehlendorff, C.; Winther, J.F.; Schmiegelow, K.; Kjær, S.K. Association Between Fertility Treatment and Cancer Risk in Children. JAMA 2019, 322, 2203–2210. [Google Scholar] [CrossRef]
  229. Papi, A.; Blasi, F.; Canonica, G.W.; Morandi, L.; Richeldi, L.; Rossi, A. Treatment strategies for asthma: Reshaping the concept of asthma management. Allergy Asthma Clin. Immunol. 2020, 16, 75. [Google Scholar] [CrossRef]
  230. Pawankar, R. Allergic diseases and asthma: A global public health concern and a call to action. World Allergy Organ. J. 2014, 7, 12. [Google Scholar] [CrossRef]
  231. Harris, R.A.; Nagy-Szakal, D.; Kellermayer, R. Human metastable epiallele candidates link to common disorders. Epigenetics 2013, 8, 157–163. [Google Scholar] [CrossRef]
  232. Reese, S.E.; Xu, C.J.; den Dekker, H.T.; Lee, M.K.; Sikdar, S.; Ruiz-Arenas, C.; Merid, S.K.; Rezwan, F.I.; Page, C.M.; Ullemar, V.; et al. Epigenome-wide meta-analysis of DNA methylation and childhood asthma. J. Allergy Clin. Immunol. 2019, 143, 2062–2074. [Google Scholar] [CrossRef]
  233. Declercq, E.; Luke, B.; Belanoff, C.; Cabral, H.; Diop, H.; Gopal, D.; Hoang, L.; Kotelchuck, M.; Stern, J.E.; Hornstein, M.D. Perinatal outcomes associated with assisted reproductive technology: The Massachusetts Outcomes Study of Assisted Reproductive Technologies (MOSART). Fertil. Steril. 2015, 103, 888–895. [Google Scholar] [CrossRef]
  234. Huang, L.; Chen, Q.; Zhao, Y.; Wang, W.; Fang, F.; Bao, Y. Is elective cesarean section associated with a higher risk of asthma? A meta-analysis. J. Asthma 2015, 52, 16–25. [Google Scholar] [CrossRef]
  235. Hart, R.; Norman, R.J. The longer-term health outcomes for children born as a result of IVF treatment. Part II--Mental health and development outcomes. Hum. Reprod. Update 2013, 19, 244–250. [Google Scholar] [CrossRef]
  236. Kettner, L.O.; Henriksen, T.B.; Bay, B.; Ramlau-Hansen, C.H.; Kesmodel, U.S. Assisted reproductive technology and somatic morbidity in childhood: A systematic review. Fertil. Steril. 2015, 103, 707–719. [Google Scholar] [CrossRef]
  237. Ericson, A.; Nygren, K.G.; Olausson, P.O.; Källén, B. Hospital care utilization of infants born after IVF. Hum. Reprod. 2002, 17, 929–932. [Google Scholar] [CrossRef]
  238. Tsabouri, S.; Lavasidis, G.; Efstathiadou, A.; Papasavva, M.; Bellou, V.; Bergantini, H.; Priftis, K.; Ntzani, E.E. Association between childhood asthma and history of assisted reproduction techniques: A systematic review and meta-analysis. Eur. J. Pediatr. 2021, 180, 2007–2017. [Google Scholar] [CrossRef]
  239. Källén, B.; Finnström, O.; Nygren, K.G.; Otterblad Olausson, P. Asthma in Swedish children conceived by in vitro fertilisation. Arch. Dis. Child. 2013, 98, 92–96. [Google Scholar] [CrossRef]
  240. García-Blanco, A.; Diago, V.; Hervás, D.; Ghosn, F.; Vento, M.; Cháfer-Pericás, C. Anxiety and depressive symptoms, and stress biomarkers in pregnant women after in vitro fertilization: A prospective cohort study. Hum. Reprod. 2018, 33, 1237–1246. [Google Scholar] [CrossRef]
  241. Wijs, L.A.; Doherty, D.A.; Keelan, J.A.; Penova-Veselinovic, B.; Burton, P.; Yovich, J.L.; Hall, G.L.; Sly, P.D.; Holt, P.G.; Hart, R.J. Asthma and allergies in a cohort of adolescents conceived with ART. Reprod. Biomed. Online 2022, 45, 1255–1265. [Google Scholar] [CrossRef]
  242. Glover, V. Prenatal stress and its effects on the fetus and the child: Possible underlying biological mechanisms. Adv. Neurobiol. 2015, 10, 269–283. [Google Scholar] [CrossRef]
  243. Ahmadi, H.; Fathi, F.; Moeini, A.; Amidi, F.; Sobhani, A. Evaluation of prooxidant-antioxidant balance in in vitro fertilization-conceived mice. Clin. Exp. Reprod. Med. 2018, 45, 82–87. [Google Scholar] [CrossRef] [PubMed]
  244. Zijlmans, M.A.; Korpela, K.; Riksen-Walraven, J.M.; de Vos, W.M.; de Weerth, C. Maternal prenatal stress is associated with the infant intestinal microbiota. Psychoneuroendocrinology 2015, 53, 233–245. [Google Scholar] [CrossRef]
  245. Moustaki, M.; Tsabouri, S.; Priftis, K.N.; Douros, K. Prenatal Stress Enhances Susceptibility to Allergic Diseases of Offspring. Endocr. Metab. Immune Disord. Drug Targets 2017, 17, 255–263. [Google Scholar] [CrossRef]
  246. Price, T.M.; Murphy, S.K.; Younglai, E.V. Perspectives: The possible influence of assisted reproductive technologies on transgenerational reproductive effects of environmental endocrine disruptors. Toxicol. Sci. 2007, 96, 218–226. [Google Scholar] [CrossRef]
  247. Baart, E.B.; Martini, E.; Eijkemans, M.J.; Van Opstal, D.; Beckers, N.G.; Verhoeff, A.; Macklon, N.S.; Fauser, B.C. Milder ovarian stimulation for in-vitro fertilization reduces aneuploidy in the human preimplantation embryo: A randomized controlled trial. Hum. Reprod. 2007, 22, 980–988. [Google Scholar] [CrossRef]
  248. Lainas, T.G.; Sfontouris, I.A.; Zorzovilis, I.Z.; Petsas, G.K.; Lainas, G.T.; Alexopoulou, E.; Kolibianakis, E.M. Flexible GnRH antagonist protocol versus GnRH agonist long protocol in patients with polycystic ovary syndrome treated for IVF: A prospective randomised controlled trial (RCT). Hum. Reprod. 2010, 25, 683–689. [Google Scholar] [CrossRef] [PubMed]
  249. Lee, S.J.; Schover, L.R.; Partridge, A.H.; Patrizio, P.; Wallace, W.H.; Hagerty, K.; Beck, L.N.; Brennan, L.V.; Oktay, K. American Society of Clinical Oncology recommendations on fertility preservation in cancer patients. J. Clin. Oncol. 2006, 24, 2917–2931. [Google Scholar] [CrossRef]
  250. Henningsen, A.A.; Gissler, M.; Skjaerven, R.; Bergh, C.; Tiitinen, A.; Romundstad, L.B.; Wennerholm, U.B.; Lidegaard, O.; Nyboe Andersen, A.; Forman, J.L.; et al. Trends in perinatal health after assisted reproduction: A Nordic study from the CoNARTaS group. Hum. Reprod. 2015, 30, 710–716. [Google Scholar] [CrossRef] [PubMed]
  251. Miller, L.M.; Hodgson, R.; Wong, T.Y.; Merrilees, M.; Norman, R.J.; Johnson, N.P. Single embryo transfer for all? Aust. N. Z. J. Obstet. Gynaecol. 2016, 56, 514–517. [Google Scholar] [CrossRef]
  252. Pandian, Z.; Marjoribanks, J.; Ozturk, O.; Serour, G.; Bhattacharya, S. Number of embryos for transfer following in vitro fertilisation or intra-cytoplasmic sperm injection. Cochrane Database Syst. Rev. 2013, 2013, Cd003416. [Google Scholar] [CrossRef]
  253. Gardner, D.K.; Lane, M.; Stevens, J.; Schoolcraft, W.B. Noninvasive assessment of human embryo nutrient consumption as a measure of developmental potential. Fertil. Steril. 2001, 76, 1175–1180. [Google Scholar] [CrossRef]
  254. Lane, M.; Gardner, D.K. Selection of viable mouse blastocysts prior to transfer using a metabolic criterion. Hum. Reprod. 1996, 11, 1975–1978. [Google Scholar] [CrossRef]
  255. Montskó, G.; Zrínyi, Z.; Janáky, T.; Szabó, Z.; Várnagy, Á.; Kovács, G.L.; Bódis, J. Noninvasive embryo viability assessment by quantitation of human haptoglobin alpha-1 fragment in the in vitro fertilization culture medium: An additional tool to increase success rate. Fertil. Steril. 2015, 103, 687–693. [Google Scholar] [CrossRef]
  256. Pallinger, E.; Bognar, Z.; Bodis, J.; Csabai, T.; Farkas, N.; Godony, K.; Varnagy, A.; Buzas, E.; Szekeres-Bartho, J. A simple and rapid flow cytometry-based assay to identify a competent embryo prior to embryo transfer. Sci. Rep. 2017, 7, 39927. [Google Scholar] [CrossRef]
  257. Feuerstein, P.; Cadoret, V.; Dalbies-Tran, R.; Guerif, F.; Bidault, R.; Royere, D. Gene expression in human cumulus cells: One approach to oocyte competence. Hum. Reprod. 2007, 22, 3069–3077. [Google Scholar] [CrossRef]
  258. Assou, S.; Haouzi, D.; Mahmoud, K.; Aouacheria, A.; Guillemin, Y.; Pantesco, V.; Rème, T.; Dechaud, H.; De Vos, J.; Hamamah, S. A non-invasive test for assessing embryo potential by gene expression profiles of human cumulus cells: A proof of concept study. Mol. Hum. Reprod. 2008, 14, 711–719. [Google Scholar] [CrossRef]
  259. Le Gac, S.; Nordhoff, V. Microfluidics for mammalian embryo culture and selection: Where do we stand now? Mol. Hum. Reprod. 2017, 23, 213–226. [Google Scholar] [CrossRef]
  260. Kushnir, V.A.; Smith, G.D.; Adashi, E.Y. The Future of IVF: The New Normal in Human Reproduction. Reprod. Sci. 2022, 29, 849–856. [Google Scholar] [CrossRef]
Figure 1. Types of Assisted Reproductive Technology. Common methods of assisted reproductive technology (ART) include: In vitro fertilization (IVF), Intrauterine insemination (IUI), and Intracytoplasmic sperm injection (ICSI). IVF and ICSI involve fertilization outside of the body, the latter with a single sperm injection into a mature egg. IUI involves direct sperm injection into the uterus. A receptive state of the endometrium is crucial to achieve implantation. Estrogen and progesterone are essential for establishing a receptive endometrium and successful pregnancy. These hormones are key factors for transition of the endometrium into a receptive state.
Figure 1. Types of Assisted Reproductive Technology. Common methods of assisted reproductive technology (ART) include: In vitro fertilization (IVF), Intrauterine insemination (IUI), and Intracytoplasmic sperm injection (ICSI). IVF and ICSI involve fertilization outside of the body, the latter with a single sperm injection into a mature egg. IUI involves direct sperm injection into the uterus. A receptive state of the endometrium is crucial to achieve implantation. Estrogen and progesterone are essential for establishing a receptive endometrium and successful pregnancy. These hormones are key factors for transition of the endometrium into a receptive state.
Ijms 24 13564 g001
Figure 2. Risks and complications associated with assisted reproductive technologies. ART procedures can induce epigenetic alterations (through three different mechanisms: DNA methylation of cytosine bases, chromatin modifications, and histone protein modifications) and obstetric complications such as placental abnormalities, multiple pregnancy and pre-term delivery. The possible long-term effects of ART include the increased risk of neurological, cardiovascular, metabolic, immunological and allergic diseases.
Figure 2. Risks and complications associated with assisted reproductive technologies. ART procedures can induce epigenetic alterations (through three different mechanisms: DNA methylation of cytosine bases, chromatin modifications, and histone protein modifications) and obstetric complications such as placental abnormalities, multiple pregnancy and pre-term delivery. The possible long-term effects of ART include the increased risk of neurological, cardiovascular, metabolic, immunological and allergic diseases.
Ijms 24 13564 g002
Table 1. Types and effects of preimplantation stress on embryo development and long-term effects.
Table 1. Types and effects of preimplantation stress on embryo development and long-term effects.
Types of StressEffects on Embryo Development and Long-Term EffectsDefinitionsReferences
In vitroIVF
  • Impaired development
  • Changes to gene expression
  • Altered ICM:TE number and ratio
  • Glucose intolerance
Adding a defined amount of sperm to the oocyte in a culture medium[40,51]
ICSI
  • Impaired development
  • Changes to gene expression
  • Altered ICM:TE number and ratio
A single spermatozoon is injected into the cytoplasm of a mature oocyte.[52,53]
Culture media composition
  • Impaired development
  • Changes to gene expression
  • Modified metabolism
  • Altered ICM:TE number and ratio
Variations in nutrient availability are sources of stresses.[54,55]
Oxygen tension
  • Impaired development
  • Changes to gene expression
  • Altered ICM:TE number and ratio
Oxygen concentration during culture and oxidative stress affect gene expression and intracellular signaling.[56]
Temperature
  • Changes to gene
  • Microtubule disassembly
Temperature fluctuations affect gamete and embryo viability.[57]
PH
  • Impaired development
  • Modified metabolism
  • Changes to gene expression
PH modulates metabolic activity, cellular proliferation, transcriptional activity, protein localization and synthesis.[36,58]
Light
  • Changes to gene expression
  • DNA fragmentation
  • Mitochondrial degradation
White light exposure of the embryo results in impaired implantation capacity.[45,49,59]
Substrate stiffness
  • Different developmental velocity
  • changes to gene expression
  • Altered ICM:TE number and ratio
fertilization and embryo development are more successful on a collagen matrix than on a standard polystyrene petri dish.[60]
Cryopreservation
  • Non-long-term and transgenerational effects
Cryoinjuries including ice crystal formation, structural damage to water bound enzymes, separation of membrane proteins from lipids, altered membrane permeability and osmotic stress due to changes in cell volume should be considered.[61,62,63]
In vivoSuboptimal diet
  • Reduced birthweight
  • Altered growth curves
  • Hypertension
Nutritional conditions in utero are associated with glucose intolerance, obesity, and cardiac dysfunction in adulthood.[64]
Maternal diseases
  • Altered ICM:TE number and ratio
Diabetes, hypertension, epilepsy, obesity, and cardiopulmonary disorders known as in vivo stress factors.[65,66]
Endocrine disruptors
  • Different developmental velocity
Poor oocyte maturation and competency, embryonic defects and poor IVF outcomes are possible complications of endocrine disorders.[67]
Toxins
  • Developmental delay
Tobacco, nicotine can cause delayed migration of embryos from the fallopian tubes into the uterus, growth retardation, pregnancy loss.[44,68]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ahmadi, H.; Aghebati-Maleki, L.; Rashidiani, S.; Csabai, T.; Nnaemeka, O.B.; Szekeres-Bartho, J. Long-Term Effects of ART on the Health of the Offspring. Int. J. Mol. Sci. 2023, 24, 13564. https://doi.org/10.3390/ijms241713564

AMA Style

Ahmadi H, Aghebati-Maleki L, Rashidiani S, Csabai T, Nnaemeka OB, Szekeres-Bartho J. Long-Term Effects of ART on the Health of the Offspring. International Journal of Molecular Sciences. 2023; 24(17):13564. https://doi.org/10.3390/ijms241713564

Chicago/Turabian Style

Ahmadi, Hamid, Leili Aghebati-Maleki, Shima Rashidiani, Timea Csabai, Obodo Basil Nnaemeka, and Julia Szekeres-Bartho. 2023. "Long-Term Effects of ART on the Health of the Offspring" International Journal of Molecular Sciences 24, no. 17: 13564. https://doi.org/10.3390/ijms241713564

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

Article Metrics

Back to TopTop