what epigenetics is and why it is thought to be important

van Otterdijk, S.D., and Michels, K.B. (2016). Transgenerational epigenetic inheritance in mammals: how good is the evidence? FASEB Journal, (30), p. 2457-2465. This is a review article that summarizes a number of research studies related to epigenetic inheritance and the state of the research in that area as of 2016. Please refer to the section on epigenetic inheritance in your textbook on page 217. This assignment is due no later than April 22nd, 2019 in Blackboard. You will not understand all of the terms or ideas covered in this article, and that’s ok. For the questions below, please refer to the appropriate sections of the article as indicated in order to answer each question. Please type your answers on a separate sheet of paper. Do not put your answers in the spaces below. 1. (20 points) Using information provided on pages 2457-2458, discuss:

a. what epigenetics is and why it is thought to be important b. what are the primary mechanisms involved in epigenetic inheritance c. what sorts of cellular changes may occur as a result of epigenetics

2. (20 points) Referring to page 2458, discuss how epigenetics might have evolutionary influences. 3. (20 points) Also on page 2458, there is a discussion of intrauterine exposures and how that might affect the fetus’s phenotype. What are some known exposures that affect the fetus in this way? 4. (20 points) Referring to page 2459, discuss how/why genomic imprinting may be the strongest evidence for transgenerational epigenetic inheritance. 5. (10 points) On pages 2462-2463, summarize the main points of the authors’ discussion and conclusions. Does transgenerational epigenetic inheritance exist? What might be some limitations of studying this in humans? 6. (10 points) Using the RMU Library database system, find a primary literature article that is related to this topic. Provide the full citation for this article as the answer to this question.

THE

JOURNAL • REVIEW • www.fasebj.org

Transgenerational epigenetic inheritance in mammals: how good is the evidence? Sanne D. van Otterdijk* and Karin B. Michels*,†,‡,1

*Institute for Prevention and Cancer Epidemiology, University Medical Center Freiburg, Freiburg, Germany; †Obstetrics and Gynecology Epidemiology Center, Department of Obstetrics, Gynecology and Reproductive Biology, Brigham and Women’s Hospital, Harvard Medical School, Boston, Massachusetts; and ‡Department of Epidemiology, Harvard School of Public Health, Boston, Massachusetts

ABSTRACT: Epigenetics plays an important role in orchestrating key biologic processes. Epigenetic marks, including DNAmethylation, histones, chromatin structure, andnoncodingRNAs, aremodified throughout life in response to environmental and behavioral influences. With each new generation, DNA methylation patterns are erased in gametesandresetafter fertilization,probably toprevent theseepigeneticmarks frombeing transferred fromparents to their offspring. However, some recent animal studies suggest an apparent resistance to complete erasure of epigenetic marks during early development, enabling transgenerational epigenetic inheritance. Whether there are similar mechanisms in humans remains unclear, with the exception of epigenetic imprinting. Nevertheless, a distinctly different mechanism—namely, intrauterine exposure to environmental stressors that may affect estab- lishment of the newly composing epigenetic patterns after fertilization—is often confused with transgenerational epigenetic inheritance. In this review, we delineate the definition of and requirement for transgenerational epi- genetic inheritance, differentiate it from the consequences of intrauterine exposure, and discuss the available evidence in both animal models and humans.—Van Otterdijk, S. D., Michels, K. B. Transgenerational epigenetic inheritance in mammals: how good is the evidence? FASEB J. 30, 2457–2465 (2016). www.fasebj.org

KEY WORDS: DNA methylation • intrauterine exposure • reprogramming • imprinting • IAP elements

Epigenetics is the meiotically and mitotically herita- ble potential for gene expression that does not involve variation in the DNA sequence (1, 2). Epigenetic events are important in orchestrating key biologic processes, such as cell differentiation, genomic imprinting, and X- chromosome inactivation (3–5). The primarymechanisms involved in epigenetic regulation include DNA methyl- ation, posttranslational histone modification, chromatin remodeling, and RNA-associated gene silencing by non- codingRNAs. Epigeneticmechanismsmediate diversified gene expression profiles to allow the generation of the variety of cells and tissues required in multicellular organisms. All cells in an organism contain essentially the same information, but different cell types vary in their phenotype, function, and expression profiles (6–9). Whereas an organism’s genetic information does not change during its life span (with the exception of acquired

mutations and DNA damage) epigenetic signatures are plastic (10–12) and can be modified in response to envi- ronmental and behavioral influences, such as nutrition, smoking, and air pollution (13–15). Alterations in the epigenome can result in subtle changes in cell differentia- tion or gene expression profiles or can result in cumulative detrimental effects in a cell and may compromise the normal function of the respective gene. Many diseases are associatedwith epigenetic changes. Epigenetic aberrations aremost extensively studied in cancer, but variation in the epigenome has also been described in cardiovascular dis- eases, autoimmune diseases, metabolic disorders, and neurodegenerative diseases (16–18).

Thus, although epigenetic variability enables relevant plasticity to adapt to environmental and lifestyle condi- tions, epigenetic aberrations may predispose to disease, and transferring the acquired epigenetic marks from parents to their offspring could affect the offspring’s de- velopment, plasticity, and disease susceptibility. The epi- genetic signature is erased in gametes and reset after fertilization, likely to prevent genome-wide epigenetic inheritance. In mammals, genome-wide DNA demethy- lationoccurs in 2 stepsduring early development. The first complete demethylation occurs in the parental gametes when the DNA methylation marks are erased in 2 steps accompanied by the restoration of developmental potency (19). An active and rapid demethylation event mediated

ABBREVIATIONS:DNMT, DNA methyltransferase; DOHaD, developmental origins of health and disease; DPPA, developmental pluripotency associ- ated; DZ, dizygotic; IAP, intracisternal A particle; lnRNA, long noncoding RNA; LTR, long terminal repeat; miRNA, microRNA; MZ, monozy- gotic; ncRNA, noncoding RNA; PGC, primordial germ cell; TET, 10–11 translocation 1 Correspondence: Obstetrics and Gynecology Epidemiology Center, Brig- ham and Women’s Hospital, Harvard Medical School, 221 Longwood Ave., Boston, MA 02115, USA. E-mail: kmichels@bwh.harvard.edu

doi: 10.1096/fj.201500083

0892-6638/16/0030-2457 © FASEB 2457

Downloaded from www.fasebj.org by (205.146.48.2) on October 08, 2018. The FASEB Journal Vol. ${article.issue.getVolume()}, No. ${article.issue.getIssueNumber()}, pp. 2457-2465.

by 10–11 translocation (TET) proteins is followed by a passive loss of DNA methylation marks during sub- sequent cell divisions. It has been thought that active demethylation mainly occurs in the paternal gametes and that the maternal gametes were mainly passively deme- thylated (20–22), because the maternal pronucleus may be protected from TET3-mediated conversion by devel- opmental pluripotency associated (DPPA)-3 (also known as stella) (23). However, in a recent study, Guo et al. (24) suggested that both active demethylation by TET3 and passive demethylation are important in both parental gametes. Other mechanisms, such as the base excision repair systems, may also operate in conjunction with the TETproteins, todriveDNAdemethylation (25). Following this first set of demethylation, reestablishment of DNA methylation marks commences during the establishment of the primordial germ cells (PGCs). The second general demethylation occurs after fertilization in the inner cell mass of the developing embryo (26).De novomethylation is established in the blastocyst associated with cellular differentiation.

Notably, differentiallymethylated regions of imprinted genes and retrotransposable elements are exempted from this second de- and remethylation step. In both mice and humans, imprinted regions undergo rapid DNA deme- thylationduring early germ-linedevelopment (27–29), but they are protected from the second round of demethyla- tion,whichenables themtomaintain theirparent-of-origin methylation state (30–32). In mouse models some other repeats, such as intracisternal A particle (IAP), endoge- nous retroviral sequence-1, and single copy sequences, have also been reported to escape epigenetic reprogram- ming (33–35). Whether this resistance to the second de- and remethylation pathway manifests an evolutionary path to ease transfer of epigenetic marks from parent to offspring remains unclear, but intergenerational inheri- tance may provide some functional advantages. For ex- ample,maintenance ofmethylationmaybe necessary to prevent transcriptional activity of the transposable el- ements and may reduce the risk of germ-line mutations through dysregulation of adjacent genes (27).

The evolutionary potential of the epigenetic code links back to theories of both Lamarck and Darwin. Lamarck proposed that the environment alters pheno- type in a heritable manner, which is consistent with the concept that environmental exposures at critical devel- opmental windows can promote epigenetic inheritance of epimutations in the germ line, which may increase phenotypic variation (36). Darwin argued that natural selection favors the survival or reproductive success of those with the greatest ability to adapt (36). This theory favors intergenerational plasticity and heritable adap- tive phenotypic variation.

Transgenerational epigenetic inheritance requires an incomplete erasure of the epigenetic signature during de- velopmental reprogramming permitting the transfer of epigenetic marks from parents onto their offspring over subsequent generations. Most studies investigating this phenomenon are based on animal models, leaving it un- clear whether transgenerational epigenetic inheritance ex- ists in humans. Moreover, germ-line inherited epigenetic

traits must be differentiated from intrauterine exposures, as both may shape the neonates epigenetic profile and can result in similarities between parents and offspring in their epigenetic code and phenotypes.

In this review, we will define the concept of trans- generational epigenetic inheritance, discuss the available data, and draw distinctions between transgenerational epigenetic inheritance and other phenomena, such as in- trauterine exposures, and their implications for the estab- lishment of the epigenome after fertilization.

INTRAUTERINE EXPOSURES

Any environmental stressor that acts during early devel- opmentmay affect the establishment of the epigenome and in turn the individual’s phenotype in the short and long term. For this reason there has been considerable interest among the Developmental Origins of Health and Disease (DOHaD) community in exploring the epigenetic under- pinningsof exposuresduringpregnancy thataffect later-life susceptibility to chronic disease (37, 38). The fetus is ex- posed in utero to some of the same environmental stressors as the mother. As a result, intrauterine experiences may induce fetal reprogramming via alterations in the epigenetic code before birth, with no inheritance from the parents.

The epigenome is thought to be particularly vulnerable to environmental factors during embryogenesis, which becomes evident from numerous studies that have re- ported serious consequences in later life caused by in- trauterine stressors. In animal studies, altered levels of the offspring’s stress response, glucosemetabolism, blood pressure, cholesterol metabolism, and cardiac energy metabolismwere related tomaternal diet, stress, and even traumatic exposures during pregnancy, and these associ- ations were complemented by differential epigenetic pat- terns and differentially expressed genes (35, 39–43). Maternal behavior may also affect offspring epigenetics. The levels ofmaternal licking, grooming, and arched-back nursing in mouse pups have been reported to influence their level of fearfulness. The impact of maternal care on the development of stress reactivity has been suggested to be mediated by changes in the levels of expression of specific genes inbrain regions that regulate behavioral and endocrine responses to stress (44), and maternal behavior has been suggested to alter DNA methylation and chro- matin structure in these pups (44, 45). However, these observations still have to be confirmed by other research groups.

In humans, parental nutritional and smoking behavior during pregnancy affect the offspring’s risk of cardiovas- cular and metabolic diseases, schizophrenia, and antiso- cial personality disorders (46–49); however, any epigenetic involvement remains less clear. Whereas maternal smok- ing has consistently been linked to demethylation of the AHRR gene, the methylation status of the human NR3C1 gene in newborns is sensitive to prenatal maternal mood, and periconceptional exposure to famine has been asso- ciated with lower DNA methylation of the IGF2 gene in adulthood(50–52), examplesofepigeneticdifferences linking prenatal exposures and adult disease outcomes are still

2458 Vol. 30 July 2016 VAN OTTERDIJK AND MICHELSThe FASEB Journal x www.fasebj.org

Downloaded from www.fasebj.org by (205.146.48.2) on October 08, 2018. The FASEB Journal Vol. ${article.issue.getVolume()}, No. ${article.issue.getIssueNumber()}, pp. 2457-2465.

missing, primarily because of the lack of availability of biospecimens at birth in longitudinal studies.

TRANSGENERATIONAL EPIGENETIC INHERITANCE

Intergenerational epigenetic inheritance is the transfer of epigenetic marks from the gametes to the embryo for 1 generation. It requires incomplete erasure of the parental epigenetic marks, thus avoiding epigenetic reprogram- ming in the gametes and during early embryo develop- ment. For transgenerational epigenetic inheritance to occur, epigenetic marks and phenotypes must be trans- ferredacross subsequentgenerations. Inagestating female, the phenotypic changes would have to be maintained for at least 4 generations. When the gestating mother (F0) is exposed to an environmental challenge, her embryo (F1) and the already developing germ line (F2) of the embryo, are also directly exposed (Fig. 1). Thus, a third generation’s phenotype may result from the grandmother’s experi- ences via intrauterine exposure and does not represent inheritance. In investigationsof exposures operatingbefore gestation via the maternal germ line or of inheritance via the paternal germ line, the third-generation offspring phe- notype is sufficient to establish transgenerational epige- netic inheritance (53, 54).

Epigenetic marks may be inherited through several pathways. We will discuss these in the sections below.

IMPRINTING

The strongest evidence for transgenerational epige- netic inheritance in mammals is genomic imprinting. Imprinted genes are expressed by only one of the 2 pa- rentally inherited alleles, whereas the other parental allele is silenced by epigenetic mechanisms in a parent- of-origin–specific manner. The parental specificity of the active allele and silenced allele has been faithfully maintained throughout generations. To establish these parental imprints, the germ cells must first lose their inherited maternal and paternal imprints, and the pa- rental imprints must then be reestablished in an allele- specific manner during gamete formation (55, 56). Maternal DNA methylation imprints are established during oogenesis at different time points, depending on the imprinted gene loci, whereas paternal imprints are established during spermatogonial differentiation in the adult testis (57). Once established these epigeneticmarks are able to resist postfertilization global epigenetic reprogramming in the preimplantation embryo through the interaction of the chromatin and DNA-modifying factors zinc finger protein (ZFP)-57 with tripartite motif containing (TRIM)-28, which are attracted tomethylated imprinting control regions (58), or by specific factors, such as DPPA3, that prevent DNA demethylation by binding H3K9me2 and blocking TET3 activity (23). Be- cause of the selective nature of epigenetic reprogram- ming in the preimplantation embryo, inheritance of parental imprints by the new embryo is permissive.

Since the imprinted regions are able to resist the second wave of reprogramming, errors that may occur during the first wave of erasure of parental imprints are maintained in the embryo. Several disorders, such as Beckwith-Wiedemann, Prader-Willi, and Angelman syndromes, are caused by loss of imprinting that results in biallelic expression of the respective imprinted gene. Although these imprintingdefectsmay also be caused by spontaneous epimutations, Buiting et al. (59) reported that loss of imprinting in humans may be the result of a failure to erase the parental imprint. In their study, im- printing defects in a subset of patients with Angelman syndrome occurred on the chromosome inherited from the maternal grandparents, whereas, in a subset of pa- tients with Prader-Willi syndrome, the imprinting defect occurred at the chromosome inherited from the paternal grandmother. While incomplete erasure of the parental imprint may explain some of these phenomena, the observed imprinting defects may also have occurred after the erasure of parental imprints or could result from the patient’s genetic background (59).

REPETITIVE RETROTRANSPOSONS

Maintaining genomic stability is vital for mammalian survival, and several diseases are a consequence of the inability to maintain genomic stability. For example, can- cer is associated with site-specific hypermethylation of CpG islands at promoters and global DNA hypo- methylation at repetitive and satellite regions, such as retrotransposons (60–62). One of the best-studied long terminal repeat (LTR) retrotransposons in mice is the IAP element. IAPs are retroviruslike repetitive DNA elements that possess an LTR region that functions as a promoter. These IAP elements have been suggested to provide a potential pathway of epigenetic inheritance. They are un- der the control of DNA methyltransferase 1 (DNMT1), as transcript levels are elevated in mouse embryos that are DNMT1deficient (63). CpG islands located close to an IAP showed consistently high methylation levels across all developmental stages (64), and methylation of the LTR sequences of most of the IAP element copies in themouse genome persisted through the wave of demethylation that occurs in the preimplementation embryo. A significant fraction of these IAP genomes remained essentially unreprogrammed, even in PGCs (27, 65, 66).

Because of the ability of IAPs to resist the secondwave of reprogramming, IAPs are a potential way of epigenetic inheritance in mammals, and several studies have in- vestigated this in mice. IAPs were reported to be essential in thevariablemethylationaffecting the inheritanceof coat color and tail length in Agouti viable yellow and Axin- fusedalleles inmice (66, 67), andmethylationof theAgouti viable yellow gene in offspring can be modulated by the availability of methyl donors in the maternal diet (68). In utero exposure to methyl donors by the F1 generation in pseudoagouti mice resulted in F2 generation mice with pseudoagouti phenotypes (69). Supplemented mice were more variable in their methylation levels, with the highest variation in mice that were supplemented over several

TRANSGENERATIONAL EPIGENETIC INHERITANCE IN MAMMALS 2459

Downloaded from www.fasebj.org by (205.146.48.2) on October 08, 2018. The FASEB Journal Vol. ${article.issue.getVolume()}, No. ${article.issue.getIssueNumber()}, pp. 2457-2465.

subsequent generations. Supplementation also affected the offspring’s disease susceptibility: supplemented off- spring were protected from developing obesity and di- abetes mellitus (70). It is important to note, however, that inheritance of these traits by subsequent and unexposed

generations has not been shown. Also, these mouse ex- periments were performed with inbred mouse strains, which may differ in their ability to reprogram epige- netic marks at IAPs after fertilization (67), and studies performed in one inbred mouse strain may not be

Figure 1. Epigenetic inheritance via the female and male germ line. If a gestating mother (F0, blue) is exposed to environmental stressors, her fetus (green) and its already developing germ line (red) are also directly exposed. As a result, phenotypes observed up to the F2 generation (red) may result from the grandmother’s experience. The F3 generation (yellow) is the first generation that has not been exposed to these environmental stressors, thus phenotypes observed in this generation could represent transgenerational epigenetic inheritance. If inheritance via the male or the maternal germ line before gestation is investigated, environmental stressors may affect the F0 generation (blue) and the developing germ line (red). As a result, the first generation that could represent transgenerational epigenetic inheritance is F2 (yellow).

2460 Vol. 30 July 2016 VAN OTTERDIJK AND MICHELSThe FASEB Journal x www.fasebj.org

Downloaded from www.fasebj.org by (205.146.48.2) on October 08, 2018. The FASEB Journal Vol. ${article.issue.getVolume()}, No. ${article.issue.getIssueNumber()}, pp. 2457-2465.

reproducible in other strains. Thus, genetic variation may be an alternative explanation for these phenomena rather than epigenetic inheritance. Moreover, the resistance of IAPs to demethylation during development may be a protective trait that prevents IAP retrotransposition, which may induce mutations (65). Thus, the maintenance of DNA methylation patterns may be a bystander rather than the driver, and it is an exceptional situation, as it does not occur on a genome-wide level.

Even though a large portion of the human genome is composed of transposable elements, a similar mecha- nism has not been identified in humans. However, trans- posable elements in humans are also highly methylated (71) and in a recent study, Tang et al. (72) reported that some retrotransposon-associated and single-copy regions in the human genome resist DNA demethylation. As a result, it is possible that factors affecting the degree of methylation cause variable gene expression in humans similar to that in mice.

RNA-DEPENDENT PROCESSES

Epigenetic inheritance may also be mediated by non- coding RNAs (ncRNAs), such as microRNA (miRNA) or long noncoding RNA (lnRNA), and RNA-dependent processes may contribute to the transmission of acquired traits in mammals. ncRNAs have key functions during early development; they control embryonic gene expres- sion, promote developmental transitions, and maintain developmental states (73).

Animal studies suggest that initial inductions resulting in a modification of the early embryonic genome to miRNAsmaybe sufficient to induce aheritablephenotype mediated through miRNAs. Mouse experiments have shown that stress before breeding affects the expression profiles in spermmiRNA and the responsiveness of stress by the offspring (74, 75). However, these alterations are present only in sperm of the F1 generation and not in later generations. Two other studies suggested that an initial induction involving a modification of the expression of critical miRNAs could be inherited both paternally and maternally by the next-generation offspring and could affect the offspring’s phenotype in mice (76, 77). Both studies found specific miRNA sequences in the sperm of the affected animals, whichwas supported by an increase in transcription in the study by Rassoulzadegan et al. (76). Wagner et al. (77) reported that this phenotype could even be transmitted for at least 3 generations. Another study by Padmanabhan et al. (78), reported that aMtrr deficiency in grandparents and parents, leading to a defect in folate metabolism, resulted in epigenetic effects on the daugh- ter’s maternal environment and their gametes, leading to congenital abnormalities. An effect of this Mtrr mutation was foundup to5generations in thewild-typeoffspring. It remains unclear which mechanism underlies this obser- vation, but it may be epigenetic inheritance, even though other possibilities cannot be excluded. For example, the initial mutation in the gene may lead to a perturbed reg- ulation of the nucleotide biosynthesis pathway that may result in deficiencies in DNA repair mechanisms and

potentially genetic mutations (79). Also, the epigenetic mark underlying the stability of the modified phenotypes observed in these studies and the exact role of themiRNAs has yet to be elucidated.

To our knowledge, no data are currently available on the potential role of lnRNAs in transgenerational epige- netic inheritance.

In human sperm, miRNA expression profiles differed between smoking and nonsmoking volunteers (80), but it remains unknown whether these miRNA profiles are maintained during early embryogenesis. In a study on human parent–offspring triads, regulatory scores of most miRNAs correlated between parents and offspring, sug- gestingapossibleheritabilityofmiRNAs(81).However, in humans, consistency in miRNA profiles across subsequent generations has not been reported, nor is there any other proof of miRNA-induced epigenetic inheritance to date.

HISTONE RETENTION

Histones are small proteins that constitute major building blocks of the chromatin structure; they have a positively charged central fold domain and terminal tails. The tails are crucial for normal function of cellular processes, as they are targets for posttranslational modifications, such as acetylation, methylation, phosphorylation, and ubiquiti- nation.Anegative chargeof thehistone tailwill result inan open chromatin structure, allowing access to the tran- scription machinery, whereas positively charged histone tails provide the ability to fold the chromatin, protecting it from transcription. Histones have been reported to be an important factor in early life development, as embryos deficient of certain histone methyltransferases display se- vere growth retardation and early lethality (82, 83).

Todate, little support is available for a role ofhistones in the process of transgenerational epigenetic inheritance. Studies in both mice and humans include only early stage embryos, and thus it remains unclear what the impact of these modifications is for the offspring in later stages of development.

In mice, specific histone modifications were found to undergo reprogramming during early embryo develop- ment. Hyperacetylated histone H4, Me(Arg17)[3H], and Me(Arg3)H4 marks were removed during the metaphase in eggs and early embryos, whereas the histone marks Me(Lys9)[3H], Me(Lys4)[3H], and Ph(Ser1)H4/H2A were reported to be stable until at least the blastocyst stage of development (84).

A study with human sperm suggested that epige- netic inheritance may transit through an incomplete replacement of histones by protamines during game- togenesis (85). In this study, correlations were observed between H3K4me in sperm and early expression in the embryo. A human embryo study suggested that con- stitutive heterochromatin marks in the embryo are transmitted and retained from human spermatozoa (86). However, it remains to be clarified whether these histone marks are maintained during later embryonic development and whether they can be propagated into subsequent generations.

TRANSGENERATIONAL EPIGENETIC INHERITANCE IN MAMMALS 2461

Downloaded from www.fasebj.org by (205.146.48.2) on October 08, 2018. The FASEB Journal Vol. ${article.issue.getVolume()}, No. ${article.issue.getIssueNumber()}, pp. 2457-2465.

DISCUSSION

Transgenerational epigenetic inheritance is a topic of great interest as the potential relevance of epigenetics for DOHaD is being scrutinized. However, many questions remain. If transgenerational inheritance is present in the formof incomplete erasure, it remainsuncertainwhether it is an error or there is an evolutionary advantage to main- tainingepigeneticmarks reflecting theparents’ experience.

It has been observed in animal experiments and in ep- idemiologic studies inhumans that the impact ofnutrition, smoking, irradiation, and even traumatic exposures in the parents andgrandparentsmayaffect the (grand)children’s phenotype or risk of diseases, and similarities in the epi- genetic profile between parent and offspring have been found (39–43, 46–52, 87–91). These observations may be caused by intrauterine exposures, germ-line–mediated transmission of phenotypic traits through genetic muta- tions,mutations in theDNArepairmechanisms, orgenetic mutations in epigenetic modifiers, rather than to the con- tinuity of epigenetic marks between generations. More- over, a shared environment may explain parental and sometimes even grandparental, similarities in the epige- netic profile, rather than transgenerational epigenetic inheritance.

A condition for transgenerational epigenetic inheri- tance is that the epigenetic mark and the associated phe- notype are maintained for at least 4 generations in a gestating female and for at least 3 generations if pre- gestational exposures or inheritance via themale germ line are investigated (53, 54). Few studies have considered the maintenance of epigenetic marks for these durations. Wagner et al. (77) observed in a mouse experiment that miRNAsweremaintained over 3 generations of offspring, and a study by Padmanabhan et al. (78) in mice suggested that an Mtrr mutation may impact epigenetic inheritance for up to 5 generations. Nonetheless, these studies were performed in animals that were genetically inbred, and it remains unclear whether these results are reproducible in other mouse strains.

The most convincing examples of potential pathways of transgenerational epigenetic inheritance come from mouse studies suggesting resistance to epigenetic reprogramming by repetitive retrotransposons or in- heritanceofmiRNAprofiles.Differences in the epigenome between humans and mice do not permit direct inference to humans (92, 93), and to date there is no evidence of transgenerational epigenetic inheritance in humans, ex- cept for the parent-of-origin specificity of genomic im- printing. Some histones have been reported to be stably inherited from the parents to the offspring during the first phases of embryonic development; correlations in the regulatory effect scores of miRNAs were observed be- tween parents and offspring; and in imprinting disorders, some of the imprinting defect in humansmay be the result of a failure to erase the parental imprint. However, in all these examples, it remains unclear whether those marks are preserved during later stages of development and in subsequent generations and other possible explanations for the observed effects cannot be excluded, such as the influence of genetic background. The genetic contribution

toDNAmethylationwas investigatedbyGertz et al. (94) in 3 humangenerations. They reported thatmost variation in DNAmethylation in the genome can be explained by ge- notype and that thegenetic influence exceeds the influence of imprinting on genome-wide methylation levels. Their data are further supported by twin studies, where evi- dence was found for within-pair epigenetic variability, including bothmethylation and gene expression analyses, in multiple tissues at birth (95–97). Methylation and gene expression differences withinmonozygotic (MZ) twin pairs were observed to be smaller than those observed for dizy- gotic (DZ) twins (95, 96), suggesting a genetic contribution tomethylationandgeneexpressionprofiles.MZdichorionic twins, each with their own placenta, displayed greater within-pair expression discordance than did MZ mono- chorionic twinswhosharedaplacenta (97), likely influenced by the intrauterine environment. These observations are in line with the concept that perceived inheritance of DNA methylation is driven by genetics and intrauterine exposure rather than transgenerational epigenetic inheritance.

FUTURE PERSPECTIVE

The study of transgenerational epigenetic inheritance in humans is more complex than in animals for several reasons. First, because transgenerational epigenetic in- heritance has to be maintained for several generations, longitudinal multigenerational studies are required with biospecimens available for 3 to 4 generations. To date, no such studies have been conducted. Second, studies in hu- mans are characterized by individual variation, stochastic differences, and ethical restrictions and thus differ pro- foundly from the controlled environment of animal stud- ies. Last, the heterozygosity of the human population makes it difficult to distinguish between genetics and epigenetics. Only when individuals are truly genetically identical and exhibit a range of phenotypes that are heri- tablemay thesebe attributed toepigenetic variation.These factors make it hard to study transgenerational epigenetic inheritance and its potential players in humans.

To fully understand transgenerational epigenetic in- heritance, the influence of intrauterine exposures and a shared postnatal environment on the epigenetic signature has to be studied in greater detail to enable a clear dis- tinction between these exposures and transgenerational epigenetic inheritance. Future studies may focus on un- derstanding how epigenetic marks during early life de- velopmentare established, inaddition to identifyepigenetic marks that are maintained over subsequent generations. Shedding light on these processes may help to understand whether, and how, parental epigenetic marks influence offspring’s health and disease susceptibility through transgenerational epigenetic inheritance and intrauterine exposures.

CONCLUSIONS

Several processes that may be part of intergenerational epigenetic inheritance have been identified in mice.

2462 Vol. 30 July 2016 VAN OTTERDIJK AND MICHELSThe FASEB Journal x www.fasebj.org

Downloaded from www.fasebj.org by (205.146.48.2) on October 08, 2018. The FASEB Journal Vol. ${article.issue.getVolume()}, No. ${article.issue.getIssueNumber()}, pp. 2457-2465.

However, much uncertainty remains as to whether these processes are truly epigenetically inherited or are influ- enced by other factors, such as intrauterine exposures and genetics, andwhether these epigenetic marks can be maintained over several subsequent generations. As a result, proof of principle of a widespread transgenera- tional epigenetic inheritance is lacking to date. Because of differences in the epigenome between mice and hu- mans and the limited number of studies performed in humans, the concept of transgenerational epigenetic in- heritance in humans remains equivocal.

REFERENCES

1. Ushijima,T.,Watanabe,N.,Okochi, E., Kaneda,A., Sugimura, T., and Miyamoto, K. (2003) Fidelity of the methylation pattern and its variation in the genome. Genome Res. 13, 868–874

2. Waterland, R. A., andMichels, K. B. (2007) Epigenetic epidemiology of thedevelopmental origins hypothesis.Annu.Rev.Nutr. 27, 363–388

3. Fedoriw, A., Mugford, J., and Magnuson, T. (2012) Genomic imprinting and epigenetic control of development. Cold Spring Harb. Perspect. Biol. 4, a008136

4. Wu, H., and Sun, Y. E. (2006) Epigenetic regulation of stem cell differentiation. Pediatr. Res. 59, 21R–25R

5. Li, Y., Tan,T., Zong,L.,He,D., Tao,W., andLiang,Q. (2012)Study of methylation of histone H3 lysine 9 and H3 lysine 27 during X chromosome inactivation in three types of cells. Chromosome Res. 20, 769–778

6. Moore,L.D.,Le,T., andFan,G.(2013)DNAmethylationand itsbasic function. Neuropsychopharmacology 38, 23–38

7. Roadmap Epigenomics Consortium. (2015) Integrative analysis of 111 reference human epigenomes. Nature 518, 317–330

8. Leung, D., Jung, I., Rajagopal, N., Schmitt, A., Selvaraj, S., Lee, A. Y., Yen, C. A., Lin, S., Lin, Y., Qiu, Y., Xie,W., Yue, F., Hariharan,M., Ray, P., Kuan, S., Edsall, L., Yang, H., Chi, N. C., Zhang, M. Q., Ecker, J. R., and Ren, B. (2015) Integrative analysis of haplotype-resolved epige- nomes across human tissues. Nature 518, 350–354

9. Ziller, M. J., Edri, R., Yaffe, Y., Donaghey, J., Pop, R., Mallard, W., Issner, R., Gifford, C. A., Goren, A., Xing, J., Gu, H., Cacchiarelli, D., Tsankov, A. M., Epstein, C., Rinn, J. L., Mikkelsen, T. S., Kohlbacher, O., Gnirke, A., Bernstein, B. E., Elkabetz, Y., andMeissner, A. (2015) Dissecting neural differentiation regulatory networks through epigenetic footprinting. Nature 518, 355–359

10. MuTHER Consortium. (2012) Epigenome-wide scans identify dif- ferentiallymethylated regions for ageandage-relatedphenotypes ina healthy ageing population. PLoS Genet. 8, e1002629

11. Gautrey,H. E., vanOtterdijk, S.D., Cordell,H. J., Newcastle 85+ Study Core Team, Mathers, J. C., and Strathdee, G. (2014) DNA methylation abnormalities at gene promoters are extensive and variable in the elderly and phenocopy cancer cells. FASEB J. 28, 3261–3272

12. Thompson, R. F., Atzmon, G., Gheorghe, C., Liang, H. Q., Lowes, C., Greally, J. M., and Barzilai, N. (2010) Tissue-specific dysregulation of DNA methylation in aging. Aging Cell 9, 506–518

13. Mathers, J. C., Strathdee, G., and Relton, C. L. (2010) Induction of epigenetic alterations by dietary and other environmental factors. Adv. Genet. 71, 3–39

14. Christensen, B. C., and Marsit, C. J. (2011) Epigenomics in environmental health. Front. Genet. 2, 84

15. Talikka, M., Sierro, N., Ivanov, N. V., Chaudhary, N., Peck, M. J., Hoeng, J., Coggins, C. R. E., and Peitsch, M. C. (2012) Genomic impact of cigarette smoke, with application to three smoking-related diseases. Crit. Rev. Toxicol. 42, 877–889

16. Van Otterdijk, S. D., Mathers, J. C., and Strathdee, G. (2013) Do age- relatedchanges inDNAmethylationplay a role in thedevelopmentof age-related diseases? Biochem. Soc. Trans. 41, 803–807

17. Jung,M., and Pfeifer, G. P. (2015) Aging andDNAmethylation. BMC Biol. 13, 7

18. Ribel-Madsen, R., Fraga, M. F., Jacobsen, S., Bork-Jensen, J., Lara, E., Calvanese,V., Fernandez,A.F., Friedrichsen,M.,Vind,B.F.,Højlund,K., Beck-Nielsen,H., Esteller,M., Vaag, A., andPoulsen, P. (2012)Genome- wide analysis of DNA methylation differences in muscle and fat from monozygotic twins discordant for type 2 diabetes. PLoS One 7, e51302

19. Seisenberger, S., Peat, J.R.,Hore,T.A., Santos,F.,Dean,W., andReik, W. (2013) Reprogramming DNAmethylation in the mammalian life cycle: building and breaking epigenetic barriers. Philos. Trans. R. Soc. Lond. B Biol. Sci. 368, 20110330

20. Gu,T. P.,Guo, F., Yang,H.,Wu,H.P.,Xu,G.F., Liu,W.,Xie,Z.G., Shi, L., He, X., Jin, S. G., Iqbal, K., Shi, Y. G., Deng, Z., Szabó, P. E., Pfeifer, G. P., Li, J., andXu,G.L. (2011)The role ofTet3DNAdioxygenase in epigenetic reprogramming by oocytes. Nature 477, 606–610

21. Rougier, N., Bourc’his, D., Gomes, D. M., Niveleau, A., Plachot, M., Pàldi, A., and Viegas-Péquignot, E. (1998) Chromosomemethylation patterns during mammalian preimplantation development. Genes Dev. 12, 2108–2113

22. Iqbal, K., Jin, S. G., Pfeifer, G. P., and Szabó, P. E. (2011) Reprogramming of the paternal genome upon fertilization involves genome-wide oxidation of 5-methylcytosine. Proc. Natl. Acad. Sci. USA 108, 3642–3647

23. Nakamura, T., Liu, Y. J., Nakashima, H., Umehara, H., Inoue, K., Matoba, S., Tachibana, M., Ogura, A., Shinkai, Y., and Nakano, T. (2012)PGC7bindshistoneH3K9me2 toprotect againstconversionof 5mC to 5hmC in early embryos. Nature 486, 415–419

24. Guo, F., Li, X., Liang,D., Li, T., Zhu, P., Guo,H.,Wu,X.,Wen, L., Gu, T. P., Hu, B., Walsh, C. P., Li, J., Tang, F., and Xu, G. L. (2014) Active andpassivedemethylationofmale and femalepronuclearDNA in the mammalian zygote. Cell Stem Cell 15, 447–458

25. Santos, F., Peat, J., Burgess, H., Rada, C., Reik, W., and Dean, W. (2013) Active demethylation in mouse zygotes involves cytosine deamination and base excision repair. Epigenetics Chromatin 6, 39

26. Smith, Z. D., Chan,M.M.,Mikkelsen, T. S., Gu,H., Gnirke, A., Regev, A., and Meissner, A. (2012) A unique regulatory phase of DNA methylation in the early mammalian embryo. Nature 484, 339–344

27. Hajkova, P., Erhardt, S., Lane, N., Haaf, T., El-Maarri, O., Reik, W., Walter, J., and Surani, M. A. (2002) Epigenetic reprogramming in mouse primordial germ cells.Mech. Dev. 117, 15–23

28. Brandeis,M., Kafri, T., Ariel,M., Chaillet, J. R.,McCarrey, J., Razin, A., and Cedar, H. (1993) The ontogeny of allele-specific methylation associatedwith imprinted genes in themouse.EMBOJ. 12, 3669–3677

29. Guo,H., Zhu, P., Yan,L., Li, R.,Hu,B., Lian, Y., Yan, J., Ren,X., Lin, S., Li, J., Jin, X., Shi, X., Liu, P., Wang, X., Wang,W., Wei, Y., Li, X., Guo, F., Wu, X., Fan, X., Yong, J., Wen, L., Xie, S. X., Tang, F., and Qiao, J. (2014) The DNA methylation landscape of human early embryos. Nature 511, 606–610

30. Sakashita, A., Kobayashi, H., Wakai, T., Sotomaru, Y., Hata, K., and Kono, T. (2014) Dynamics of genomic 5-hydroxymethylcytosine during mouse oocyte growth. Genes Cells 19, 629–636

31. Kafri, T., Gao, X., and Razin, A. (1993) Mechanistic aspects of genome-wide demethylation in the preimplantation mouse embryo. Proc. Natl. Acad. Sci. USA 90, 10558–10562

32. Oswald, J., Engemann, S., Lane, N., Mayer, W., Olek, A., Fundele, R., Dean,W., Reik, W., andWalter, J. (2000) Active demethylation of the paternal genome in the mouse zygote. Curr. Biol. 10, 475–478

33. Borgel, J., Guibert, S., Li, Y., Chiba, H., Schübeler, D., Sasaki, H., Forné, T., and Weber, M. (2010) Targets and dynamics of promoter DNA methylation during early mouse development. Nat. Genet. 42, 1093–1100

34. Hackett, J. A., Sengupta, R., Zylicz, J. J., Murakami, K., Lee, C., Down, T. A., and Surani, M. A. (2013) Germline DNA demethylation dynamics and imprint erasure through 5-hydroxymethylcytosine. Science 339, 448–452

35. Radford,E. J., Ito,M., Shi,H.,Corish, J.A., Yamazawa,K., Isganaitis, E., Seisenberger, S., Hore, T. A., Reik, W., Erkek, S., Peters, A. H., Patti, M. E., and Ferguson-Smith, A. C. (2014) In utero effects: in utero undernourishment perturbs the adult sperm methylome and intergenerational metabolism. Science 345, 1255903

36. Skinner, M. K. (2015) Environmental epigenetics and a unified theory of the molecular aspects of evolution: a neo-lamarckian con- cept that facilitates neo-Darwinian evolution. Genome Biol. Evol. 7, 1296–1302

37. Sookoian, S., Gianotti, T. F., Burgueño, A. L., and Pirola, C. J. (2013) Fetal metabolic programming and epigenetic modifications: a systems biology approach. Pediatr. Res. 73, 531–542

38. Rhee,K.E., Phelan, S., andMcCaffery, J. (2012)Early determinantsof obesity: genetic, epigenetic, and in utero influences. Int. J. Pediatr. 2012, 463850

39. Alkemade, F. E., van Vliet, P., Henneman, P., van Dijk, K. W., Hierck, B. P., van Munsteren, J. C., Scheerman, J. A., Goeman, J. J., Havekes, L.M.,Gittenberger-deGroot,A.C., vandenElsen, P. J., andDeRuiter, M. C. (2010) Prenatal exposure to apoE deficiency and postnatal

TRANSGENERATIONAL EPIGENETIC INHERITANCE IN MAMMALS 2463

Downloaded from www.fasebj.org by (205.146.48.2) on October 08, 2018. The FASEB Journal Vol. ${article.issue.getVolume()}, No. ${article.issue.getIssueNumber()}, pp. 2457-2465.

hypercholesterolemia are associated with altered cell-specific lysine methyltransferase and histone methylation patterns in the vascula- ture. Am. J. Pathol. 176, 542–548

40. Pruis, M. G. M., Lendvai, A., Bloks, V. W., Zwier, M. V., Baller, J. F. W., de Bruin, A., Groen, A. K., and Plösch, T. (2014) Maternal western diet primes non-alcoholic fatty liver disease in adult mouse offspring. Acta Physiol. (Oxf.) 210, 215–227

41. Huang, Y., He, Y., Sun, X., He, Y., Li, Y., and Sun, C. (2014) Maternal high folic acid supplement promotes glucose intolerance and insulin resistance in male mouse offspring fed a high-fat diet. Int. J. Mol. Sci. 15, 6298–6313

42. St-Cyr, S., andMcGowan, P. O. (2015) Programming of stress-related behavior and epigenetic neural gene regulation in mice offspring throughmaternal exposure to predator odor. Front. Behav.Neurosci. 9, 145

43. Maeyama, H., Hirasawa, T., Tahara, Y., Obata, C., Kasai, H., Moriishi, K., Mochizuki, K., and Kubota, T. (2015) Maternal restraint stress during pregnancy in mice induces 11b-HSD1-associated metabolic changes in the liversof theoffspring. J. Dev.Orig.HealthDis.6, 105–114

44. Francis, D., Diorio, J., Liu, D., andMeaney, M. J. (1999) Nongenomic transmission across generations of maternal behavior and stress responses in the rat. Science 286, 1155–1158

45. Weaver, I.C.G.,Cervoni,N.,Champagne,F.A.,D’Alessio,A.C., Sharma, S., Seckl, J. R., Dymov, S., Szyf, M., andMeaney, M. J. (2004) Epigenetic programming by maternal behavior. Nat. Neurosci. 7, 847–854

46. Ravelli, A. C., van der Meulen, J. H., Michels, R. P., Osmond, C., Barker, D. J., Hales, C. N., and Bleker, O. P. (1998) Glucose tolerance in adults after prenatal exposure to famine. Lancet 351, 173–177

47. Bygren, L. O., Tinghög, P., Carstensen, J., Edvinsson, S., Kaati, G., Pembrey, M. E., and Sjöström, M. (2014) Change in paternal grandmothers’ early food supply influenced cardiovascular mortality of the female grandchildren. BMC Genet. 15, 12

48. Roseboom, T. J., Painter, R. C., van Abeelen, A. F., Veenendaal,M. V., and de Rooij, S. R. (2011) Hungry in the womb: what are the consequences?Lessons fromtheDutch famine.Maturitas70, 141–145

49. Harris, H. R., Willett, W. C., and Michels, K. B. (2013) Parental smoking during pregnancy and risk of overweight and obesity in the daughter. Int. J. Obes. 37, 1356–1363

50. Joubert, B. R., Håberg, S. E., Nilsen, R. M., Wang, X., Vollset, S. E., Murphy, S. K., Huang, Z., Hoyo, C., Midttun, Ø., Cupul-Uicab, L. A., Ueland, P. M., Wu, M. C., Nystad, W., Bell, D. A., Peddada, S. D., and London, S. J. (2012)450Kepigenome-wide scan identifies differential DNA methylation in newborns related to maternal smoking during pregnancy. Environ. Health Perspect. 120, 1425–1431

51. Oberlander, T. F., Weinberg, J., Papsdorf, M., Grunau, R., Misri, S., and Devlin, A. M. (2008) Prenatal exposure to maternal depression, neonatal methylation of human glucocorticoid receptor gene (NR3C1) and infant cortisol stress responses. Epigenetics 3, 97–106

52. Heijmans, B. T., Tobi, E. W., Stein, A. D., Putter, H., Blauw, G. J., Susser, E. S., Slagboom, P. E., and Lumey, L. H. (2008) Persistent epigenetic differences associated with prenatal exposure to famine in humans. Proc. Natl. Acad. Sci. USA 105, 17046–17049

53. Skinner, M. K. (2008) What is an epigenetic transgenerational phenotype? F3 or F2. Reprod. Toxicol. 25, 2–6

54. Grossniklaus, U., Kelly, W. G., Ferguson-Smith, A. C., Pembrey, M., and Lindquist, S. (2013) Transgenerational epigenetic inheritance: how important is it? Nat. Rev. Genet. 14, 228–235

55. Wang,L., Zhang, J.,Duan, J.,Gao,X., Zhu,W.,Lu,X., Yang,L., Zhang, J., Li, G., Ci, W., Li, W., Zhou, Q., Aluru, N., Tang, F., He, C., Huang, X., andLiu, J. (2014) Programming and inheritance of parental DNA methylomes in mammals. Cell 157, 979–991

56. Barlow, D. P., and Bartolomei, M. S. (2014) Genomic imprinting in mammals. Cold Spring Harb. Perspect. Biol. 6, a018382

57. Kerjean, A., Dupont, J.-M., Vasseur, C., Le Tessier, D., Cuisset, L., Pàldi, A., Jouannet, P., and Jeanpierre, M. (2000) Establishment of the paternal methylation imprint of the human H19 and MEST/ PEG1 genes during spermatogenesis.Hum. Mol. Genet. 9, 2183–2187

58. Zuo, X., Sheng, J., Lau, H. T., McDonald, C.M., Andrade,M., Cullen, D. E., Bell, F. T., Iacovino,M., Kyba,M., Xu,G., andLi, X. (2012) Zinc finger protein ZFP57 requires its co-factor to recruit DNA methyl- transferases and maintains DNA methylation imprint in embryonic stem cells via its transcriptional repression domain. J. Biol. Chem. 287, 2107–2118

59. Buiting, K., Gross, S., Lich, C., Gillessen-Kaesbach, G., el-Maarri, O., and Horsthemke, B. (2003) Epimutations in Prader-Willi and Angelman syndromes: a molecular study of 136 patients with an im- printing defect. Am. J. Hum. Genet. 72, 571–577

60. Meng, H., Cao, Y., Qin, J., Song, X., Zhang, Q., Shi, Y., and Cao, L. (2015) DNA methylation, its mediators and genome integrity. Int. J. Biol. Sci. 11, 604–617

61. Luczak, M. W., and Jagodziński, P. P. (2006) The role of DNA methylation in cancer development. Folia Histochem. Cytobiol. 44, 143–154

62. Crichton, J. H., Dunican, D. S., Maclennan, M., Meehan, R. R., and Adams, I. R. (2014) Defending the genome from the enemy within: mechanisms of retrotransposon suppression in the mouse germline. Cell. Mol. Life Sci. 71, 1581–1605

63. Walsh, C. P., Chaillet, J. R., and Bestor, T. H. (1998) Transcription of IAP endogenous retroviruses is constrained by cytosine methylation. Nat. Genet. 20, 116–117

64. Seisenberger, S., Andrews, S., Krueger, F., Arand, J.,Walter, J., Santos, F., Popp, C., Thienpont, B., Dean, W., and Reik, W. (2012) The dynamics of genome-wide DNA methylation reprogramming in mouse primordial germ cells.Mol. Cell 48, 849–862

65. Lane, N., Dean,W., Erhardt, S., Hajkova, P., Surani, A.,Walter, J., and Reik, W. (2003) Resistance of IAPs to methylation reprogramming may provide a mechanism for epigenetic inheritance in the mouse. Genesis 35, 88–93

66. Morgan, H. D., Sutherland, H. G., Martin, D. I., and Whitelaw, E. (1999) Epigenetic inheritance at the agouti locus in the mouse. Nat. Genet. 23, 314–318

67. Rakyan, V. K., Chong, S., Champ, M. E., Cuthbert, P. C., Morgan, H. D., Luu, K. V., and Whitelaw, E. (2003) Transgenerational inheritance of epigenetic states at the murine Axin(Fu) allele occurs after maternal and paternal transmission. Proc. Natl. Acad. Sci. USA 100, 2538–2543

68. Wolff, G. L., Kodell, R. L., Moore, S. R., and Cooney, C. A. (1998) Maternal epigenetics and methyl supplements affect agouti gene expression in Avy/a mice. FASEB J. 12, 949–957

69. Cropley, J. E., Suter,C.M., Beckman,K. B., andMartin,D. I. K. (2006) Germ-line epigenetic modification of the murine A vy allele by nu- tritional supplementation. Proc.Natl. Acad. Sci. USA 103, 17308–17312

70. Li, C. C. Y., Cropley, J. E., Cowley,M. J., Preiss, T., Martin, D. I. K., and Suter, C. M. (2011) A sustained dietary change increases epigenetic variation in isogenic mice. PLoS Genet. 7, e1001380

71. Okae,H., Chiba,H.,Hiura,H.,Hamada,H., Sato, A., Utsunomiya, T., Kikuchi, H., Yoshida, H., Tanaka, A., Suyama, M., and Arima, T. (2014) Genome-wide analysis of DNA methylation dynamics during early human development. PLoS Genet. 10, e1004868

72. Tang, W. W. C., Dietmann, S., Irie, N., Leitch, H. G., Floros, V. I., Bradshaw, C. R., Hackett, J. A., Chinnery, P. F., and Surani, M. A. (2015)A unique gene regulatory network resets the human germline epigenome for development. Cell 161, 1453–1467

73. Pauli, A., Rinn, J. L., and Schier, A. F. (2011) Non-coding RNAs as regulators of embryogenesis. Nat. Rev. Genet. 12, 136–149

74. Gapp, K., Jawaid, A., Sarkies, P., Bohacek, J., Pelczar, P., Prados, J., Farinelli, L.,Miska, E., andMansuy, I.M. (2014) Implicationof sperm RNAs in transgenerational inheritance of the effects of early trauma in mice. Nat. Neurosci. 17, 667–669

75. Rodgers, A. B., Morgan, C. P., Bronson, S. L., Revello, S., and Bale, T. L. (2013) Paternal stress exposure alters spermmicroRNA content and reprograms offspring HPA stress axis regulation. J. Neurosci. 33, 9003–9012

76. Rassoulzadegan,M., Grandjean, V., Gounon, P., Vincent, S., Gillot, I., and Cuzin, F. (2006) RNA-mediated non-mendelian inheritance of an epigenetic change in the mouse. Nature 441, 469–474

77. Wagner, K. D., Wagner, N., Ghanbarian, H., Grandjean, V., Gounon, P., Cuzin, F., and Rassoulzadegan, M. (2008) RNA induction and inheritance of epigenetic cardiac hypertrophy in themouse. Dev. Cell 14, 962–969

78. Padmanabhan, N., Jia, D., Geary-Joo, C., Wu, X., Ferguson-Smith, A.C., Fung,E.,Bieda,M.C., Snyder, F. F.,Gravel,R.A.,Cross, J.C., and Watson, E.D. (2013)Mutation in folatemetabolismcauses epigenetic instability and transgenerational effects on development. Cell 155, 81–93

79. Heard, E., and Martienssen, R. A. (2014) Transgenerational epigenetic inheritance: myths and mechanisms. Cell 157, 95–109

80. Marczylo,E. L., Amoako,A.A., Konje, J.C.,Gant,T.W., andMarczylo, T. H. (2012) Smoking induces differential miRNA expression in human spermatozoa: a potential transgenerational epigenetic concern? Epigenetics 7, 432–439

81. Geeleher, P., Huang, S. R., Gamazon, E. R., Golden, A., and Seoighe, C. (2012) The regulatory effect of miRNAs is a heritable genetic trait in humans. BMC Genomics 13, 383

2464 Vol. 30 July 2016 VAN OTTERDIJK AND MICHELSThe FASEB Journal x www.fasebj.org

Downloaded from www.fasebj.org by (205.146.48.2) on October 08, 2018. The FASEB Journal Vol. ${article.issue.getVolume()}, No. ${article.issue.getIssueNumber()}, pp. 2457-2465.

82. Tachibana,M., Sugimoto,K., Nozaki,M., Ueda, J., Ohta, T., Ohki,M., Fukuda, M., Takeda, N., Niida, H., Kato, H., and Shinkai, Y. (2002) G9a histone methyltransferase plays a dominant role in euchromatic histone H3 lysine 9 methylation and is essential for early embryogenesis. Genes Dev. 16, 1779–1791

83. Dodge, J. E., Kang, Y.-K., Beppu,H., Lei, H., andLi, E. (2004)Histone H3-K9 methyltransferase ESET is essential for early development. Mol. Cell. Biol. 24, 2478–2486

84. Sarmento, O. F., Digilio, L. C., Wang, Y., Perlin, J., Herr, J. C., Allis, C. D., and Coonrod, S. A. (2004) Dynamic alterations of specific histone modifications during early murine development. J. Cell Sci. 117, 4449–4459

85. Hammoud, S. S., Nix, D. A., Zhang, H., Purwar, J., Carrell, D. T., and Cairns, B. R. (2009) Distinctive chromatin in human sperm packages genes for embryo development. Nature 460, 473–478

86. vandeWerken,C., vanderHeijden,G.W.,Eleveld,C.,Teeuwssen,M., Albert,M., Baarends,W.M., Laven, J. S., Peters, A.H., andBaart, E. B. (2014) Paternal heterochromatin formation in human embryos is H3K9/HP1 directed and primed by sperm-derived histone modifi- cations. Nat. Commun. 5, 5868

87. Dias, B. G., and Ressler, K. J. (2014) Parental olfactory experience influences behavior and neural structure in subsequent generations. Nat. Neurosci. 17, 89–96

88. Braunschweig,M., Jagannathan, V.,Gutzwiller, A., andBee,G. (2012) Investigations on transgenerational epigenetic response down the male line in F2 pigs. PLoS One 7, e30583

89. Dunn, G. A., and Bale, T. L. (2011) Maternal high-fat diet effects on third-generation female body size via the paternal lineage. Endocri- nology 152, 2228–2236

90. Carone, B. R., Fauquier, L., Habib, N., Shea, J. M., Hart, C. E., Li, R., Bock, C., Li, C., Gu, H., Zamore, P. D., Meissner, A., Weng, Z., Hofmann, H. A., Friedman, N., and Rando, O. J. (2010) Paternally induced transgenerational environmental reprogramming of metabolic gene expression in mammals. Cell 143, 1084–1096

91. Goodspeed, D., Seferovic, M. D., Holland, W., Mcknight, R. A., Summers, S.A., Branch,D.W., Lane,R.H., andAagaard,K.M. (2015)

Essential nutrient supplementation prevents heritable metabolic disease in multigenerational intrauterine growth-restricted rats. FASEB J. 29, 807–819

92. Casas, E., andVavouri, T. (2014) Sperm epigenomics: challenges and opportunities. Front. Genet. 5, 330

93. Chavez, S. L.,McElroy, S. L., Bossert, N. L., De Jonge, C. J., Rodriguez, M. V., Leong, D. E., Behr, B., Westphal, L. M., and Reijo Pera, R. A. (2014) Comparison of epigenetic mediator expression and function in mouse and human embryonic blastomeres. Hum. Mol. Genet. 23, 4970–4984

94. Gertz, J., Varley, K. E., Reddy, T. E., Bowling, K. M., Pauli, F., Parker, S. L., Kucera, K. S., Willard, H. F., andMyers, R. M. (2011) Analysis of DNA methylation in a three-generation family reveals widespread genetic influence on epigenetic regulation. PLoS Genet. 7, e1002228

95. Gordon,L., Joo, J.E., Powell, J. E.,Ollikainen,M.,Novakovic,B., Li,X., Andronikos, R., Cruickshank, M. N., Conneely, K. N., Smith, A. K., Alisch, R. S., Morley, R., Visscher, P. M., Craig, J. M., and Saffery, R. (2012)NeonatalDNAmethylationprofile inhuman twins is specified by a complex interplay between intrauterine environmental and genetic factors, subject to tissue-specific influence. Genome Res. 22, 1395–1406

96. Ollikainen, M., Smith, K. R., Joo, E. J., Ng, H. K., Andronikos, R., Novakovic, B., Abdul Aziz, N. K., Carlin, J. B., Morley, R., Saffery, R., and Craig, J. M. (2010) DNA methylation analysis of multiple tissues from newborn twins reveals both genetic and intrauterine components to variation in the human neonatal epigenome. Hum. Mol. Genet. 19, 4176–4188

97. Gordon, L., Joo, J. H., Andronikos, R., Ollikainen, M., Wallace, E. M., Umstad, M. P., Permezel, M., Oshlack, A., Morley, R., Carlin, J. B., Saffery, R., Smyth, G. K., and Craig, J. M. (2011) Expression discordance of monozygotic twins at birth: effect of intrauterine environment and a possible mechanism for fetal programming. Epigenetics 6, 579–592

Received for publication December 18, 2015. Accepted for publication March 21, 2016.

TRANSGENERATIONAL EPIGENETIC INHERITANCE IN MAMMALS 2465

Downloaded from www.fasebj.org by (205.146.48.2) on October 08, 2018. The FASEB Journal Vol. ${article.issue.getVolume()}, No. ${article.issue.getIssueNumber()}, pp. 2457-2465.