You are on page 1of 15

Neurobiology of Disease 132 (2019) 104591

Contents lists available at ScienceDirect

Neurobiology of Disease
journal homepage: www.elsevier.com/locate/ynbdi

Review

Multigenerational epigenetic inheritance: One step forward, two generations T


back
⁎ ⁎⁎
Jennifer J. Tuscher , Jeremy J. Day
Department of Neurobiology, Evelyn F. McKnight Brain Institute, University of Alabama at Birmingham, Birmingham, AL 35294, USA

A R T I C LE I N FO A B S T R A C T

Keywords: Modifications to DNA and histone proteins serve a critical regulatory role in the developing and adult brain, and
Epigenetics over a decade of research has established the importance of these “epigenetic” modifications in a wide variety of
Gene expression brain functions across the lifespan. Epigenetic patterns orchestrate gene expression programs that establish the
Heritability phenotypic diversity of various cellular classes in the central nervous system, play a key role in experience-
Transgenerational
dependent gene regulation in the adult brain, and are commonly implicated in neurodevelopmental, psychiatric,
DNA methylation
and neurodegenerative disease states. In addition to these established roles, emerging evidence indicates that
Non-coding RNA
Histone modifications epigenetic information can potentially be transmitted to offspring, giving rise to inter- and trans-generational
epigenetic inheritance phenotypes. However, our understanding of the cellular events that participate in this
information transfer is incomplete, and the ability of this transfer to overcome complete epigenetic repro-
gramming during embryonic development is highly controversial. This review explores the existing literature on
multigenerational epigenetic mechanisms in the central nervous system. First, we focus on the cellular me-
chanisms that may perpetuate or counteract this type of information transfer, and consider how epigenetic
modification in germline and somatic cells regulate important aspects of cellular and organismal development.
Next, we review the potential phenotypes resulting from ancestral experiences that impact gene regulatory
modifications, including how these changes may give rise to unique metabolic phenotypes. Finally, we discuss
several caveats and technical limitations that influence multigenerational epigenetic effects. We argue that
studies reporting multigenerational epigenetic changes impacting the central nervous system must be inter-
preted with caution, and provide suggestions for how epigenetic information transfer can be mechanistically
disentangled from genetic and environmental influences on brain function.

1. Introduction 2013; Karpova et al., 2017; Clayton et al., 2019), and the formation and
long-term storage of memory (Miller and Sweatt, 2007; Lubin et al.,
Epigenetic modifications provide an essential mechanism for the 2008; Gupta et al., 2010; Day and Sweatt, 2011; Day et al., 2013;
fine tuning of gene expression in response to extracellular signals and Zovkic et al., 2014). Despite the well-established role of epigenetic
changes in the environment. These alterations can occur in the form of regulation of a number of these neural processes, whether experience-
chromatin remodeling, direct covalent modifications to the DNA itself, dependent epigenetic marks acquired in an individual can be inherited
post-translational modifications to histone proteins, and activity of non- by the subsequent generation remains an open topic of debate.
coding RNAs. In the central nervous system, epigenetic mechanisms are Accumulating evidence supports the notion that external stimuli,
responsible for shaping early developmental programming (Fagiolini including environmental exposure to chemicals, extreme heat, psy-
et al., 2009; Feng et al., 2010; Bale, 2015), responses to stressful or chological stressors, traumatic experience, substances of abuse, and
appetitive stimuli (Renthal and Nestler, 2008; McEwen et al., 2012; Day dietary intake, can all shape epigenetic state at the cellular level in the
et al., 2013; Stankiewicz et al., 2013; Klengel and Binder, 2015), ac- exposed individual (Waddington, 1953; Francis et al., 1999; Anway
tivity-dependent changes in synaptic plasticity (Cortes-Mendoza et al., et al., 2005; Baccarelli and Bollati, 2009; Mathers et al., 2010; Gapp


Correspondence to: J. J. Tuscher, Department of Neurobiology, UAB The University of Alabama at Birmingham, SHEL 930, 1825 University Blvd, Birmingham, AL
35294-2182, USA.
⁎⁎
Correspondence to: J. J. Day, Department of Neurobiology, Evelyn F. McKnight Brain Institute, UAB The University of Alabama at Birmingham, SHEL 910, 1825
University Blvd, Birmingham, AL 35294-2182, USA.
E-mail addresses: jtuscher@uab.edu (J.J. Tuscher), jjday@uab.edu (J.J. Day).

https://doi.org/10.1016/j.nbd.2019.104591
Received 2 April 2019; Received in revised form 22 July 2019; Accepted 26 August 2019
Available online 27 August 2019
0969-9961/ © 2019 Elsevier Inc. All rights reserved.
J.J. Tuscher and J.J. Day Neurobiology of Disease 132 (2019) 104591

Fig. 1. Intergenerational vs. Transgenerational Inheritance. A, Intergenerational inheritance involves exposure of the parent generation (F0, male or non-pregnant
female) to an environmental factor or external stimuli (e.g., drugs of abuse, high fat diet, stress) that leads to an epigenetic alteration in the parent and parental
germline cells (F1; e.g., sperm or oocytes). Epigenetic modifications may be observed in somatic tissues of the F1 adult, but do not persist in the F2 generation. B, For
epigenetic effects to be considered transgenerational, modifications acquired in the F0 generation would need to be observed in the F1 and F2 generations, and
potentially beyond. For pregnant females that experience environmentally-induced epigenetic changes, the pregnant dam (F0), fetus (F1), and fetal germline cells
(F2) would all be exposed to the external stimulus. In this case, intergenerational effects encompass transmission from F0 to F2, and only persistence to the F3
generation would be considered a transgenerational effect.

et al., 2014; Norouzitallab et al., 2014; Vickers, 2014; Klengel and well documented in plant species, which do not require the passage of
Binder, 2015; Cadet, 2016; Nilsson et al., 2018; Walker and Nestler, epigenetic marks to occur through the germline (Quadrana and Colot,
2018). Alterations in epigenetic profile have also been observed in the 2016). The intriguing notion that mammalian gene expression pro-
offspring of exposed individuals (i.e., intergenerational inheritance grams could similarly be influenced by external stimuli to alter phe-
(Perez and Lehner, 2019)). However, whether the transfer of such notypic traits has more recently become a topic of great interest and
epigenetic marks can occur strictly via germline-dependent transmis- debate in biomedical sciences. This is particularly true for instances in
sion, in the absence of any direct exposure to the driving external sti- which DNA sequence-based inheritance (i.e., Mendelian genetics)
mulus, remains controversial. This review will first define the distinc- cannot fully explain the pattern heritability of certain phenotypic traits
tion between intergenerational and transgenerational epigenetic across multiple generations. Despite the passage of epigenetic mod-
inheritance (Skinner, 2008; Perez and Lehner, 2019), as these terms are ifications becoming an attractive candidate mechanism for the in-
frequently conflated in the literature. Next, cellular transmission of heritance of such traits, clear evidence that observable phenotypes are a
epigenetic modifications will be discussed, followed by putative me- direct result of germline-dependent transmission of epigenetic altera-
chanisms of heritable, multigenerational epigenetic transmission. Fi- tions has also been severely lacking.
nally, preclinical evidence for multigenerational propagation of epige- This pursuit is complicated by several remarkable features of re-
netic marks in mammals will be reviewed, as well as the translational productive biology. In mammals, haploid gamete cells (ovum and
implications of such work. For the purposes of this review, multi- sperm) each possess their own unique epigenetic landscapes prior to
generational inheritance encompasses both intergenerational in- fertilization. These patterns are dominated by repressive marks,
heritance (from F0 to F1), and transgenerational inheritance (F0 to F2 yielding transcriptionally inactive cells and (for sperm) a highly com-
or beyond; Fig. 1). pacted genome architecture in which the majority of histones have been
replaced by protamines (Balhorn, 2007). However, upon fertilization
these landscapes are rapidly altered, and established epigenetic marks
2. Germline reprogramming and intergenerational vs. are removed. This reprogramming is critical to establishing a totipotent
transgenerational epigenetic inheritance state in the zygote, but also necessary to ensure the genetic and epi-
genetic blueprint responsible for determining the fate and function of
Epigenetic inheritance across multiple generations has been most

2
J.J. Tuscher and J.J. Day Neurobiology of Disease 132 (2019) 104591

thousands of distinct cell types cannot simply be undone by parental regarding the putative mechanisms through which such effects would
experience. Without normal reprogramming of the germline, proper be perpetuated. In addition to direct alterations to the DNA itself, a host
cellular differentiation cannot be achieved. However, the erasure of of epigenetic factors, including DNA methylation, histone post-trans-
epigenetic marks does not occur with equal efficiency at all sites across lational modifications (PTMs), and small non-coding RNAs, can be
the genome (Guibert et al., 2012; Seisenberger et al., 2012; Tang et al., passed from one generation to the next via mitosis and meiosis. Below
2015), providing a small window of opportunity for some epigenetic we discuss potential epigenetic routes for cellular inheritance of gene
modifications to escape early periods of reprogramming. The matter of expression profiles, including DNA methylation, post-translation his-
which loci are protected during these phases of epigenomic repro- tone modifications, and small non-coding RNAs. Evidence for the
gramming, and why some regions escape while others do not, remains transmission of associated heritable phenotypic states will be outlined
an important open question. Another point that merits consideration is in later sections of the review.
why might modifications imposed by the environment be stably main-
tained across multiple generations? Such a mechanism that allows en- 3.1. Cellular transmission of epigenetic marks via DNA methylation
vironmental stimuli to influence gene expression programs in response
to external cues without directly altering the DNA has the potential to To discern how epigenetic inheritance across generations may
be adaptive. As such, understanding the complex interactions between occur, it is first useful to understand how epigenetic marks are trans-
the genome, epigenome, and environment may have important im- mitted at the cellular level. DNA methylation involves the covalent
plications for the heritability of disease states or adaptive fitness of the addition of a methyl group to cytosines in the 5th position of pyr-
species in general. imidine rings in the DNA backbone, forming 5-methylcytosine (5mC).
Although the transfer of environmentally induced epigenetic mod- This predominantly occurs within cytosine-guanine dinucleotides
ifications through mitotic and meiotic processes is relatively well es- (CpGs) in most cells, and is dynamically regulated by DNA methyl-
tablished in non-mammalian species (e.g., plants, nematodes, droso- transferases (DNMTs; (Holliday and Pugh, 1975; Bird, 1992)). The
phila (Cavalli and Paro, 1998, 1999; Quadrana and Colot, 2016; stable inheritance of DNA methylation patterns requires not only high
Serobyan and Sommer, 2017; Minkina and Hunter, 2018; Moore et al., fidelity DNA replication, but also faithful maintenance of cytosine
2019; Posner et al., 2019)), mammalian research on multigenerational methyl marks. DNA replication is incredibly accurate, with an error
epigenetic inheritance has proved considerably more challenging. This occurring only once per 107–108 bases copied (Kunkel, 2004). During
is in part due to a number of caveats that must be carefully considered replication, methylation marks on the parent DNA strand are en-
when designing and interpreting experiments. First, it is important to zymatically copied to the complimentary daughter strand by main-
distinguish between intergenerational vs. transgenerational epigenetic tenance methyltransferase DNMT1, which coordinates the transfer of
inheritance. The importance of this distinction extends beyond se- methyl groups to hemi-methylated CpG sites throughout cellular re-
mantics, as the mechanisms that govern epigenetic alterations to germ plication (Bird, 1992). This provides a potential route for methylation
cells as a result of direct exposure to an environmental stimulus may marks at hemi-methylated sites to be reestablished and perpetuated
differ from the mechanism through which acquired epigenetic marks across multiple rounds of cell division. Methylation of non-methylated
are maintained and propagated during cellular division and across sites can also occur throughout the life cycle of the cell, and this process
generations. Intergenerational inheritance involves exposure of the is coordinated by de novo methyltransferases DNMT3a and DNMT3b
parent generation (F0, male or non-pregnant female) to an environ- (Okano et al., 1999). A key issue regarding the transfer of epigenetic
mental factor (e.g., drugs of abuse, high fat diet, stress, etc.) that leads marks is the stability of methylation patterns across multiple rounds of
to an epigenetic alteration in the parent and likely the parental germ- cell division. DNA methylation is considerably less accurate than DNA
line cells (F1; e.g., sperm or oocytes). Observation of epigenetic marks replication with a ~96% fidelity rate, which roughly translates to 1
or phenotypic traits in offspring through this type of F0 to F1 trans- error for every 25 methylation marks copied per cell division (Laird
mission would be considered an intergenerational effect (see Fig. 1A). et al., 2004). This elevated error rate introduces variation in methyla-
Unlike intergenerational inheritance, transgenerational inheritance re- tion profiles due to faulty copying during replication and cell division,
lies strictly on germline transfer of epigenetic marks in the absence of which in turn yields cell populations with diverse methylation patterns
any direct environmental exposure. As such, for an effect to be con- (Silva et al., 1993). Thus, even at the cellular level within a single in-
sidered transgenerational, an inherited epigenetic mark and corre- dividual there are multiple opportunities for errors to be introduced
sponding trait would need to persist in the F2 generation, and poten- into methylation patterns, preventing highly accurate transmission
tially beyond (Fig. 1B). For pregnant females that experience during cell division. Taken together, this suggests that although DNMTs
environmentally-induced epigenetic changes, the pregnant dam (F0), may largely preserve methylation programs across the cell life cycle,
fetus (F1), and fetal germline cells (F2) would all be exposed to the this process is inherently less accurate than DNA replication. This issue
external stimulus. In this case, intergenerational effects encompass creates one of several substantial hurdles that must be overcome in
transmission from F0 to F2 (Fig. 1A), and only persistence to the F3 order for methylation marks to be stably transmitted not only during
generation would be considered a transgenerational effect (Fig. 1B). At cell division, but also across multiple generations.
present, limited evidence exists to support either mode of transge- The processes of active and passive DNA demethylation are also key
nerational epigenetic inheritance in mammals. Understanding the pu- to understanding the cellular maintenance of methylation programs.
tative mechanisms through which epigenetic alterations may be passed Cytosines can be actively demethylated by TET (ten eleven transloca-
from generation to generation first requires an understanding of how tion) enzymes, which are responsible for the addition of a hydroxyl
such marks are maintained at the cellular level during DNA replication group to 5-methylcytosines, forming 5hmC (Tahiliani et al., 2009; Ito
and cellular division. Therefore, the cellular transmission of epigenetic et al., 2010). This is followed by a series of oxidation steps resulting in
modifications will be discussed in the following section, followed by a several intermediates including 5fC (5-formylcytosine) and 5CaC (5-
description of the current evidence in support of transgenerational carboxylcytosine), prior to base excision and repair with an unmodified
epigenetic inheritance and its potential underlying mechanisms. cytosine (He et al., 2011; Ito et al., 2011). It remains to be determined
whether these intermediates represent passive transitional stages on the
3. Putative cellular mechanisms for epigenetic inheritance road to demethylation, or if each modification serves its own distinct
function. Passive demethylation occurs when DNMT1 activity is not
Over the past several decades, much research has focused on the present or inhibited and 5mC marks are not able to be maintained
transmission of phenotypic traits or disease states via non-genetic during the replication process, leading to gradual dilution of the me-
modes of inheritance, and consequently, much debate has ensued thylation profile over time (Rougier et al., 1998; Lee et al., 2014). Both

3
J.J. Tuscher and J.J. Day Neurobiology of Disease 132 (2019) 104591

Fig. 2. Methylation levels during epigenetic reprogramming and development. The initial wave of mammalian methylome reprogramming occurs after fertilization
in the pronuclei of the paternal and maternal genomes via distinct demethylation mechanisms. Paternal pronuclei undergo active demethylation mediated via TET
enzymes, whereas maternal pronuclei are passively demethylated. Global methylation levels remain low in the zygote and blastocyst stage, until widespread re-
establishment of methylation by de novo DNMTs 3a and 3b occurs in the embryo after implantation, during the differentiation process. Once methylation programs
are established in differentiated cells of the developing embryo, cell-type specific patterns will be maintained in somatic cells during maturation and throughout the
lifespan by DNMT1. Methylome reprogramming also occurs in the primordial germline cells of the developing embryo. Unlike the distinct mechanisms experienced
by the paternal and maternal pronuclei shortly after fertilization, both active and passive modes of demethylation participate in primordial germ cell reprogramming.

active and passive demethylation mechanisms contribute to widespread Rougier et al., 1998). These periods of widespread demethylation fol-
post-fertilization demethylation in the developing zygote, and thus in- lowed by re-methylation in the post-implantation embryo create a
fluence the propagation of methylation patterns inherited from the substantial obstacle for the inheritance of any methylation marks that
maternal and paternal genomes. Each type of demethylation process may have been established in the parental epigenome.
will be discussed in greater detail in the following section. While this sequence of methylation events appears to be the general
rule of thumb, notable examples of persistent DNA methylation after
3.2. Multigenerational epigenetic inheritance via DNA methylation post-fertilization reprogramming have been documented. For example,
in mice, DNA methylation persists at specific differentially methylated
Given its relatively stable perpetuation during cell division, DNA regions (DMRs) in oocytes during post-fertilization periods that are
methylation has long been considered a putative route through which typically characterized by widespread hypomethylation (Smith et al.,
epigenetic patterns could be transferred from one generation to an- 2012). Certain DMRs also remain partially methylated during the pre-
other. However, the methylome undergoes heavy reprogramming at implantation stage in human oocytes (Guo et al., 2014b), indicating a
least twice during development in mammals (Lee et al., 2014), making possible route for epigenetic inheritance via maternally transmitted
many skeptical about the likelihood of transgenerational epigenetic methylation marks at these sites. Retrotransposable elements known as
inheritance via DNA methylation. The initial wave of mammalian me- intracisternal A particles (IAPs) may also be resistant to DNA de-
thylome reprogramming occurs in the pronuclei of the paternal and methylation during embryonic development, as some of these loci re-
maternal genomes via distinct demethylation mechanisms (Fig. 2; main mostly hypermethylated while primordial germ cells undergo
Rougier et al., 1998; Mayer et al., 2000; Oswald et al., 2000; Santos widespread demethylation (Lane et al., 2003; Guibert et al., 2012;
et al., 2002; Iqbal et al., 2011; Smith et al., 2012). Active demethylation Seisenberger et al., 2012). Persistent DNA methylation within gene
in paternal gametes is mediated via TET enzymes, which are re- bodies at intragenic enhancers has also been reported at the pre-im-
sponsible for converting 5mC to an unmethylated state through a series plantation stage in humans. Given that intragenic enhancer elements
of successive oxidation steps (Morgan et al., 2005; Guo et al., 2014a). are also subject to control by DNA methylation state, this represents a
Concurrently, the maternally inherited genome undergoes passive de- potential platform for methylation-based epigenetic inheritance of
methylation prior to implantation, wherein methyl marks fail to be certain cis-regulatory states (Birnbaum et al., 2012; Blattler et al.,
maintained during replication, and are thus gradually diluted in the 2014).
zygote (Mayer et al., 2000; Guo et al., 2014a). Global methylation le- Methylome reprogramming also occurs in the primordial germline
vels remain low in the blastocyst stage, until widespread re-establish- cells of the developing embryo (Fig. 2; (Smith et al., 2012). Unlike the
ment of methylation by de novo DNMTs 3a and 3b occurs in the embryo distinct mechanisms experienced by the paternal and maternal pronu-
after implantation (Fig. 2; (Santos et al., 2002; Smith et al., 2012). This clei shortly after fertilization, both active and passive modes of de-
re-establishment of methylation patterns by DNMT3a and 3b governs methylation participate in primordial germ cell reprogramming. More
cell fate and function during the differentiation process. Once methy- specifically, primordial germline demethylation is first exerted through
lation programs are established in differentiated cells of the developing active TET-mediated demethylation, followed by subsequent replica-
embryo, cell-type specific patterns will be maintained in somatic cells tion-dependent passive demethylation (Guibert et al., 2012;
during maturation and throughout the lifespan by DNMT1 (Fig. 2; Seisenberger et al., 2012; Hackett et al., 2013; Hill et al., 2018).

4
J.J. Tuscher and J.J. Day Neurobiology of Disease 132 (2019) 104591

However, similar to the post-implantation embryo, not all loci appear to modifications would allow for semi-conservative propagation of histone
be subject to methylation reprogramming. For example, some im- modification patterns. However, it is unclear how parent histone marks
printed genes and retrotransposons are also able escape demethylation are established on daughter histones that have been recently in-
in the primordial germline (Hackett et al., 2013). Although imprint corporated into nucleosomes of the newly synthesized DNA strand. This
control regions typically undergo the initial wave of demethylation may be facilitated in part by the recruitment of histone modifying en-
during germline development (Brandeis et al., 1993; Hajkova et al., zymes to the replication fork during DNA synthesis, or histone mod-
2002; Guo et al., 2014a), these sites can escape methyl mark erasure ifying enzymes may be able to interact directly with DNA in a sequence-
during the second period of demethylation (Kafri et al., 1993; Oswald specific manner, allowing the enzyme to remain associated with the
et al., 2000; Sakashita et al., 2014), providing a putative route for DNA during replication (Budhavarapu et al., 2013). In support of this
methylation patterns to be transmitted from parent to offspring. Despite notion, the histone modifying enzyme methyltransferase G9a, which
periods of largescale demethylation during embryonic development, interacts with DNMT1 to coordinate H3K9 dimethylation, localizes to
over 115,000 regions remain hypermethylated during methylation re- the replication fork during DNA synthesis (Esteve et al., 2006). Simi-
programming of the primordial germline in human development, and larly, PRC2 (Polycomb Repressive Complex 2), which methylates un-
many of these same regions similarly escape demethylation in mice modified H3K27 sites when recruited by previously established tri-
(Tang et al., 2015). Interestingly, a number of genes that escape me- methylated H3K27 marks, is also found near replication foci (Hansen
thylation reprogramming are expressed in the brain, and are linked to et al., 2008; Margueron et al., 2009). These examples suggest certain
the regulation of neurobiological processes and metabolism (Tang et al., histone modifying enzymes are physically recruited to replication sites,
2015). These gene targets are prime candidates for future investigation and may participate in the reinstatement of post-translational mod-
of how the transmission of their corresponding epigenetic marks may ifications on daughter histones. Thus, although many details regarding
set the stage for transgenerational inheritance of metabolic disease how histone PTMs are maintained during replication remain to be es-
states or neurological disorders. tablished, this evidence suggests that certain histone modifying en-
Collectively, these findings suggest at least some fraction of genomic zymes can be recruited to the site of DNA replication and may interact
loci (e.g., particularly imprint control regions, DMRs, IAPs, transposons, with other epigenetic modifiers (such as DNMTs) to maintain estab-
and some enhancer elements), may be more resistant to the global lished epigenetic patterns during DNA replication and across cell divi-
methylation reprogramming that occurs during early development. sion.
Why and how these particular regions of the genome are able to escape
methylation reprogramming remains largely unknown. Nevertheless, 3.4. Multigenerational epigenetic inheritance via histone modifications
such sites provide a putative route for DNA methylation patterns to be
conserved and transferred from the parent genome to future genera- Similar to global DNA methylation reprogramming, histones and
tions. their associated modifications experience chromatin remodeling events
that prove challenging for maintaining and transferring epigenetic
3.3. Cellular transmission of epigenetic marks via post-translational histone marks from one generation to the next. For example, during mamma-
modifications lian spermatogenesis, most histones are evacuated from nucleosomes
and replaced by small arginine-rich nuclear proteins known as prota-
Histone modifications also alter availability of DNA to the necessary mines in sperm to facilitate the compaction of paternal DNA (Balhorn,
transcriptional machinery required for regulating gene expression. 2007; Casas and Vavouri, 2014). This process serves to protect and
Histones are modified at the N-terminal portion of their tails that ex- stabilize the DNA, but also results in the removal of histones and any
tends beyond the nucleosome to interact with regulatory proteins, corresponding post-translational modification to histone tails. Although
neighboring histones, and the DNA itself (Marmorstein, 2001; Tsankova this reduces the likelihood of epigenetic transmission of histone mod-
et al., 2007). A variety of post-translational modifications (PTMs) to ifications, recent studies suggest protamines are also post-translation-
histone tails have been observed in the literature, including methyla- ally modified, and therefore may also be capable of transferring epi-
tion, phosphorylation, ubiquitination, SUMOylation, and acetylation, genetic information from one generation to the next (Brunner et al.,
which is perhaps the most well characterized. As with DNA methyla- 2014). Whether such modifications can be epigenetically inherited
tion, relatively stable transmission of any of these epigenetic marks across generations remains to be investigated. Despite the apparent
during DNA replication and cell division would be required for histone hurdle of widespread histone removal during sperm maturation, a small
alterations to be heritable. However, the brief lifespan of most histone percentage of histone modifications are preserved in humans and mice
modifications, and limited knowledge regarding the fidelity of histone (Hammoud et al., 2009; Brykczynska et al., 2010). This has been ob-
modifying enzymes, has dampened enthusiasm for histone PTMs as a served with both active (H3K4me2, H3K4me3) and repressive
plausible route for transgenerational epigenetic inheritance. For ex- (H3K27me3, H3K9me3, H4K20me3) histone marks, however specific
ample, histone lysine methylation, a relatively stable histone mod- PTMs and the loci at which they occur may differ by species. In humans,
ification, can be maintained on a timescale of hours-to-days (Zee et al., evidence suggests repressive histone marks H3K9me3 and H4K20me3
2010). Histone acetylation and phosphorylation states can fluctuate can be transferred from sperm to zygote, and these alterations are ac-
within minutes (Jackson et al., 1975; Chestier and Yaniv, 1979). Thus, companied by highly condensed constitutive heterochromatin (van de
the transient half-life of most histone modifications calls into question Werken et al., 2014). Some retention of histone marks has also been
the likelihood that such marks could be easily propagated during re- reported near the promoter region of key developmental and house-
plication and cell division. For modifications that do survive long en- keeping genes (Erkek et al., 2013). These findings suggest epigenetic
ough to be transmitted during replication, the mechanisms through transfer of histone modifications may occur via the paternal epigenome,
which parent histone marks are established on daughter histones re- although the opportunity for such contribution is limited given the
main poorly understood. Nonetheless, recent work in yeast has de- small percentage of histone PTMs preserved in sperm. Alternatively,
monstrated that certain histone modifications, such as H3K9 methyla- paternal chromatin compaction in the developing zygote may be driven
tion, can be stably propagated during mitosis and meiosis in a manner by maternal donation of the polycomb repressive complex (Puschendorf
independent of DNA methylation (Ragunathan et al., 2015). One pu- et al., 2008). This putative mechanism suggests heritable changes in the
tative mechanism involves the reassembly of two displaced parent paternal epigenome can also be regulated by maternal contribution of
histones into the nucleosome of the newly synthesized DNA strand, repressive complexes that shape the chromatin landscape during em-
along with two new daughter histones (Annunziato, 2005). This partial bryogenesis.
recycling of histones and their corresponding post-translational In oocytes, histones are largely retained during development (Gu

5
J.J. Tuscher and J.J. Day Neurobiology of Disease 132 (2019) 104591

et al., 2010). Thus, stable transmission of histone marks from one across generations in humans, several lines of research suggest that
generation to another maybe be more plausible via a maternal route. In epigenetic factors may also confer risk for the development of certain
support of this notion, recent examination of the maternal epigenome metabolic conditions. Exposure to high fat or low protein diets in the
has uncovered patterns of H3K4me3 and H3K27me3 marks in mouse parent generation, as well as intrauterine exposure to maternal caloric
oocytes that appear to impact embryonic development (Dahl et al., restriction, have all been linked to disruption of metabolic function in
2016; Zhang et al., 2016; Zheng et al., 2016; Inoue et al., 2017; Xu and offspring, in some cases beyond the F2 generation. To date, the vast
Xie, 2018). For example, a reduction in active histone methylation majority of multigenerational studies have focused on the inheritance
mark H3K4me2 at the promoter regions of genes that shape develop- of metabolic phenotypes in offspring, with little characterization of
mental processes and metabolic function have been linked to altered heritable epigenetic changes to somatic cells in the brain regions that
gene expression in F1 generation embryos. This modification induced regulate energy homeostasis and satiety. Nonetheless, some evidence
by lysine specific demethylase 1 (LSD1) overexpression was associated does suggest metabolic health of future progeny can be impacted by
with developmental defects that were present in F1 offspring and could poor nutrition in the parent generation. Specific examples of how high
be transferred across 3 generations (Siklenka et al., 2015). Although fat or low protein diet in the parent generation can influence metabolic
limited at present, these studies suggest that both active and repressive function of offspring will be discussed in the following sections.
histone marks may be perpetuated across generations in mammals, and
that these transferable epigenetic marks may yield important implica- 4.2. High fat diet
tions for embryogenesis and multigenerational inheritance.
Parental exposure to a diet high in fat content has been linked to
3.5. Cellular transmission of epigenetic marks by non-coding RNAs phenotypic transmission of impaired metabolic function in offspring. A
recent study involving the in vitro transfer of oocytes and sperm from
Small non-coding RNAs may also act as carriers of heritable epige- parents both fed a high fat diet to a healthy foster mother reported
netic information. These small RNAs are comprised of a variety of RNA significant weight gain in F1 offspring relative to offspring from mice
species under 200 nucleotides in length, and include micro RNAs fed a normal diet (Huypens et al., 2016). Specifically, female offspring
(miRNAs), PIWI-interacting RNAs (piRNAs), and transfer RNA frag- from obese parents gained the most weight, and this pattern of weight
ments (tRFs). In comparison to DNA methylation and histone mod- gain in the offspring persisted regardless of whether the dam, sire, or
ifications, relatively little is known about the role of small non-coding both parents received the high fat diet. Although potential epigenetic
RNAs in germline-dependent epigenetic transfer across generations. modifications accompanying the observed phenotypes were not eval-
Nonetheless, a variety of small non-coding RNA species have been uated in this study, the authors did report that obesity in female off-
documented in sperm and oocytes, which may be transferred to the spring was accompanied by insulin resistance (Huypens et al., 2016),
zygote at fertilization (Rassoulzadegan et al., 2006; Pauli et al., 2011; suggesting future studies should investigate the epigenetic landscape
Liu et al., 2012; Chen et al., 2016; Cropley et al., 2016; Sharma et al., surrounding genes important for insulin regulation. Pregnant female
2016; Yang et al., 2016). Although the precise mechanisms involved mice exposed to a fatty diet resulted in decreased insulin sensitivity and
remain to be defined, a host of exogenous factors such as exposure to increased body size in F1 and F2 offspring, an observation authors at-
pathogens, stress, elevated temperature, and dietary composition have tributed to both maternal and paternal lineage despite gestational ex-
been reported to produce alterations in RNA species found in germline posure (Dunn and Bale, 2011). Interestingly, in the F3 generation, in-
cells (Rassoulzadegan et al., 2006; Rechavi et al., 2011; Gapp et al., creased body size was only observed in female offspring, and appeared
2014; Rodgers et al., 2015; Chen et al., 2016; Cropley et al., 2016; to occur by a paternal mode of transmission. Although epigenetic
Sharma et al., 2016). These alterations may contribute to inheritance of modifications were not directly measured, the increased body weight
certain metabolic phenotypes linked to high fat diet, which will be phenotype was accompanied by a disrupted pattern of expression in
discussed further in the following section. Although limited, emerging paternally imprinted genes involved in growth regulation in liver tissue
evidence suggests tRNA fragments, piRNAs, and miRNAs may also collected from F3 female offspring (Dunn and Bale, 2011). Taken to-
contribute to epigenetic silencing at certain transcriptional domains. gether, the above studies suggest both maternal and paternal modes of
For example, miRNA inheritance has been linked to altered embryonic transmission may contribute to inheritance of altered metabolic phe-
development (Tang et al., 2007; Liu et al., 2012), and transmission of notypes. The fact that increased body size and altered gene expression
stress-induced behavioral phenotypes (Gapp et al., 2014; Rodgers et al., patterns persisted in the F3 generation suggest transgenerational me-
2015). Finally, research focused on tRNA fragments has begun to chanisms of inheritance may be at play, although more definitive
identify a role for these non-coding RNAs in the intergenerational characterization of accompanying epigenetic alterations to somatic and
transfer of metabolic phenotypes (Chen et al., 2016; Cropley et al., germline cells for each generation will be important for substantiating
2016; Sharma et al., 2016). In sum, like DNA methylation and histone this claim.
modifications, non-coding RNAs also have the potential to convey Maternal transmission of epigenetic inheritance can be particularly
epigenetic information from parental germline to their offspring and challenging to study, as diet-induced phenotypic traits in offspring
future generations. could arise from nuclear or mitochondrial DNA contributions, effects of
maternal diet and metabolism during gestation (e.g., diabetes, obesity),
4. Effect of epigenetic inheritance on phenotypic traits and brain or epigenetic modifications to female germ cells. One benefit of fo-
function cusing solely on paternal transmission is that sperm carries only genetic
and epigenetic material, eliminating some of the potential confounds
4.1. Nutritional intake and inheritance of metabolic phenotypes inherent to studies of this nature. Studies investigating exclusively pa-
ternal transmission after exposure to a high fat diet have observed
The prevalence of metabolic disorders is on the rise in the United weight gain, hypertension, and glucose intolerance in both F1 and F2
States and worldwide (Finucane et al., 2011; Ogden et al., 2014), generations (Masuyama et al., 2016). Further, these metabolic impair-
making the elucidation of cellular and molecular contributions to these ments were accompanied by H3K9ac and H4K20me1 in the promoter
diseases particularly relevant. Mendelian inheritance does not ade- regions of adiponectin and leptin genes in adipose tissue from adult F1
quately account for the apparent heritability of observed phenotypic offspring, genes involved in insulin resistance (Fasshauer and Paschke,
traits and metabolic disorders. Although environmental exposure to 2003) and satiety and energy homeostasis (Myers Jr., 2004), respec-
things like shared diet and other sociological and cultural factors most tively. Similar work examining paternal transmission of phenotypic
certainly contribute to the passage of impaired metabolic function metabolic traits found that sperm from obese male mice exhibited

6
J.J. Tuscher and J.J. Day Neurobiology of Disease 132 (2019) 104591

altered DNA methylation patterns and corresponding changes in insulin increased anxiety and depression-like behavior, as well as elevated
signaling genes, effects also observed in the pancreatic islets of F1 and plasma levels of the stress hormone corticosterone (Dietz et al., 2011).
F2 offspring (Wei et al., 2014). In addition to DNA methylation and However, most of these traits could not be recapitulated with in vitro
histone PTMs, small non-coding RNAs have also been implicated in the fertilization using sperm from defeated sires, suggesting these effects
epigenetic transfer of metabolic phenotypes. For example, injection of are likely not due to transmission of epigenetic marks across the
total RNA extracted from the sperm of mice on a high fat diet into germline. However, other studies involving chronic paternal stress re-
zygotes yielded impaired metabolic function in F1 offspring (Zhang sulted in hypoactivation of the hypothalamic-pituitary-adrenal (HPA)
et al., 2018). This effect was not observed when sperm was donated axis in offspring, and alterations in glucocorticoid-responsive genes in
from obese mice with genetic deletion of Dnmt2, a methyltransferase the bed nucleus of stria terminalis and paraventricular nucleus (Rodgers
that alters the stability of tRNAs. Such findings suggest that high fat- et al., 2013). Changes in gene expression were also accompanied by an
diet induced changes in metabolic function may be transferred to the F1 upregulation in 9 different miRNAs in the sperm of stressed vs. non-
generation via paternal donation of non-coding RNAs in sperm. stressed sires. Given that miRNAs in sperm can impact post-fertilization
Nevertheless, it currently remains unclear whether high fat diet-in- gene expression during early development (Rassoulzadegan et al.,
duced epigenetic alterations also occur in brain tissue, and if such 2006), and that DNMT3a (a key regulator of imprinted gene methyla-
marks are heritable across multiple generations. tion) was identified as one of the top predicted miRNA targets, these
findings provide a potential mode of action through which paternal
4.3. Caloric restriction and low protein diets stress can epigenetically alter germ cells. However, whether changes to
stress responsivity and the expression of stress-associated genes extends
Caloric restriction has also been shown to alter the epigenome in a beyond the F1 generation remains an open question.
heritable way. Male offspring of calorically restricted dams have low Paternal transmission of traumatic experience has also been in-
birth weights and suffer from several metabolic defects (Radford et al., vestigated using fear conditioning paired with specific scent cues.
2012). When dietary restrictions were in place during primordial Olfactory receptors are selectively expressed by sensory neurons of the
germline reprogramming, this led to DNA hypomethylation at 111 re- olfactory epithelium, with each individual sensory neuron expressing
gions in F1 sperm (Radford et al., 2012). However, changes in methy- only one of the hundreds (in humans) or thousands (in mice) of ge-
lation patterns at specific loci in F1 sperm were not observable in brain netically encoded olfactory receptors (Godfrey et al., 2004; Monahan
or liver cells of F2 embryos, suggesting this particular epigenetic effect et al., 2017). Given the relatively defined nature of olfactory receptor
is only intergenerational in nature. As with high fat diet studies, small activation by distinct odor stimuli, this system presents an opportunity
non-coding RNA species such as miRNAs and tRNA fragments are likely for elegant investigations into transgenerational epigenetic mechan-
also involved in the transmission of phenotypic traits related to meta- isms. In a landmark study by Brian Dias and Kerry Ressler, mouse sires
bolism. For example, low protein diets have been linked to altered tRNA were fear conditioned in the presence of a specific odor cue (acet-
fragment abundance, and this effect can be paternally transmitted via ophenone), which is known to activate the M71 odor receptor encoded
sperm to alter embryogenesis (Chen et al., 2016; Cropley et al., 2016; by the Olfr151 gene (Dias and Ressler, 2014). Surprisingly, this learned
Sharma et al., 2016). Other work examining the role of non-coding association was transferred to F1 and F2 progeny, measured as an
RNAs in epigenetic transmission of metabolic traits reported elevated elevated “startle” response to the conditioned odor but not to another
cholesterol synthesis in the liver of IVF-derived offspring from sires on a control odor. These inherited fear responses were accompanied by al-
low protein diet (Sharma et al., 2016). Phenotypic changes in metabolic tered promoter methylation of Olfr151 in F0 sperm, and neuroanato-
function were accompanied by dysregulation of hepatic cholesterol mical differences in the olfactory system of F1 and F2 offspring, in-
biosynthesis genes in the F1 generation. These effects appear to be cluding more M71 receptor expressing olfactory sensory neurons and
mediated in part by diet-induced changes in the abundance of tRNA- larger M71 glomeruli in the olfactory bulb (Dias and Ressler, 2014;
Gly-GCC fragments in F0 sperm, which repress genes normally regu- Fig. 3A). Interestingly, in vitro fertilization (IVF) and cross-fostering
lated by the endogenous retroelement MERVL (Mouse endogenous approaches were incorporated in this study to substantiate the idea that
retrovirus with leucine tRNA primer). Collectively, such work suggests observed phenotypic behavioral responses and molecular changes were
low protein diet-induced alterations in non-coding RNAs can be pa- not likely due to social transmission from the parent generation. Per-
ternally transmitted to subsequent generations via germ cells. At pre- sistence of behavioral sensitivity to the conditioned scent, and neu-
sent, these effects appear to be mostly intergenerational in nature. roanatomical alterations in the olfactory bulb of IVF-derived F2 pro-
Notably, non-coding RNA species may prove an attractive candidate for geny, as well as hypomethylation near the odorant receptor gene
future research on transgenerational epigenetic inheritance of meta- Olfr151 in sperm from the F0 generation and F1 odor naïve males,
bolic traits, as they may not be subject to the same type of epigenetic provides compelling evidence for epigenetic transmission of these
reprogramming events experienced by DNA methylation and histone phenotypic traits. Although it is unclear at present how olfactory cues
modifications during early development. and corresponding alterations in somatic cells of the olfactory system
Taken together, the above studies provide evidence for inter- led to methylation changes in sperm of the F0 and F1 generation, this
generational inheritance of metabolic phenotypes that occur as con- work provides intriguing preliminary evidence that information re-
sequence of diet-induced changes in the F0 generation. Present findings garding stressful signals encountered in the environment may be in-
suggest intergenerational alterations in metabolic function are medi- corporated into the germline and passed on to offspring.
ated in part by non-coding RNA species transferred to offspring via Maternal stress during pregnancy leads to elevated production of
paternal germline cells. However, whether diet-induced epigenetic al- stress hormones and this heightened activation of the HPA axis has been
terations observed in germ cells are also present in somatic cells of the linked to increased stress sensitivity, cognitive deficits, and dysregula-
CNS (i.e., in the brain), and the heritability of such marks across mul- tion of stress-related gene expression in offspring (Mueller and Bale,
tiple generations, remain open questions. 2007, 2008). When maternal stress is experienced during early preg-
nancy (i.e., within the first week), male, but not female, offspring dis-
5. Stress exposure and inheritance of altered stress responsivity play worse performances in tail suspension and forced swim tests, in-
dicative of depression-like behavior (Mueller and Bale, 2007, 2008).
Activation of the endocrine system by maternal or paternal exposure Early gestation stress also resulted in elevated production of the stress
to stressful events can alter gene expression and have long-term beha- hormone corticosterone, increased corticotropin-releasing factor in the
vioral consequences in offspring. Male mice that experienced chronic amygdala, and decreased glucocorticoid receptor expression in the
social defeat or variable stress prior to breeding yielded progeny with hippocampus of male offspring (Mueller and Bale, 2008). Further,

7
J.J. Tuscher and J.J. Day Neurobiology of Disease 132 (2019) 104591

Fig. 3. Putative mechanisms for transgenerational epigenetic inheritance of neuronal and behavioral phenotypes. A, Multigenerational epigenetic transmission of a
startle response after paternal olfactory fear conditioning. Fear conditioning in the presence of acetophenone odorant leads to reduced methylation of the Olfr151 gene in
the olfactory bulb of the paternal F0 generation (left) and in sperm (F1; right). Hypomethylation of Olfr151 (a gene that encodes M71 odor receptors) is associated
with an increase in the number of olfactory sensory neurons expressing M71 receptors, and larger corresponding glomeruli in the olfactory bulb of the F0 generation
(left). More M71-expressing olfactory sensory neurons and larger M71 glomeruli are also observed in adult F1 and F2 progeny without any direct exposure to the
acetophenone odor cue. Neuroanatomical adaptations observed in the F1/F2 generation are thought to increase sensitivity to acetophenone when offspring are later
exposed to this specific odorant, resulting in elevated startle to the odor cue in the absence of direct pairing with fear conditioning. B, Multigenerational epigenetic
transmission of a cocaine resistant phenotype after paternal cocaine exposure. Cocaine self-administration leads to increased H3K9K14ac2 at Bdnf promoter IV and
elevated BDNF protein and mRNA transcripts in somatic cells of the mPFC in the F0 generation (left), as well as increased H3K9K14ac2 at all Bdnf promoters in F0
sperm (F1 generation; right). During adulthood, parallel changes in BDNF protein, mRNA, and H3 acetylation are observed in the mPFC of F1 offspring. Elevated
BDNF levels in the mPFC have previously been linked to a reduction in the reinforcing effects of cocaine, and subsequently, a decline in drug-seeking behavior. Thus,
elevated Bdnf mRNA and protein levels in the mPFC of F1 progeny prior to drug exposure are thought to confer a protective phenotype by blunting the rewarding
properties of cocaine and reducing the propensity for drug seeking.

8
J.J. Tuscher and J.J. Day Neurobiology of Disease 132 (2019) 104591

methylation of the glucocorticoid receptor promoter was also elevated sensitization (Crozatier et al., 2003), and impaired synapse maturation
in hippocampal tissue, suggesting epigenetic mechanisms are likely during development (Bellone et al., 2011). Although such work pro-
involved, although most likely as a result of direct exposure to stress vides valuable translational information regarding the deleterious ef-
hormones in utero, and not through germline inheritance of such fects of cocaine exposure during pregnancy, numerous studies focused
modifications. In guinea pigs, maternal stress induced by repeated ex- on epigenetic transmission of behavioral traits in the context of ad-
posure to flashing light similarly increased anxiety-related behaviors diction have opted to focus on paternal transmission, as this circum-
and basal corticosterone levels in offspring while reducing glucocorti- vents potential interpretational confounds that can arise in transge-
coid receptor expression in the hippocampus when the stressor oc- nerational inheritance studies. For example, past drug experience in
curred during mid, but not late, pregnancy (Kapoor and Matthews, female rats can negatively impact maternal care for pups later in life
2005; Kapoor et al., 2009). Taken together, these studies suggest timing (Johnson et al., 2011), and a single instance of maternal separation has
of perturbation during pregnancy may play a key role in the extent to been shown to influence drug intake in pups (Martini and Valverde,
which developmental processes will be disrupted or induce long-term 2012). Thus, investigation of paternal drug use on neurobiological
epigenetic changes in the neural tissues of their progeny. At any rate, processes and behavioral phenotypes in offspring from drug naïve dams
the vast majority of studies examining the effects of maternal trauma on evades these potential confounds. This approach also rules out the
stress vulnerability in offspring to date involve direct in utero exposure potential for transfer of epigenetic marks by direct drug exposure to the
to the F1 (developing embryo), and F2 (embryo primordial germline) F1 or primordial germline of F2 generations in utero.
generations, and as such are intergenerational in nature. Additional At present, a majority of the evidence for transmission of phenotypic
research will be required to determine whether epigenetic modifica- traits across generations as a consequence of paternal or maternal drug
tions that occur as a consequence of maternal stress are transgenera- use is behavioral and correlational in nature. Nonetheless, a variety of
tional, extending out to the F3 offspring. behavioral phenotypes have been observed in offspring from cocaine
Another important consideration for research examining the con- experienced sires, including hyperlocomotion, impaired spatial
sequence of parental stress on offspring is that maternal stress can also memory, and elevated anxiety levels (Abel et al., 1989; Fischer et al.,
alter the level of maternal care dams provide to their pups after birth 2017). Some of the earliest work to report intergenerational transmis-
(Ivy et al., 2008), in addition to affecting the hormonal milieu of the sion of addiction-related phenotypic traits found that both male and
placenta. Poor maternal care during this early postnatal period is female adolescents (P16) from cocaine-sired offspring displayed hy-
traumatic in its own right (Weaver et al., 2004; Rice et al., 2008; Ivy perlocomotive activity (Abel et al., 1989). More recently, sex-specific
et al., 2010; Korosi and Baram, 2010; McClelland et al., 2011), and can effects on locomotion have been reported in male, but not female,
thus act as its own direct exposure to stress early in life, making in- offspring of cocaine-sired mice (Fischer et al., 2017). Anxiety-like be-
terpretation of transgenerational vs. intergenerational social transmis- haviors have also commonly been observed in the progeny of cocaine-
sion of such effects on behavior and gene expression challenging. For experienced rodents. For example, male offspring from cocaine-sired
example, stressed mothers provide poor maternal care, and the stress of rats reportedly display defensive burying and novelty-induced hypo-
poor maternal care acts as its own direct insult to the developing pup. phagia, which are considered measures of anxiety-like behavior (White
This elevates stress sensitivity in adult F1 progeny, likely affecting their et al., 2016). Similarly, male offspring of rats receiving chronic cocaine
own ability to provide adequate maternal care, thus perpetuating the injections also demonstrated increased anxiety levels relative to saline
cycle. An important mechanistic distinction is that this continued pas- controls, as measured by amount of time in the open arm of the ele-
sage of a stress sensitive phenotype and any accompanying epigenetic vated plus maze (Fischer et al., 2017).
alterations likely occurs from direct exposure to a stressful postnatal Despite these intriguing observations, it should be noted that at least
experience in each generation, making this a repeating cycle of inter- one study measuring locomotion was not able to recapitulate a similar
generational effects, rather than transgenerational effects, which re- hyperactive phenotype in adult mouse progeny (P60) sired by cocaine-
quire transfer of epigenetic modifications via the germline in the ab- experienced fathers (Killinger et al., 2012). Similarly, this same study
sence of direct exposure. To avoid these type of experimental also observed no impact on paternally transmitted anxiety-like traits
confounds, some investigators have opted to focus on patrilineal in- after cocaine experience in the elevated plus maze or open field test
heritance, as male rodents are not typically involved in rearing off- (Killinger et al., 2012). One potential difference that could account for
spring in a laboratory setting. Exposure to paternal stress prior to these discrepancies is that sex-specific statistical analyses were used for
breeding is thought to minimize potential social or behavioral con- some studies (Fischer et al., 2017), while others collapsed sex during
founds that can be imposed during early postnatal periods. However, it analysis (Killinger et al., 2012), which would alter power for detecting
is worth noting that mate quality can impact maternal investment of behavioral effects within each sex. Differences in amount of handling,
female rodents in caring for their offspring (Drickamer et al., 2000; behavioral protocols, and total amount of cocaine exposure may have
Mashoodh et al., 2012). Therefore, even subtle interactions between the also contributed to these incongruent observations. Despite these dis-
sire and dam during or after mating could theoretically impact maternal crepancies, at least some research suggests that paternal exposure to
behavior and postnatal care of offspring, and thus it should not be as- cocaine can lead to elevated locomotor activity and other anxiety-like
sumed that separation of an exposed sire from the dam after mating will traits in F1 offspring, and that male offspring may be particularly sus-
guarantee avoidance of any potential behavioral confounds. ceptible to these behavioral abnormalities.
In addition to hyperactivity effects, paternal exposure to cocaine
6. Substance abuse and inheritance of behavioral phenotypes may also disrupt normal memory function in F1 progeny, as impair-
ments have been observed in the performance of certain memory tasks
Exposure to drugs of abuse can modify the epigenome and alter in cocaine-sired rodents. For example, adolescent (P35) offspring from
neural plasticity, and each of these phenomena is thought to establish cocaine-sired rats demonstrated perseverative behavior in a T-Maze
and maintain the addicted state (Renthal and Nestler, 2008; Heller task (Abel et al., 1989), and similar impairments were observed in
et al., 2014). The evidence for epigenetic modifications and changes in cocaine-sired CD1 mouse offspring in the 5-Arm Maze, a spatial
gene expression in reward circuitry of the drug-exposed individual is working memory task (He et al., 2006). In a self-administration para-
well established, but more recently epigenetic modifications have also digm, adult male offspring also exhibited impaired spatial memory in
been documented in the sperm and oocytes after drug exposure the object location task (Wimmer et al., 2017). Although most of this
(Vassoler and Sadri-Vakili, 2014). Prenatal (i.e., in utero) exposure to evidence suggests spatial memory tasks are particularly susceptible to
cocaine in rodent addiction models can result in elevated self-admin- disruption by paternal drug use, others have reported no impairment in
istration of cocaine (Keller Jr. et al., 1996), altered stimulant-induced the Morris Water Maze, another type of hippocampus-dependent spatial

9
J.J. Tuscher and J.J. Day Neurobiology of Disease 132 (2019) 104591

task (Killinger et al., 2012; Fischer et al., 2017). Nonetheless, the bal- protein levels (Vassoler et al., 2013; Collo et al., 2014). Interestingly,
ance of studies to date suggest tasks that recruit the hippocampus may Bdnf levels in the offspring of drug-exposed sires is reported to posi-
be more heavily influenced by the consequences of paternal exposure to tively correlate with amount of cocaine intake in the parent generation
cocaine. However, several factors may have contributed to the observed (Le et al., 2017). This aligns with previous research suggesting elevated
differences in memory task performance. For example, Killinger and mPFC BDNF levels coincide with reduced drug-seeking behavior, and
Fischer used experimenter administered cocaine, whereas Wimmer and may play a key role in mediating the protective effects some have ob-
He utilized self-administration models, allowing the subject to lever served in the offspring of cocaine-sired rodents (Berglind et al., 2007;
press for cocaine reward. Differences in these experimental approaches Vassoler et al., 2013;Fig. 3B). Elevated DNA methylation has also been
should therefore be considered when interpreting the data, as each observed in the sperm of highly motivated “addict” vs. non-addict co-
approach would likely engage distinct brain regions to a different ex- caine-exposed F0 rats, a pattern which persisted in their F1 descendants
tent. Self-administration relies on learning about contingencies between (Le et al., 2017). Consistent with these distinct methylation patterns,
behavioral responses, environmental cues, and the rewarding proper- transcriptome sequencing of the Nucleus Accumbens (NAc) in the F1
ties of the drug, whereas experimenter administration is more passive. generation revealed that reduced global methylation in F0 sperm co-
Further, self-administration studies likely resulted in more overall co- incided with significantly lower expression levels of genes important for
caine experience vs. the experimenter delivered studies, and this could developmental regulation, gap junction formation, and axon guidance
also be an important factor in paternal transmission of certain drug- in the addict-like phenotype vs. non-addict F1 progeny (Le et al., 2017).
induced phenotypic behaviors. Overall, these findings suggest highly motivated drug-seeking behavior
Perhaps not surprisingly, cocaine-sired offspring of both sexes dis- may alter DNA methylation in a way that is distinct from methylation
play behavioral differences in response to their own experience with that occurs as a result of drug exposure alone, and that these epigenetic
cocaine. One example of this is stimulant-induced hyperlocomotion, modifications may be transmitted to F1 offspring. Additional research
suggesting increased sensitivity to cocaine experience (Fischer et al., will be required to determine if heritability of the epigenetic patterns
2017). Cocaine exposure in the parent generation may shift the dose- observed in the F1 generation extends to F2 progeny.
response curve for female mice, which show reduced magnitude in their Collectively, this body of work provides a preliminary glimpse into
conditioned place preferences, and in response to a lower dose of co- the generational effects of cocaine exposure on anxiety, memory, sen-
caine than usual (5 mg/kg vs. 10 mg/kg; Fischer et al., 2017). Male sitivity to drugs of abuse, and corresponding alterations to the epi-
offspring appear to experience the effects of parental exposure to co- genome and gene expression. Although converging evidence does sug-
caine slightly differently. For example, cocaine-sired male rats acquire gest transmission of behavioral phenotypes, much work remains to be
self-administration more slowly, and took less total drug overall done to see if these observations extend beyond the F1 generation,
(Vassoler et al., 2013), suggesting a sex-specific protective effect against making them true transgenerational rather than intergenerational ef-
drug seeking in male progeny. fects. Further, characterization of which epigenetic marks are trans-
In a model that divided rats into groups based on motivation level to ferred from drug experienced parent to offspring, the brain regions and
voluntarily self-administer cocaine, researchers found that the highly cell types they occur in, how such modifications survive epigenetic
motivated behavioral phenotype (termed “Addict” rats in the study) reprogramming during development, and how such alterations would
could be transmitted to F1 and F2 generations (Le et al., 2017). Inter- manifest as the observed behavioral phenotype remain to be explained.
estingly, the heritable addict-like behavioral phenotype was linked to Thus, future studies will be important to further dissect the multi-
high motivation to drug seek, rather than higher levels of cocaine intake generational transmission of these phenotypic traits, and the epigenetic
alone, as transmission of this behavioral phenotype was not observed in remodeling events that may subserve them.
offspring when the F0 generation received passive experimenter-ad-
ministered cocaine injections (Le et al., 2017). This addict-like pheno- 7. Current limitations and challenges, and approaches to move
type was also not observed when the F0 generation sired offspring prior forward
to drug experience, suggesting the phenotype is acquired as a con-
sequence of contingent drug experience, and is not likely a result of Many gaps remain in our understanding of how epigenetic mod-
direct genetic inheritance. ifications acquired in the parent generation might contribute to the
Cocaine use is known to shape changes in gene expression and transmission of phenotypic traits in their offspring. Despite evidence
plasticity in the reward circuitry that underlies addiction-related be- that certain behaviors and disease states seemingly persist across gen-
haviors. However, currently evidence for which corresponding epige- erations, the idea that these traits are caused by germline-dependent
netic modifications are also observed in the brains of the F1 generation inheritance of epigenetic marks remains controversial. This is in part
after paternal drug experience, and the potential mechanisms through due to a number of issues that remain unresolved, including un-
which this might occur, are quite limited. One of the earliest studies to ambiguous proof of heritability, demonstration of survival of epigenetic
link cocaine-induced epigenetic alterations to heritable changes in be- marks during epigenome reprogramming, their selective maintenance
havior found a significant decrease in Dnmt1 and increased Dnmt3a in in specific somatic tissues, and the extent to which these inherited
the seminiferous tubules of the testes in male mice trained to self-ad- modifications are causal of an observable phenotype. The following
minister cocaine (He et al., 2006). Although epigenetic modifications section highlights current limitations and key issues in transgenera-
were not measured in adult offspring, several morphological and be- tional epigenetic inheritance research, and proposes guidelines for how
havioral abnormalities were noted, including decreased cerebral vo- future research might address these existing gaps.
lume, reduced head diameter, impaired attention and working memory.
This study was one of the first to show drug experience alters not only 7.1. Technical limitations and challenges in experimental design
the self-administering rodent, but also their germline cells, hinting at a
potential mechanism through which cocaine-induced changes to the Epigenetic factors help shape cell fate and function, and accord-
epigenome could be transferred to the next generation. A more recent ingly, methylation patterns likely differ between distinct subpopula-
study also observed epigenetic alterations to the sperm of cocaine-ex- tions of cells both within the brain and throughout the periphery. As
perienced sires, this time in the form of elevated H3 acetylation at Bdnf such, the tissue collected to generate bulk homogenate for most EWAS
promoters I, IV, and VI (Vassoler et al., 2013). These effects in the F0 studies likely differs in composition, with some cell types being over-
generation were accompanied by sex-specific increases in H3 acetyla- represented while others remain underrepresented (Jaffe and Irizarry,
tion at Bdnf promoter IV in the medial prefrontal cortex (mPFC) of 2014). Thus, observations made from methylation patterns consisting
cocaine-sired male offspring, as well as increased Bdnf mRNA and largely of one cell type may not generalize to other cell populations.

10
J.J. Tuscher and J.J. Day Neurobiology of Disease 132 (2019) 104591

Single-cell level resolution of methylation and histone profiles in future 7.4. Unambiguous demonstration of heritability
studies will be necessary to overcome this technical limitation, and
provide a more accurate reflection of the heterogeneity of epigenetic One challenge transgenerational epigenetic inheritance studies have
patterns across unique cell populations. yet to overcome is definitively demonstrating that epigenetic marks
In addition to the technical limitations of currently available ap- present in the germline of the parent generation are truly inherited by
proaches, careful attention must be paid to experimental design in the F1 generation, rather than removed and re-established in offspring.
order to reduce potential variability and contribution of non-epigenetic During early prenatal development, epigenetic marks present in the
factors to studies of this nature. For example, preclinical research parent germline (e.g., DNA methylation and histone PTMs) are largely
should consider using inbred rodent strains with tightly regulated en- removed in the developing zygote after fertilization. At present, it re-
vironments to reduce the potential for genetic and environmental mains unclear how specific epigenetic modifications in the maternal
confounds. Further, if transgenerational effects are to be evaluated, and paternal gametes would survive this largescale removal of epige-
studies must be extended long enough to allow for evaluation of epi- netic marks. Further, it is not clear how epigenetic modifications that
genetic modifications and phenotypic traits in the appropriate genera- do survive this period of reprogramming would be transferred from
tion (i.e., F2 for male or non-pregnant female exposure, F3 for exposure undifferentiated embryonic cells to specific differentiated somatic tis-
to pregnant females; see Fig. 1). For some studies, an attractive alter- sues (i.e., neurons) in the adult offspring. Although DNMTs assist in
native may involve the use of in vitro fertilization and transfer to foster preserving methylation marks during cellular differentiation and in
mothers, which allows for strict observation of transmission of epige- somatic tissues, the accuracy of DNA methylation during replication is
netic marks via the gametes and reduces the possible influence of al- imperfect (~96% fidelity rate). Although this may sound relatively
tered behavioral effects in the parent generation (e.g., changes in ma- high, in reality it allows for errors to be introduced into methylation
ternal care) indirectly altering behavior or epigenetic changes in patterns across multiple rounds of DNA replication and cell division.
progeny. Therefore, accurate sustained transmission of such epigenetic mod-
ifications, which would be necessary for the transgenerational passage
of such marks, would be difficult to achieve. Even less is known about
7.2. Crosstalk between different epigenetic modifications how post-translational modifications of histones are maintained during
DNA replication and cell division, so this would also be an issue for
The choreography of various epigenetic factors that act in concert to studies focused on a histone-mediated mode of epigenetic inheritance.
initiate transcriptional programs and define cellular fate are complex Given the high probability for the loss of both histone PTMs and DNA
and interconnected. Combinatorial approaches will be required to more methylation modifications over multiple rounds of replication and cell
accurately define methylation state in conjunction with histone mod- division, future studies should focus on providing convincing evidence
ifications by using approaches that incorporate bisulfite sequencing that epigenetic marks observed in the parent germline are also observed
information with ChIP data from the same cellular populations at several other key developmental stages (gametes > zygote >
(Brinkman et al., 2012; Statham et al., 2012). Because epigenetic blastocyst > post-implantation embryo > F1 offspring) with parallel
modifications are dynamic and reversible, intermediate states of DNA evidence of the heritable mark across these same stages in the F2
methylation (e.g., 5fC and 5caC) may also need to be taken into con- generation. Bisulfite sequencing, methylated DNA immunoprecipitation
sideration in EWAS studies to capture a deeper level of resolution on (MeDIP), or chromatin immunoprecipitation (ChIP)-based approaches
these dynamic states. At present, it is unclear whether these are truly with antibodies against specific histone marks may be best suited to
intermediate states, or if they serve unique functions of their own. track histone PTM or DNA methylation profiles across generations.
Sequencing data (e.g., RNA-seq, ChIP-seq) demonstrating gene expres-
sion profiles and phenotypic traits expected to accompany the observed
7.3. Contribution of genetic influence on epigenetic landscape pattern of epigenetically inherited marks would further strengthen such
work, and provide a plausible proximal mechanism linking inheritance
In addition to the complex interplay amongst different epigenetic of these marks to biological functions in the affected cell population.
factors, there is also the potential for indirect genetic contribution to
the transmission of epigenetic marks. That is, normal variation in DNA
sequence between individuals may underlie some DNA methylation 7.5. Transfer of germline modifications to target somatic tissues
differences observed between individuals, making it particularly chal-
lenging to disentangle genetic vs. epigenetic transmission of epigenetic In order for epigenetic modifications laid down in the germline to
marks and corresponding effects on phenotypic traits (Gibbs et al., influence brain function, it is likely that these modifications would have
2010; Bell et al., 2011; Gertz et al., 2011). For this reason, epigenetic to survive through early development in order to alter gene expression
inheritance studies should include DNA sequencing controls whenever in target neurons that control the phenotype in question. However, it is
possible to ensure observed effects are not an indirect result of genetic presently unclear what determines which somatic tissues will maintain
rather than epigenetic differences between individuals (Lappalainen the epigenetic marks originating from undifferentiated stem cells. It
and Greally, 2017). seems unlikely that such modifications would be transferred from the
This consideration is also important for monitoring gene variants germline to all somatic tissues, so how are such marks passed on only to
that could lead to an indirect epimutation that impacts transcription specific cell populations of cells but not others? For example, if sub-
patterns in the sense or antisense orientation at the region of interest. stance abuse leads to an epigenetic alteration at gene X only in nucleus
This phenomenon was recently reported to accompany a rare disease accumbens neuron populations, how does this mark 1) get transferred
across several generations, wherein genetic mutation in one gene re- from parental neural tissue to the germline, and 2) from the germline
sulted in elimination of its usual transcription termination signal, and containing undifferentiated cells back to just a distinct subpopulation of
subsequently led to atypical methylation at a gene neighboring the neurons in a specific brain region (but not in other somatic tissues)? It is
region of interest (Chan et al., 2006; Ligtenberg et al., 2009; Gueant easier to imagine how this might occur for phenotypes that are encoded
et al., 2018). Although ultimately methylation marks were altered at by highly specific expression of specific genes in specific cell popula-
the disease-linked gene, this change was an indirect consequence of tions (e.g., olfactory receptors), but it is far more challenging to un-
direct mutation in the DNA sequence, rather than through gamete in- derstand how this selectivity could occur for more complex polygenic
heritance of the methylation mark, and as such is not truly re- or omnigenic phenotypes.
presentative of epigenetic inheritance.

11
J.J. Tuscher and J.J. Day Neurobiology of Disease 132 (2019) 104591

7.6. Evidence demonstrating the causal nature of epigenetic alterations about epigenetic mechanisms in somatic and dividing cells over the past
several decades, the mechanisms that drive heritable, germline-de-
In addition to the limited proof concerning unambiguous herit- pendent transmission of epigenetic marks remain poorly understood. At
ability, evidence demonstrating causality is also currently lacking. present, germline-dependent transmission appears to be the exception
Future studies should include rigorous evidence that epigenetic al- rather than the rule for mammalian species. However, while currently
terations passed from one generation to the next are the root cause of limited, these exceptions have provided dramatic evidence for the po-
the observed phenotype, rather than a co-occurrence. Recent advances tential implications for transgenerational epigenetic contributions to a
in epigenetic editing techniques will be particularly important for de- number of global public health issues, including obesity, substance use
monstrating both necessity and sufficiency of key epigenetic mod- disorders, and stress response in the face of traumatic experience.
ifications in the manifestation of observable phenotypic traits. For ex- Future research will be required not only for providing more conclusive
ample, CRISPR/dCas9-based strategies allow for epigenetic modifying evidence that epigenetic alterations acquired in one generation can be
enzymes, including DNMTs and HATs, to be tethered to a catalytically stably inherited and propagated to future generations, but also for
inactive dCas9 protein (Liu et al., 2016; Vojta et al., 2016; Kwon et al., unravelling the molecular mechanisms that make this type of in-
2017; Pflueger et al., 2018; Savell et al., 2019). This dCas9/epigenetic formation transfer possible.
modifier construct can then be directly targeted to specific loci
throughout the genome. These type of directed epigenetic alterations Acknowledgements
will allow for subsequent changes in gene expression, behavior, and
phenotypic traits to be observed in response to specific epigenetic This work was supported by NIH grants DA039650, DA034681, and
modifications, which can then be monitored across multiple genera- MH114990 (J.J.D.). Illustrations were generated using BioRender.
tions. Such work will be essential for providing more rigorous and
compelling evidence regarding the causal nature of transgenerational References
epigenetic inheritance.
Abel, E.L., Moore, C., Waselewsky, D., Zajac, C., Russell, L.D., 1989. Effects of cocaine
8. Conclusions hydrochloride on reproductive function and sexual behavior of male rats and on the
behavior of their offspring. J. Androl. 10, 17–27.
Annunziato, A.T., 2005. Split decision: what happens to nucleosomes during DNA re-
Although it has quickly captured the attention of the neuroscience plication? J. Biol. Chem. 280, 12065–12068.
community, research on multigenerational epigenetic inheritance in Anway, M.D., Cupp, A.S., Uzumcu, M., Skinner, M.K., 2005. Epigenetic transgenerational
actions of endocrine disruptors and male fertility. Science 308, 1466–1469.
mammalian species remains in its infancy. Despite evidence of inter- Baccarelli, A., Bollati, V., 2009. Epigenetics and environmental chemicals. Curr. Opin.
generational and transgenerational phenotypes in many different model Pediatr. 21, 243–251.
systems, a number of critical questions remain to be answered. Perhaps Bale, T.L., 2015. Epigenetic and transgenerational reprogramming of brain development.
Nat. Rev. Neurosci. 16, 332–344.
most importantly, it is currently unclear what biological problem this Balhorn, R., 2007. The protamine family of sperm nuclear proteins. Genome Biol. 8, 227.
form of inheritance solves. Here, it is worth considering whether gen- Bell, J.T., Pai, A.A., Pickrell, J.K., Gaffney, D.J., Pique-Regi, R., Degner, J.F., Gilad, Y.,
erational passage of epigenetic changes make sense from an evolu- Pritchard, J.K., 2011. DNA methylation patterns associate with genetic and gene
expression variation in HapMap cell lines. Genome Biol. 12, R10.
tionary biology perspective. For example, if an external stimulus (e.g.,
Bellone, C., Mameli, M., Luscher, C., 2011. In utero exposure to cocaine delays postnatal
drug use, stress, poor nutrition) results in overwhelming negative synaptic maturation of glutamatergic transmission in the VTA. Nat. Neurosci. 14,
consequences for the individual and their ability to create optimally fit 1439–1446.
progeny, why might such epigenetic modifications and consequent Berglind, W.J., See, R.E., Fuchs, R.A., Ghee, S.M., Whitfield Jr., T.W., Miller, S.W.,
McGinty, J.F., 2007. A BDNF infusion into the medial prefrontal cortex suppresses
phenotypes be allowed to persist for multiple generations? Why would cocaine seeking in rats. Eur. J. Neurosci. 26, 757–766.
the same malleable system that allowed for epigenetic marks to be Bierut, L.J., Dinwiddie, S.H., Begleiter, H., Crowe, R.R., Hesselbrock, V., Nurnberger Jr.,
deposited in the first place not respond in a manner more adaptive for J.I., Porjesz, B., Schuckit, M.A., Reich, T., 1998. Familial transmission of substance
dependence: alcohol, marijuana, cocaine, and habitual smoking: a report from the
the species? Although it is worth noting that some evidence in the ro- collaborative study on the genetics of alcoholism. Arch. Gen. Psychiatry 55, 982–988.
dent literature indicates drug experience may confer a protective phe- Bird, A., 1992. The essentials of DNA methylation. Cell 70, 5–8.
notype in offspring, this is not necessarily reflective of what happens in Birnbaum, R.Y., Clowney, E.J., Agamy, O., Kim, M.J., Zhao, J., Yamanaka, T., Pappalardo,
Z., Clarke, S.L., Wenger, A.M., Nguyen, L., Gurrieri, F., Everman, D.B., Schwartz, C.E.,
humans. If anything, parental drug experiences confer more – not less – Birk, O.S., Bejerano, G., Lomvardas, S., Ahituv, N., 2012. Coding exons function as
vulnerability for substance use disorders in offspring generations tissue-specific enhancers of nearby genes. Genome Res. 22, 1059–1068.
(Bierut et al., 1998; Goldman et al., 2005; Vink, 2016). This problem is Blattler, A., Yao, L., Witt, H., Guo, Y., Nicolet, C.M., Berman, B.P., Farnham, P.J., 2014.
Global loss of DNA methylation uncovers intronic enhancers in genes showing ex-
inherent to all epigenetics studies, as the nature of the system is
pression changes. Genome Biol. 15, 469.
somehow both stable (i.e., transmissible across generations) yet plastic Brandeis, M., Kafri, T., Ariel, M., Chaillet, J.R., McCarrey, J., Razin, A., Cedar, H., 1993.
in response to external stimuli. The ontogeny of allele-specific methylation associated with imprinted genes in the
mouse. EMBO J. 12, 3669–3677.
A second concept that is valuable to consider is the relative weight
Brinkman, A.B., Gu, H., Bartels, S.J., Zhang, Y., Matarese, F., Simmer, F., Marks, H., Bock,
of epigenetic changes in comparison to genetic influences. Genetic C., Gnirke, A., Meissner, A., Stunnenberg, H.G., 2012. Sequential ChIP-bisulfite se-
contributions to observable phenotypes in mammals are powerful, nu- quencing enables direct genome-scale investigation of chromatin and DNA methy-
merous, and extremely reproducible. Key examples include primary sex lation cross-talk. Genome Res. 22, 1128–1138.
Brunner, A.M., Nanni, P., Mansuy, I.M., 2014. Epigenetic marking of sperm by post-
determination by Y-linked genes, severe neurodevelopmental disorders translational modification of histones and protamines. Epigenetics Chromatin 7, 2.
arising from de novo mutations, and the highly penetrant nature of Brykczynska, U., Hisano, M., Erkek, S., Ramos, L., Oakeley, E.J., Roloff, T.C., Beisel, C.,
familial forms of neuropsychiatric or neurodegenerative disease states. Schubeler, D., Stadler, M.B., Peters, A.H., 2010. Repressive and active histone me-
thylation mark distinct promoters in human and mouse spermatozoa. Nat. Struct.
Are epigenetic modifications simply occurring on top of the large set of Mol. Biol. 17, 679–687.
robust genetic factors that contribute to development and adult func- Budhavarapu, V.N., Chavez, M., Tyler, J.K., 2013. How is epigenetic information main-
tions of the nervous system, providing only slight tweaks or temporary tained through DNA replication? Epigenetics Chromatin 6, 32.
Cadet, J.L., 2016. Epigenetics of stress, addiction, and resilience: therapeutic implica-
changes? Or are epigenetic modifications able to overturn or usurp tions. Mol. Neurobiol. 53, 545–560.
genetically defined programs, establishing new qualitatively different Casas, E., Vavouri, T., 2014. Sperm epigenomics: challenges and opportunities. Front.
transcriptional landscapes and behavioral phenotypes? Addressing this Genet. 5, 330.
Cavalli, G., Paro, R., 1998. The Drosophila Fab-7 chromosomal element conveys epige-
question will be essential to determining the position of epigenetic netic inheritance during mitosis and meiosis. Cell 93, 505–518.
mechanisms in the hierarchy of molecular mechanisms that contribute Cavalli, G., Paro, R., 1999. Epigenetic inheritance of active chromatin after removal of the
to brain development, function, and disease. main transactivator. Science 286, 955–958.
Chan, T.L., Yuen, S.T., Kong, C.K., Chan, Y.W., Chan, A.S., Ng, W.F., Tsui, W.Y., Lo, M.W.,
Although we have gleaned a considerable amount of information

12
J.J. Tuscher and J.J. Day Neurobiology of Disease 132 (2019) 104591

Tam, W.Y., Li, V.S., Leung, S.Y., 2006. Heritable germline epimutation of MSH2 in a genes. Nat. Rev. Genet. 6, 521–532.
family with hereditary nonpolyposis colorectal cancer. Nat. Genet. 38, 1178–1183. Gu, L., Wang, Q., Sun, Q.Y., 2010. Histone modifications during mammalian oocyte
Chen, Q., Yan, M., Cao, Z., Li, X., Zhang, Y., Shi, J., Feng, G.H., Peng, H., Zhang, X., maturation: dynamics, regulation and functions. Cell Cycle 9, 1942–1950.
Zhang, Y., Qian, J., Duan, E., Zhai, Q., Zhou, Q., 2016. Sperm tsRNAs contribute to Gueant, J.L., et al., 2018. APRDX1 mutant allele causes a MMACHC secondary epimu-
intergenerational inheritance of an acquired metabolic disorder. Science 351, tation in cblC patients. Nat. Commun. 9, 67.
397–400. Guibert, S., Forne, T., Weber, M., 2012. Global profiling of DNA methylation erasure in
Chestier, A., Yaniv, M., 1979. Rapid turnover of acetyl groups in the four core histones of mouse primordial germ cells. Genome Res. 22, 633–641.
simian virus 40 minichromosomes. Proc. Natl. Acad. Sci. U. S. A. 76, 46–50. Guo, F., Li, X., Liang, D., Li, T., Zhu, P., Guo, H., Wu, X., Wen, L., Gu, T.P., Hu, B., Walsh,
Clayton, D.F., Anreiter, I., Aristizabal, M., Frankland, P.W., Binder, E.B., Citri, A., 2019. C.P., Li, J., Tang, F., Xu, G.L., 2014a. Active and passive demethylation of male and
The role of the genome in experience-dependent plasticity: extending the analogy of female pronuclear DNA in the mammalian zygote. Cell Stem Cell 15, 447–459.
the genomic action potential. Proc. Natl. Acad. Sci. U. S. A. https://doi.org/10.1073/ Guo, H., et al., 2014b. The DNA methylation landscape of human early embryos. Nature
pnas.1820837116. 511, 606–610.
Collo, G., Cavalleri, L., Spano, P., 2014. Structural plasticity in mesencephalic dopami- Gupta, S., Kim, S.Y., Artis, S., Molfese, D.L., Schumacher, A., Sweatt, J.D., Paylor, R.E.,
nergic neurons produced by drugs of abuse: critical role of BDNF and dopamine. Lubin, F.D., 2010. Histone methylation regulates memory formation. J. Neurosci. 30,
Front. Pharmacol. 5, 259. 3589–3599.
Cortes-Mendoza, J., Diaz de Leon-Guerrero, S., Pedraza-Alva, G., Perez-Martinez, L., Hackett, J.A., Sengupta, R., Zylicz, J.J., Murakami, K., Lee, C., Down, T.A., Surani, M.A.,
2013. Shaping synaptic plasticity: the role of activity-mediated epigenetic regulation 2013. Germline DNA demethylation dynamics and imprint erasure through 5-hy-
on gene transcription. Int. J. Dev. Neurosci. 31, 359–369. droxymethylcytosine. Science 339, 448–452.
Cropley, J.E., Eaton, S.A., Aiken, A., Young, P.E., Giannoulatou, E., Ho, J.W.K., Buckland, Hajkova, P., Erhardt, S., Lane, N., Haaf, T., El-Maarri, O., Reik, W., Walter, J., Surani,
M.E., Keam, S.P., Hutvagner, G., Humphreys, D.T., Langley, K.G., Henstridge, D.C., M.A., 2002. Epigenetic reprogramming in mouse primordial germ cells. Mech. Dev.
Martin, D.I.K., Febbraio, M.A., Suter, C.M., 2016. Male-lineage transmission of an 117, 15–23.
acquired metabolic phenotype induced by grand-paternal obesity. Mol. Metab. 5, Hammoud, S.S., Nix, D.A., Zhang, H., Purwar, J., Carrell, D.T., Cairns, B.R., 2009.
699–708. Distinctive chromatin in human sperm packages genes for embryo development.
Crozatier, C., Guerriero, R.M., Mathieu, F., Giros, B., Nosten-Bertrand, M., Kosofsky, B.E., Nature 460, 473–478.
2003. Altered cocaine-induced behavioral sensitization in adult mice exposed to Hansen, K.H., Bracken, A.P., Pasini, D., Dietrich, N., Gehani, S.S., Monrad, A., Rappsilber,
cocaine in utero. Brain Res. Dev. Brain Res. 147, 97–105. J., Lerdrup, M., Helin, K., 2008. A model for transmission of the H3K27me3 epige-
Dahl, J.A., Jung, I., Aanes, H., Greggains, G.D., Manaf, A., Lerdrup, M., Li, G., Kuan, S., Li, netic mark. Nat. Cell Biol. 10, 1291–1300.
B., Lee, A.Y., Preissl, S., Jermstad, I., Haugen, M.H., Suganthan, R., Bjoras, M., He, F., Lidow, I.A., Lidow, M.S., 2006. Consequences of paternal cocaine exposure in
Hansen, K., Dalen, K.T., Fedorcsak, P., Ren, B., Klungland, A., 2016. Broad histone mice. Neurotoxicol. Teratol. 28, 198–209.
H3K4me3 domains in mouse oocytes modulate maternal-to-zygotic transition. Nature He, Y.F., Li, B.Z., Li, Z., Liu, P., Wang, Y., Tang, Q., Ding, J., Jia, Y., Chen, Z., Li, L., Sun,
537, 548–552. Y., Li, X., Dai, Q., Song, C.X., Zhang, K., He, C., Xu, G.L., 2011. Tet-mediated for-
Day, J.J., Sweatt, J.D., 2011. Cognitive neuroepigenetics: a role for epigenetic mechan- mation of 5-carboxylcytosine and its excision by TDG in mammalian DNA. Science
isms in learning and memory. Neurobiol. Learn. Mem. 96, 2–12. 333, 1303–1307.
Day, J.J., Childs, D., Guzman-Karlsson, M.C., Kibe, M., Moulden, J., Song, E., Tahir, A., Heller, E.A., et al., 2014. Locus-specific epigenetic remodeling controls addiction- and
Sweatt, J.D., 2013. DNA methylation regulates associative reward learning. Nat. depression-related behaviors. Nat. Neurosci. 17, 1720–1727.
Neurosci. 16, 1445–1452. Hill, P.W.S., Leitch, H.G., Requena, C.E., Sun, Z., Amouroux, R., Roman-Trufero, M.,
Dias, B.G., Ressler, K.J., 2014. Parental olfactory experience influences behavior and Borkowska, M., Terragni, J., Vaisvila, R., Linnett, S., Bagci, H., Dharmalingham, G.,
neural structure in subsequent generations. Nat. Neurosci. 17, 89–96. Haberle, V., Lenhard, B., Zheng, Y., Pradhan, S., Hajkova, P., 2018. Epigenetic re-
Dietz, D.M., Laplant, Q., Watts, E.L., Hodes, G.E., Russo, S.J., Feng, J., Oosting, R.S., programming enables the transition from primordial germ cell to gonocyte. Nature
Vialou, V., Nestler, E.J., 2011. Paternal transmission of stress-induced pathologies. 555, 392–396.
Biol. Psychiatry 70, 408–414. Holliday, R., Pugh, J.E., 1975. DNA modification mechanisms and gene activity during
Drickamer, L.C., Gowaty, P.A., Holmes, C.M., 2000. Free female mate choice in house development. Science 187, 226–232.
mice affects reproductive success and offspring viability and performance. Anim. Huypens, P., Sass, S., Wu, M., Dyckhoff, D., Tschop, M., Theis, F., Marschall, S., Hrabe de
Behav. 59, 371–378. Angelis, M., Beckers, J., 2016. Epigenetic germline inheritance of diet-induced obe-
Dunn, G.A., Bale, T.L., 2011. Maternal high-fat diet effects on third-generation female sity and insulin resistance. Nat. Genet. 48, 497–499.
body size via the paternal lineage. Endocrinology 152, 2228–2236. Inoue, A., Jiang, L., Lu, F., Suzuki, T., Zhang, Y., 2017. Maternal H3K27me3 controls DNA
Erkek, S., Hisano, M., Liang, C.Y., Gill, M., Murr, R., Dieker, J., Schubeler, D., van der methylation-independent imprinting. Nature 547, 419–424.
Vlag, J., Stadler, M.B., Peters, A.H., 2013. Molecular determinants of nucleosome Iqbal, K., Jin, S.G., Pfeifer, G.P., Szabo, P.E., 2011. Reprogramming of the paternal
retention at CpG-rich sequences in mouse spermatozoa. Nat. Struct. Mol. Biol. 20, genome upon fertilization involves genome-wide oxidation of 5-methylcytosine.
868–875. Proc. Natl. Acad. Sci. U. S. A. 108, 3642–3647.
Esteve, P.O., Chin, H.G., Smallwood, A., Feehery, G.R., Gangisetty, O., Karpf, A.R., Carey, Ito, S., D'Alessio, A.C., Taranova, O.V., Hong, K., Sowers, L.C., Zhang, Y., 2010. Role of
M.F., Pradhan, S., 2006. Direct interaction between DNMT1 and G9a coordinates Tet proteins in 5mC to 5hmC conversion, ES-cell self-renewal and inner cell mass
DNA and histone methylation during replication. Genes Dev. 20, 3089–3103. specification. Nature 466, 1129–1133.
Fagiolini, M., Jensen, C.L., Champagne, F.A., 2009. Epigenetic influences on brain de- Ito, S., Shen, L., Dai, Q., Wu, S.C., Collins, L.B., Swenberg, J.A., He, C., Zhang, Y., 2011.
velopment and plasticity. Curr. Opin. Neurobiol. 19, 207–212. Tet proteins can convert 5-methylcytosine to 5-formylcytosine and 5-carbox-
Fasshauer, M., Paschke, R., 2003. Regulation of adipocytokines and insulin resistance. ylcytosine. Science 333, 1300–1303.
Diabetologia 46, 1594–1603. Ivy, A.S., Brunson, K.L., Sandman, C., Baram, T.Z., 2008. Dysfunctional nurturing beha-
Feng, S., Jacobsen, S.E., Reik, W., 2010. Epigenetic reprogramming in plant and animal vior in rat dams with limited access to nesting material: a clinically relevant model
development. Science 330, 622–627. for early-life stress. Neuroscience 154, 1132–1142.
Finucane, M.M., Stevens, G.A., Cowan, M.J., Danaei, G., Lin, J.K., Paciorek, C.J., Singh, Ivy, A.S., Rex, C.S., Chen, Y., Dube, C., Maras, P.M., Grigoriadis, D.E., Gall, C.M., Lynch,
G.M., Gutierrez, H.R., Lu, Y., Bahalim, A.N., Farzadfar, F., Riley, L.M., Ezzati, M., G., Baram, T.Z., 2010. Hippocampal dysfunction and cognitive impairments pro-
Global Burden of Metabolic Risk Factors of Chronic Diseases Collaborating G, 2011. voked by chronic early-life stress involve excessive activation of CRH receptors. J.
National, regional, and global trends in body-mass index since 1980: systematic Neurosci. 30, 13005–13015.
analysis of health examination surveys and epidemiological studies with 960 country- Jackson, V., Granner, D.K., Chalkley, R., 1975. Deposition of histones onto replicating
years and 9.1 million participants. Lancet 377, 557–567. chromosomes. Proc. Natl. Acad. Sci. U. S. A. 72, 4440–4444.
Fischer, D.K., Rice, R.C., Martinez Rivera, A., Donohoe, M., Rajadhyaksha, A.M., 2017. Jaffe, A.E., Irizarry, R.A., 2014. Accounting for cellular heterogeneity is critical in epi-
Altered reward sensitivity in female offspring of cocaine-exposed fathers. Behav. genome-wide association studies. Genome Biol. 15, R31.
Brain Res. 332, 23–31. Johnson, N.L., Carini, L., Schenk, M.E., Stewart, M., Byrnes, E.M., 2011. Adolescent
Francis, D., Diorio, J., Liu, D., Meaney, M.J., 1999. Nongenomic transmission across opiate exposure in the female rat induces subtle alterations in maternal care and
generations of maternal behavior and stress responses in the rat. Science 286, transgenerational effects on play behavior. Front. Psychiatry 2, 29.
1155–1158. Kafri, T., Gao, X., Razin, A., 1993. Mechanistic aspects of genome-wide demethylation in
Gapp, K., Jawaid, A., Sarkies, P., Bohacek, J., Pelczar, P., Prados, J., Farinelli, L., Miska, the preimplantation mouse embryo. Proc. Natl. Acad. Sci. U. S. A. 90, 10558–10562.
E., Mansuy, I.M., 2014. Implication of sperm RNAs in transgenerational inheritance Kapoor, A., Matthews, S.G., 2005. Short periods of prenatal stress affect growth, beha-
of the effects of early trauma in mice. Nat. Neurosci. 17, 667–669. viour and hypothalamo-pituitary-adrenal axis activity in male Guinea pig offspring.
Gertz, J., Varley, K.E., Reddy, T.E., Bowling, K.M., Pauli, F., Parker, S.L., Kucera, K.S., J. Physiol. 566, 967–977.
Willard, H.F., Myers, R.M., 2011. Analysis of DNA methylation in a three-generation Kapoor, A., Kostaki, A., Janus, C., Matthews, S.G., 2009. The effects of prenatal stress on
family reveals widespread genetic influence on epigenetic regulation. PLoS Genet. 7, learning in adult offspring is dependent on the timing of the stressor. Behav. Brain
e1002228. Res. 197, 144–149.
Gibbs, J.R., van der Brug, M.P., Hernandez, D.G., Traynor, B.J., Nalls, M.A., Lai, S.L., Karpova, N.N., Sales, A.J., Joca, S.R., 2017. Epigenetic basis of neuronal and synaptic
Arepalli, S., Dillman, A., Rafferty, I.P., Troncoso, J., Johnson, R., Zielke, H.R., plasticity. Curr. Top. Med. Chem. 17, 771–793.
Ferrucci, L., Longo, D.L., Cookson, M.R., Singleton, A.B., 2010. Abundant quantita- Keller Jr., R.W., LeFevre, R., Raucci, J., Carlson, J.N., Glick, S.D., 1996. Enhanced cocaine
tive trait loci exist for DNA methylation and gene expression in human brain. PLoS self-administration in adult rats prenatally exposed to cocaine. Neurosci. Lett. 205,
Genet. 6, e1000952. 153–156.
Godfrey, P.A., Malnic, B., Buck, L.B., 2004. The mouse olfactory receptor gene family. Killinger, C.E., Robinson, S., Stanwood, G.D., 2012. Subtle biobehavioral effects produced
Proc. Natl. Acad. Sci. U. S. A. 101, 2156–2161. by paternal cocaine exposure. Synapse 66, 902–908.
Goldman, D., Oroszi, G., Ducci, F., 2005. The genetics of addictions: uncovering the Klengel, T., Binder, E.B., 2015. Epigenetics of stress-related psychiatric disorders and gene

13
J.J. Tuscher and J.J. Day Neurobiology of Disease 132 (2019) 104591

x environment interactions. Neuron 86, 1343–1357. Dnmt3b are essential for de novo methylation and mammalian development. Cell 99,
Korosi, A., Baram, T.Z., 2010. Plasticity of the stress response early in life: mechanisms 247–257.
and significance. Dev. Psychobiol. 52, 661–670. Oswald, J., Engemann, S., Lane, N., Mayer, W., Olek, A., Fundele, R., Dean, W., Reik, W.,
Kunkel, T.A., 2004. DNA replication fidelity. J. Biol. Chem. 279, 16895–16898. Walter, J., 2000. Active demethylation of the paternal genome in the mouse zygote.
Kwon, D.Y., Zhao, Y.T., Lamonica, J.M., Zhou, Z., 2017. Locus-specific histone deacety- Curr. Biol. 10, 475–478.
lation using a synthetic CRISPR-Cas9-based HDAC. Nat. Commun. 8, 15315. Pauli, A., Rinn, J.L., Schier, A.F., 2011. Non-coding RNAs as regulators of embryogenesis.
Laird, C.D., Pleasant, N.D., Clark, A.D., Sneeden, J.L., Hassan, K.M., Manley, N.C., Vary Nat. Rev. Genet. 12, 136–149.
Jr., J.C., Morgan, T., Hansen, R.S., Stoger, R., 2004. Hairpin-bisulfite PCR: assessing Perez, M.F., Lehner, B., 2019. Intergenerational and transgenerational epigenetic in-
epigenetic methylation patterns on complementary strands of individual DNA mo- heritance in animals. Nat. Cell Biol. 21, 143–151.
lecules. Proc. Natl. Acad. Sci. U. S. A. 101, 204–209. Pflueger, C., Tan, D., Swain, T., Nguyen, T., Pflueger, J., Nefzger, C., Polo, J.M., Ford, E.,
Lane, N., Dean, W., Erhardt, S., Hajkova, P., Surani, A., Walter, J., Reik, W., 2003. Lister, R., 2018. A modular dCas9-SunTag DNMT3A epigenome editing system
Resistance of IAPs to methylation reprogramming may provide a mechanism for overcomes pervasive off-target activity of direct fusion dCas9-DNMT3A constructs.
epigenetic inheritance in the mouse. Genesis 35, 88–93. Genome Res. 28, 1193–1206.
Lappalainen, T., Greally, J.M., 2017. Associating cellular epigenetic models with human Posner, R., Toker, I.A., Antonova, O., Star, E., Anava, S., Azmon, E., Hendricks, M.,
phenotypes. Nat. Rev. Genet. 18, 441–451. Bracha, S., Gingold, H., Rechavi, O., 2019. Neuronal small RNAs control behavior
Le, Q., Yan, B., Yu, X., Li, Y., Song, H., Zhu, H., Hou, W., Ma, D., Wu, F., Zhou, Y., Ma, L., transgenerationally. Cell 177 (1814–1826), e1815.
2017. Drug-seeking motivation level in male rats determines offspring susceptibility Puschendorf, M., Terranova, R., Boutsma, E., Mao, X., Isono, K., Brykczynska, U., Kolb, C.,
or resistance to cocaine-seeking behaviour. Nat. Commun. 8, 15527. Otte, A.P., Koseki, H., Orkin, S.H., van Lohuizen, M., Peters, A.H., 2008. PRC1 and
Lee, H.J., Hore, T.A., Reik, W., 2014. Reprogramming the methylome: erasing memory Suv39h specify parental asymmetry at constitutive heterochromatin in early mouse
and creating diversity. Cell Stem Cell 14, 710–719. embryos. Nat. Genet. 40, 411–420.
Ligtenberg, M.J., Kuiper, R.P., Chan, T.L., Goossens, M., Hebeda, K.M., Voorendt, M., Lee, Quadrana, L., Colot, V., 2016. Plant transgenerational epigenetics. Annu. Rev. Genet. 50,
T.Y., Bodmer, D., Hoenselaar, E., Hendriks-Cornelissen, S.J., Tsui, W.Y., Kong, C.K., 467–491.
Brunner, H.G., van Kessel, A.G., Yuen, S.T., van Krieken, J.H., Leung, S.Y., Radford, E.J., Isganaitis, E., Jimenez-Chillaron, J., Schroeder, J., Molla, M., Andrews, S.,
Hoogerbrugge, N., 2009. Heritable somatic methylation and inactivation of MSH2 in Didier, N., Charalambous, M., McEwen, K., Marazzi, G., Sassoon, D., Patti, M.E.,
families with Lynch syndrome due to deletion of the 3′ exons of TACSTD1. Nat. Ferguson-Smith, A.C., 2012. An unbiased assessment of the role of imprinted genes in
Genet. 41, 112–117. an intergenerational model of developmental programming. PLoS Genet. 8,
Liu, W.M., Pang, R.T., Chiu, P.C., Wong, B.P., Lao, K., Lee, K.F., Yeung, W.S., 2012. e1002605.
Sperm-borne microRNA-34c is required for the first cleavage division in mouse. Proc. Ragunathan, K., Jih, G., Moazed, D., 2015. Epigenetics. Epigenetic inheritance uncoupled
Natl. Acad. Sci. U. S. A. 109, 490–494. from sequence-specific recruitment. Science 348, 1258699.
Liu, X.S., Wu, H., Ji, X., Stelzer, Y., Wu, X., Czauderna, S., Shu, J., Dadon, D., Young, R.A., Rassoulzadegan, M., Grandjean, V., Gounon, P., Vincent, S., Gillot, I., Cuzin, F., 2006.
Jaenisch, R., 2016. Editing DNA methylation in the mammalian genome. Cell 167 RNA-mediated non-mendelian inheritance of an epigenetic change in the mouse.
(233–247), e217. Nature 441, 469–474.
Lubin, F.D., Roth, T.L., Sweatt, J.D., 2008. Epigenetic regulation of BDNF gene tran- Rechavi, O., Minevich, G., Hobert, O., 2011. Transgenerational inheritance of an acquired
scription in the consolidation of fear memory. J. Neurosci. 28, 10576–10586. small RNA-based antiviral response in C. elegans. Cell 147, 1248–1256.
Margueron, R., Justin, N., Ohno, K., Sharpe, M.L., Son, J., Drury 3rd, W.J., Voigt, P., Renthal, W., Nestler, E.J., 2008. Epigenetic mechanisms in drug addiction. Trends Mol.
Martin, S.R., Taylor, W.R., De Marco, V., Pirrotta, V., Reinberg, D., Gamblin, S.J., Med. 14, 341–350.
2009. Role of the polycomb protein EED in the propagation of repressive histone Rice, C.J., Sandman, C.A., Lenjavi, M.R., Baram, T.Z., 2008. A novel mouse model for
marks. Nature 461, 762–767. acute and long-lasting consequences of early life stress. Endocrinology 149,
Marmorstein, R., 2001. Structure and function of histone acetyltransferases. Cell. Mol. 4892–4900.
Life Sci. 58, 693–703. Rodgers, A.B., Morgan, C.P., Bronson, S.L., Revello, S., Bale, T.L., 2013. Paternal stress
Martini, M., Valverde, O., 2012. A single episode of maternal deprivation impairs the exposure alters sperm microRNA content and reprograms offspring HPA stress axis
motivation for cocaine in adolescent mice. Psychopharmacology 219, 149–158. regulation. J. Neurosci. 33, 9003–9012.
Mashoodh, R., Franks, B., Curley, J.P., Champagne, F.A., 2012. Paternal social enrich- Rodgers, A.B., Morgan, C.P., Leu, N.A., Bale, T.L., 2015. Transgenerational epigenetic
ment effects on maternal behavior and offspring growth. Proc. Natl. Acad. Sci. U. S. programming via sperm microRNA recapitulates effects of paternal stress. Proc. Natl.
A. 109 (Suppl. 2), 17232–17238. Acad. Sci. U. S. A. 112, 13699–13704.
Masuyama, H., Mitsui, T., Eguchi, T., Tamada, S., Hiramatsu, Y., 2016. The effects of Rougier, N., Bourc'his, D., Gomes, D.M., Niveleau, A., Plachot, M., Paldi, A., Viegas-
paternal high-fat diet exposure on offspring metabolism with epigenetic changes in Pequignot, E., 1998. Chromosome methylation patterns during mammalian pre-
the mouse adiponectin and leptin gene promoters. Am. J. Physiol. Endocrinol. Metab. implantation development. Genes Dev. 12, 2108–2113.
311, E236–E245. Sakashita, A., Kobayashi, H., Wakai, T., Sotomaru, Y., Hata, K., Kono, T., 2014. Dynamics
Mathers, J.C., Strathdee, G., Relton, C.L., 2010. Induction of epigenetic alterations by of genomic 5-hydroxymethylcytosine during mouse oocyte growth. Genes Cells 19,
dietary and other environmental factors. Adv. Genet. 71, 3–39. 629–636.
Mayer, W., Niveleau, A., Walter, J., Fundele, R., Haaf, T., 2000. Demethylation of the Santos, F., Hendrich, B., Reik, W., Dean, W., 2002. Dynamic reprogramming of DNA
zygotic paternal genome. Nature 403, 501–502. methylation in the early mouse embryo. Dev. Biol. 241, 172–182.
McClelland, S., Korosi, A., Cope, J., Ivy, A., Baram, T.Z., 2011. Emerging roles of epi- Savell, K.E., Bach, S.V., Zipperly, M.E., Revanna, J.S., Goska, N.A., Tuscher, J.J., Duke,
genetic mechanisms in the enduring effects of early-life stress and experience on C.G., Sultan, F.A., Burke, J.N., Williams, D., Ianov, L., Day, J.J., 2019. A neuron-
learning and memory. Neurobiol. Learn. Mem. 96, 79–88. optimized CRISPR/dCas9 activation system for robust and specific gene regulation.
McEwen, B.S., Eiland, L., Hunter, R.G., Miller, M.M., 2012. Stress and anxiety: structural eNeuro 6.
plasticity and epigenetic regulation as a consequence of stress. Neuropharmacology Seisenberger, S., Andrews, S., Krueger, F., Arand, J., Walter, J., Santos, F., Popp, C.,
62, 3–12. Thienpont, B., Dean, W., Reik, W., 2012. The dynamics of genome-wide DNA me-
Miller, C.A., Sweatt, J.D., 2007. Covalent modification of DNA regulates memory for- thylation reprogramming in mouse primordial germ cells. Mol. Cell 48, 849–862.
mation. Neuron 53, 857–869. Serobyan, V., Sommer, R.J., 2017. Developmental systems of plasticity and trans-gen-
Minkina, O., Hunter, C.P., 2018. Intergenerational transmission of gene regulatory in- erational epigenetic inheritance in nematodes. Curr. Opin. Genet. Dev. 45, 51–57.
formation in Caenorhabditis elegans. Trends Genet. 34, 54–64. Sharma, U., Conine, C.C., Shea, J.M., Boskovic, A., Derr, A.G., Bing, X.Y., Belleannee, C.,
Monahan, K., Schieren, I., Cheung, J., Mumbey-Wafula, A., Monuki, E.S., Lomvardas, S., Kucukural, A., Serra, R.W., Sun, F., Song, L., Carone, B.R., Ricci, E.P., Li, X.Z.,
2017. Cooperative interactions enable singular olfactory receptor expression in Fauquier, L., Moore, M.J., Sullivan, R., Mello, C.C., Garber, M., Rando, O.J., 2016.
mouse olfactory neurons. Elife 6. Biogenesis and function of tRNA fragments during sperm maturation and fertilization
Moore, R.S., Kaletsky, R., Murphy, C.T., 2019. Piwi/PRG-1 argonaute and TGF-beta in mammals. Science 351, 391–396.
mediate transgenerational learned pathogenic avoidance. Cell 177, 1827–1841 Siklenka, K., Erkek, S., Godmann, M., Lambrot, R., McGraw, S., Lafleur, C., Cohen, T., Xia,
(e1812). J., Suderman, M., Hallett, M., Trasler, J., Peters, A.H., Kimmins, S., 2015. Disruption
Morgan, H.D., Santos, F., Green, K., Dean, W., Reik, W., 2005. Epigenetic reprogramming of histone methylation in developing sperm impairs offspring health transgener-
in mammals. Hum. Mol. Genet. 14 (1), R47–R58. ationally. Science 350, aab2006.
Mueller, B.R., Bale, T.L., 2007. Early prenatal stress impact on coping strategies and Silva, A.J., Ward, K., White, R., 1993. Mosaic methylation in clonal tissue. Dev. Biol. 156,
learning performance is sex dependent. Physiol. Behav. 91, 55–65. 391–398.
Mueller, B.R., Bale, T.L., 2008. Sex-specific programming of offspring emotionality after Skinner, M.K., 2008. What is an epigenetic transgenerational phenotype? F3 or F2.
stress early in pregnancy. J. Neurosci. 28, 9055–9065. Reprod. Toxicol. 25, 2–6.
Myers Jr., M.G., 2004. Leptin receptor signaling and the regulation of mammalian phy- Smith, Z.D., Chan, M.M., Mikkelsen, T.S., Gu, H., Gnirke, A., Regev, A., Meissner, A.,
siology. Recent Prog. Horm. Res. 59, 287–304. 2012. A unique regulatory phase of DNA methylation in the early mammalian em-
Nilsson, E.E., Sadler-Riggleman, I., Skinner, M.K., 2018. Environmentally induced epi- bryo. Nature 484, 339–344.
genetic transgenerational inheritance of disease. Environ. Epigenet 4 (dvy016). Stankiewicz, A.M., Swiergiel, A.H., Lisowski, P., 2013. Epigenetics of stress adaptations in
Norouzitallab, P., Baruah, K., Vandegehuchte, M., Van Stappen, G., Catania, F., Vanden the brain. Brain Res. Bull. 98, 76–92.
Bussche, J., Vanhaecke, L., Sorgeloos, P., Bossier, P., 2014. Environmental heat stress Statham, A.L., Robinson, M.D., Song, J.Z., Coolen, M.W., Stirzaker, C., Clark, S.J., 2012.
induces epigenetic transgenerational inheritance of robustness in parthenogenetic Bisulfite sequencing of chromatin immunoprecipitated DNA (BisChIP-seq) directly
Artemia model. FASEB J. 28, 3552–3563. informs methylation status of histone-modified DNA. Genome Res. 22, 1120–1127.
Ogden, C.L., Carroll, M.D., Flegal, K.M., 2014. Prevalence of obesity in the United States. Tahiliani, M., Koh, K.P., Shen, Y., Pastor, W.A., Bandukwala, H., Brudno, Y., Agarwal, S.,
JAMA 312, 189–190. Iyer, L.M., Liu, D.R., Aravind, L., Rao, A., 2009. Conversion of 5-methylcytosine to 5-
Okano, M., Bell, D.W., Haber, D.A., Li, E., 1999. DNA methyltransferases Dnmt3a and hydroxymethylcytosine in mammalian DNA by MLL partner TET1. Science 324,

14
J.J. Tuscher and J.J. Day Neurobiology of Disease 132 (2019) 104591

930–935. Dymov, S., Szyf, M., Meaney, M.J., 2004. Epigenetic programming by maternal be-
Tang, F., Kaneda, M., O'Carroll, D., Hajkova, P., Barton, S.C., Sun, Y.A., Lee, C., havior. Nat. Neurosci. 7, 847–854.
Tarakhovsky, A., Lao, K., Surani, M.A., 2007. Maternal microRNAs are essential for Wei, Y., Yang, C.R., Wei, Y.P., Zhao, Z.A., Hou, Y., Schatten, H., Sun, Q.Y., 2014.
mouse zygotic development. Genes Dev. 21, 644–648. Paternally induced transgenerational inheritance of susceptibility to diabetes in
Tang, W.W., Dietmann, S., Irie, N., Leitch, H.G., Floros, V.I., Bradshaw, C.R., Hackett, mammals. Proc. Natl. Acad. Sci. U. S. A. 111, 1873–1878.
J.A., Chinnery, P.F., Surani, M.A., 2015. A unique gene regulatory network resets the White, S.L., Vassoler, F.M., Schmidt, H.D., Pierce, R.C., Wimmer, M.E., 2016. Enhanced
human germline epigenome for development. Cell 161, 1453–1467. anxiety in the male offspring of sires that self-administered cocaine. Addict. Biol. 21,
Tsankova, N., Renthal, W., Kumar, A., Nestler, E.J., 2007. Epigenetic regulation in psy- 802–810.
chiatric disorders. Nat. Rev. Neurosci. 8, 355–367. Wimmer, M.E., Briand, L.A., Fant, B., Guercio, L.A., Arreola, A.C., Schmidt, H.D., Sidoli,
van de Werken, C., van der Heijden, G.W., Eleveld, C., Teeuwssen, M., Albert, M., S., Han, Y., Garcia, B.A., Pierce, R.C., 2017. Paternal cocaine taking elicits epigenetic
Baarends, W.M., Laven, J.S., Peters, A.H., Baart, E.B., 2014. Paternal hetero- remodeling and memory deficits in male progeny. Mol. Psychiatry 22, 1641–1650.
chromatin formation in human embryos is H3K9/HP1 directed and primed by sperm- Xu, Q., Xie, W., 2018. Epigenome in early mammalian development: inheritance, re-
derived histone modifications. Nat. Commun. 5, 5868. programming and establishment. Trends Cell Biol. 28, 237–253.
Vassoler, F.M., Sadri-Vakili, G., 2014. Mechanisms of transgenerational inheritance of Yang, Q., Lin, J., Liu, M., Li, R., Tian, B., Zhang, X., Xu, B., Liu, M., Zhang, X., Li, Y., Shi,
addictive-like behaviors. Neuroscience 264, 198–206. H., Wu, L., 2016. Highly sensitive sequencing reveals dynamic modifications and
Vassoler, F.M., White, S.L., Schmidt, H.D., Sadri-Vakili, G., Pierce, R.C., 2013. Epigenetic activities of small RNAs in mouse oocytes and early embryos. Sci. Adv. 2, e1501482.
inheritance of a cocaine-resistance phenotype. Nat. Neurosci. 16, 42–47. Zee, B.M., Levin, R.S., Xu, B., LeRoy, G., Wingreen, N.S., Garcia, B.A., 2010. In vivo
Vickers, M.H., 2014. Early life nutrition, epigenetics and programming of later life dis- residue-specific histone methylation dynamics. J. Biol. Chem. 285, 3341–3350.
ease. Nutrients 6, 2165–2178. Zhang, B., et al., 2016. Allelic reprogramming of the histone modification H3K4me3 in
Vink, J.M., 2016. Genetics of addiction: future focus on gene x environment interaction? early mammalian development. Nature 537, 553–557.
J. Stud. Alcohol Drugs 77, 684–687. Zhang, Y., et al., 2018. Dnmt2 mediates intergenerational transmission of paternally
Vojta, A., Dobrinic, P., Tadic, V., Bockor, L., Korac, P., Julg, B., Klasic, M., Zoldos, V., acquired metabolic disorders through sperm small non-coding RNAs. Nat. Cell Biol.
2016. Repurposing the CRISPR-Cas9 system for targeted DNA methylation. Nucleic 20, 535–540.
Acids Res. 44, 5615–5628. Zheng, H., Huang, B., Zhang, B., Xiang, Y., Du, Z., Xu, Q., Li, Y., Wang, Q., Ma, J., Peng,
Waddington, C.H., 1953. Genetic assimilation of an acquired character. Evolution 7, X., Xu, F., Xie, W., 2016. Resetting epigenetic memory by reprogramming of histone
118–126. modifications in mammals. Mol. Cell 63, 1066–1079.
Walker, D.M., Nestler, E.J., 2018. Neuroepigenetics and addiction. Handb. Clin. Neurol. Zovkic, I.B., Paulukaitis, B.S., Day, J.J., Etikala, D.M., Sweatt, J.D., 2014. Histone H2A.Z
148, 747–765. subunit exchange controls consolidation of recent and remote memory. Nature 515,
Weaver, I.C., Cervoni, N., Champagne, F.A., D'Alessio, A.C., Sharma, S., Seckl, J.R., 582–586.

15

You might also like