<<

Child , January/February 2010, Volume 81, Number 1, Pages 41–79

Epigenetics and the Biological Definition of · Environment Interactions

Michael J. Meaney McGill University

Variations in reflect the influence of environmental conditions during development on cellular functions, including that of the . The recent integration of into developmental psychobiol- ogy illustrates the processes by which environmental conditions in early structurally alter DNA, provid- ing a physical basis for the influence of the perinatal environmental signals on phenotype over the life of the individual. This review focuses on the enduring effects of naturally occurring variations in maternal care on and phenotype to provide an example of environmentally driven plasticity at the level of the DNA, revealing the interdependence of gene and environmental in the regulation of phenotype.

The –nurture debate is essentially a question prevention and intervention programs. In the social of the determinants of individual differences in the sciences, and particularly , there has expression of specific traits among members of the generally been an understandable bias in favor of same species. The origin of the terms nature and explanations derived from the ‘‘nurture’’ perspec- nurture has been credited to Richard Mulcaster, a tive, which emphasizes the capacity for environ- British teacher who imagined these influences as mentally induced plasticity in brain structure and collaborative forces that shape child development . Nevertheless, over recent decades there (West & King, 1987). History has conspired to per- has been a gradual, if at times reluctant acceptance vert Mulcaster’s intent, casting genetic and environ- that genomic variations contribute to individual dif- mental influences as independent agents in the ferences in brain development and function. And field of development. The intent of this article is to thus the integration of into psychology and examine the biological context in which genetic and psychiatry has become one of the more active environmental signals actually operate. The result research areas in the study of neural function. Now of this analysis produces an impression that lies that we have come to accept the idea that genomic close to Mulcaster’s formulation. variation influences brain development and func- Any successful attempt to constructively leverage tion in , it is time to consider what this the remarkable advances of the genomic era will should mean in the reformulation of intervention depend upon our ability to understand genetic programs that target children and their families. influences and their interactions with the environ- The idea that variation in the expression of neu- mental context within which they operate. Hinder- ral functions would lie outside the dominion of ing such efforts is the rather arcane notion that we genomic influences is biologically untenable. Like can partition the causes of individual differences any tissue, the development of the mammalian into distinct genetic or environmental spheres of brain and its function occurs as a result of coordi- influence. This issue is of particular importance in nated influences that include heritable variation in the area of child development where research that genomic sequence. What remains as a major chal- examines the origins of individual differences in lenge for the psychological sciences is the develop- complex traits often forms the basis for extensive ment of a conceptual framework from which to intervention or prevention programs (Fisher, Gun- meaningfully understand the results of studies that nar, Chamberlain, & Reid, 2000; Olds et al., 1998). examine brain development and function at the Our assumptions concerning the processes of child level of the . However, we remain mired development influence the design of such pro- in additive models of explanation within which grams. Thus, one of the great challenges at this time phenotypic development rolls out as the cumula- is that of integrating into the design of tive by product of genetic + environmental

Correspondence concerning this article should be addressed to Michael J. Meaney, Douglas Mental Health University Institute, 2010, Copyright the Author(s) McGill University, 6875 LaSalle Blvd, Montreal, QC, Canada H4H Journal Compilation 2010, Society for Research in Child Development, Inc. 1R3. Electronic mail may be sent to [email protected]. All rights reserved. 0009-3920/2010/8101-0005 42 Meaney influences, with each discipline adding indepen- ences in neural function (Griffiths & Stotz, 2007; dently to the equation. Science in the postgenomic Griffiths & Tabery, 2008; Moss, 2005). era requires a more sophisticated, biologically The issue becomes even more complicated when informed understanding of the interplay between considering the interaction between gene and envi- gene and environment in defining individual ronment, since the statistical approaches to the differences at the level of complex traits. study of Gene · Environment interactions are com- The fundamental question is that of the relation plicated and not yet fully developed. Indeed, since between genotype and phenotype: How do varia- the time of Fisher and Haldane there has been tions in the genotype influence selected traits at any debate concerning the adequacy of analysis of vari- point over the lifespan? This question is addressed ance (ANOVA) models in detecting Gene · Envi- very differently in and the social sci- ronment interactions (Wahlsten, 1990; see also ences than it is in the biological sciences. Indeed, the Griffiths & Stotz, 2007; Griffiths & Tabery, 2008, for actual meaning and interpretation of the term gene historical reviews). The primary objective of this can differ across research domains (Griffiths & Ta- article is to examine the Gene · Environment inter- bery, 2008). For epidemiologists or those employing action from the biological perspective. For whatever epidemiological approaches, genotype–phenotype the difficulties in statistically defining Gene · Envi- relations are defined by statistical probabilities. A ronment interactions, research in reveals gene for aggression is thus defined as a genomic that the genome cannot possibly operate indepen- region within which a defined variation in sequence dent of its environmental context. The biological predicts a specific phenotypic outcome, in this case perspective reveals the futility of the nature– one associated with aggressive or actions. nurture debate and of additive models of genetic Here, genotype–phenotype relations are statistical. and environmental influences in defining pheno- For the biological scientist, genotype–phenotype type. Moreover, recent studies from the field of relations are defined by the actual physical operation epigenetics provide insight into biological mecha- of a genomic region in relation to the proximal cellu- nisms whereby Gene · Environment interactions lar events that directly mediate the behavioral varia- can biochemically alter the genome and thus stably tion as well as the more distal influences that influence individual differences in neural function. regulate the relevant cellular signals. Thus, a gene for Developmentalists studying the enduring effects of aggression would be defined as a genomic region early experience on brain development and func- from which a transcribed product actively partici- tion have long anticipated such processes (Bateson pates in a biological pathway that directly mediates et al., 2004; Gottlieb, 1997, 1998; Schneirla, 1966). the expression of aggressive behaviors. The studies This approach envisions development as an active of Young and colleagues (Donaldson & Young, 2008; process of that occurs as a function of Hammock & Young, 2005) on the sequence variation the continuous dialog between the genome and its in the vasopressin 1A gene provide an environment. With the integration of epigenetics excellent example. Among voles, subspecies differ- into developmental psychobiology, we can now ences in social behaviors are associated with a begin to define the physical basis for this interactive sequence variation in the vasopressin 1A receptor process. gene and differential expression of the vasopressin 1A receptor. Reversing the pattern of vasopressin 1A receptor expression eliminates the differences in Statistical Approaches to the Study of the selected, vasopressin-mediated social behaviors (see ‘‘Instrumental’’ Gene Robinson, Fernald, & Clayton, 2008, for additional examples). These studies describe a causal geno- Research examining the potential influence of heri- type–phenotype relation. Within the biological sci- table variations in DNA sequences on brain devel- ences, the mere statistical relation between a opment and function in humans has relied largely genomic variation and phenotypic outcome would upon two approaches. The first is a classical epide- not be considered as sufficient basis for the establish- miological approach that involves association or ment of a functional link between the gene (or its linkage analyses that essentially use correlational product) and the relevant variation in phenotype. approaches to describe genotype–phenotype rela- Such differences in perspective have led to rather tions. Much as epidemiological research statistically lively debates between, for example, researchers in associates specific exposures with health outcomes, quantitative behavioral genetics and researchers such studies correlate variation in examining the biological basis for individual differ- sequence at specific sites in the genome with a Epigenetics 43 phenotypic outcome. This is certainly a reasonable precise studies that directly examine the functional approach in the initial phase of examining the significance of that genomic variant, as well as the potential role for a specific genomic variant (Plomin particular circumstances under which the geno- & Rutter, 1998). The advent of high-throughput type–phenotype relation is apparent (Plomin & technology permits genome-wide association stud- Rutter, 1998). ies that can potentially scan 1 million or more sin- A second statistical approach is one that has gle nucleotide polymorphisms in individual evoked greater controversy. Quantitative behavioral subjects. While there are inherent statistical compli- genetics does not examine genotype–phenotype cations, such as controlling for multiple compari- relations at the level of specific genomic variants sons, the notion of being able to examine the but rather attempts to partition the variance in the relation between 106 sequence variants with a sin- expression of a specific phenotypic traits across a gle trait of interest across a sufficiently large sample population into that associated with ‘‘genetic’’ or is incredibly seductive. Nevertheless, such studies ‘‘environmental’’ influences (Plomin, DeFries, & do not inform at the level of causal mechanism. A McClearn, 1990). Such studies actually examine correlation between a genomic variant and a devel- so-called genetic effects without ever directly exam- opmental outcome is no less correlational and no ining the genome let alone variation in the genome more explanatory (or mechanistic) than is the corre- (Sokolowski & Wahlsten, 2001). Indeed within lation between, say, level of education or a dietary quantitative behavioral genetics the issue that is constituent and the same developmental outcome. directly under study is that of . Genetic The popular press often trumpets such findings effects are simply implied on the basis of data (usually simply echoing the institute’s own press derived from studies of the heritability of traits. release) as the identification of a gene for an out- This statistical approach, based on the development come of interest. Studies that, for example, statisti- of ANOVA models by Fisher and Haldane at the cally identify a relation between a genomic turn of the last century, has a long history in agri- and the variation in the occurrence culture and sciences as the basis for selec- of conduct disorder are interpreted as successfully tive breeding for specific traits (Griffiths & Tabery, revealing the gene for conduct disorder. Such inter- 2008; Wahlsten, 2003). Such endeavors do not rely pretations drive many scientists slightly mad, since on the assumption that heritability necessarily the data are merely correlational and do not imply reflects a genomic mechanism for the transmission a cause–effect relation between the genomic variant of individual differences in the trait from parent to and conduct disorder. The gene might indeed be offspring. What is critical for is expressed only in the liver and influence the devel- the ability to reliably estimate the passage of traits opment or activity of the relevant neural systems from the parent to offspring. Traits with greater that regulate social behavior only through a very heritability are better candidates for selective breed- general metabolic pathway that fuels brain devel- ing. The actual mechanism of inheritance is irrele- opment in early life. There may be no direct effect vant. I think that the situation in the biological on any neuronal population whatsoever, let alone sciences is mixed. For some, heritability simply the complex neural network that underlies the abil- reflects familial transmission. For others, it has ity to regulate affect and behavior. Nevertheless, come to be equated with a true estimate of the genomic variant is indeed statistically associ- variation due to variation in genomic sequence. ated with conduct disorder. The controversy sur- Estimates of heritability are commonly equated rounding the claims from association or linkage with ‘‘genetic’’ influences. It is this assumption that studies lie in the interpretation of the data and the proves controversial. implicit confusion of correlation with causation. Johannsen (1911) first introduced the terms The data are safely assumed to provide the identifi- genotype and phenotype and the relation between the cation of potential influence over the phenotype in two became perhaps the defining question in biol- question. However annoying the misuse of such ogy. In the study of genotype–phenotype relations, correlational data in public or professional dis- the term phenotype refers to all aspects of gene course, the findings serve as a starting point for function and may thus be used to describe any research examining the possible functional relation observable characteristic of an , such as its between the genomic variant and the phenotype. In , biochemical or physiological proper- the case of our example, the nature of the relation ties, or behavior (Johannsen, 1911; Lewontin, 1974). between the genomic variant of interest and con- The gene had actually been the subject of scientific duct disorder must then become a matter of more analysis well before the early 20th century and long 44 Meaney before the identification of DNA as the physical population and not as an explanation for individual basis for the ‘‘.’’ Over this period, the patterns of development (Griffiths & Tabery, 2008). gene was considered as the hypothetical physical Studies of the heritability of selected traits thus constituent of the , the transmission of which yield statistical and not biological evidence for a from parent to offspring served as the basis for genetic influence. The irritation derives from the (Griffiths & Stotz, 2007; Grif- interpretation of such findings, and in particular fiths & Tabery, 2008). The gene was defined as the the notion of ‘‘main effects,’’ which implies that the ‘‘unit of .’’ And thus it was understandable genome and its environment can operate indepen- that during this period estimates of heredity were dently. And of course the angst should be no less assumed to reflect a genetic influence. This was for environmental main effects than for those asso- true by definition. Griffiths and Tabery (2008) thus ciated with genotype. Again the issue is that of refer to this entity as the ‘‘traditional gene’’ interpretation and an appreciation of what such —the substance of inheritance. For the early findings actually do and do not represent. pioneers in genetics such as Morgan, and many studies derive estimates of heritability. studying patterns of inheritance to this day, the Such findings are of considerable importance for actual physical identity and operation of the gene is their own sake, particularly for those interested in irrelevant (Falk, 1986). This tradition continues issues surrounding familial transmission of vulner- within the study of quantitative behavioral genetics ability for illness. An indication of high heritability for which the gene is simply viewed as the unit of for a specific trait logically leads to the question of inheritance, with apparently little concern for the whether genomic variations at the level of sequence actual biology that might mediate the relation of might serve as the mechanism for the passage of the genomic variant, cellular function and the trait individual differences in phenotype from parent to of interest. offspring (prior to investing valuable research A classical research approach in quantitative funds in genome-wide searchers, it is indeed useful behavioral genetics has employed studies of to establish that the variation in the trait is actually to infer ‘‘genetic’’ influences. Twin studies rely on subject to familial transmission). But the idea that the level of similarity in specific traits between estimates of heritability are the equivalent of a monozygotic and dizygotic twins to derive esti- genetic effect is problematic. First, such estimates mates of heritability (h2). As noted earlier, such ignore Gene · Environment interactions. Indeed, measures were developed in animal sciences using heritability equates with a main effect ‘‘genetic’’ parent–offspring regressions to study intergenera- influences only if one assumes no Gene · Environ- tional transmission of specific traits. The intent of ment contribution. Second, the estimate of heritabil- such studies is not to identify the causal interplay ity reflects the importance of genomic mechanisms between gene and environment in defining the only if one assumes that there are no other biologi- development and expression of a particular trait in cal mechanisms for inheritance. This is simply an individual but rather to estimate the relative untrue. There are multiple potential mechanisms of contribution of genetic and environmental influ- inheritance, involving, for example, the passage of ences in defining the variation in the trait across a epigenetic marks (Chong & Whitelaw, 2004) population (Plomin et al., 1990). The focus on ‘‘rela- through the , the passage of maternal RNA tive contributions’’ of gene and environment as molecules into the , the potential passage of independent forces is the hallmark of quantitative from parent to offspring, the bio- behavioral genetics and assumes that such forces chemical state of the gametes at the time of concep- can act independently of one another. In following tion, and the transmission of nutrients, bacteria, or the biometric tradition of Fisher and Haldane, the antibodies from maternal circulation to that of the variance in the expression of a offspring, and so on (Bettegowda & Smith, 2007; across a population is examined using ANOVA Boulinier & Staszewski, 2008; Chong & Whitelaw, models that partition variance in phenotype accord- 2004; Grindstaff, Brodie, & Ketterson, 2003; Hassel- ing to genetic and environmental factors. Note that quist & Nilsson, 2009; Rassoulzadegan et al., 2006; within such analyses the ‘‘genetic’’ influences are Shorter & Lindquist, 2005). Variations in genomic defined statistically and not biologically, and of sequence are simply not the only mechanism of course the same is true for the definition of inheritance. All of these factors can and do influ- so-called environmental influences. Moreover, ence the phenotype of the offspring. This is a seri- the resulting analyses are an attempt to understand ous problem for those assuming that heritability the source of variation in phenotype across a equates with main effects of genomic variation. Epigenetics 45

Apart from the methodological and biological Throughout portions of this review I will largely pitfalls with such inferences (Lewontin, 1974; Soko- focus on one gene, that encoding for the glucocorti- lowski & Wahlsten, 2001; Wahlsten, 1990), the argu- coid receptor, both as an exemplar, for which it is ment presented here is that the notion of a main rather well suited, and as a subject for our own effect of either ‘‘gene’’ or ‘‘environment’’ on the research. The belongs to the development of individual differences in complex nuclear receptor superfamily and is transcribed traits is biologically fallacious. Instead, the develop- from a single mRNA. The glucocorticoid receptor ment of the individual is best considered as the is found both in the cytoplasm and in the emergent property of a constant interplay between nucleus of the cell. The receptor belongs to a class the genome and its environment (Gottlieb, 1991, of intracellular proteins referred to as -acti- 1997). This article briefly examines how the study vated factors. This simply refers to the of gene activity or expression reveals why main fact that the binding of the hormone (the ligand) to effects models are biologically implausible and then the receptor alters the configuration of the glucocor- reviews studies on the effects of maternal care on ticoid receptor, which then permits binding to spe- glucocorticoid receptor expression in the rat as cific sites on the DNA and the regulation of gene a model for the understanding one process by transcription. Glucocorticoids, a hallmark of the which Gene · Environment interactions might endocrine stress response, are hormones secreted operate to influence the development of individual by the adrenal cortex that bind to the glucocorticoid differences. receptor to mediate various cellular processes including those involved in cardiovascular activity, appetite, metabolism, immune responses, electro- lyte balance, as well as neuronal function and The Molecular Gene and Its Operation behavior (e.g., Dallman, Akana, Strack, Hanson, & The challenge within is Sebastian, 1995; Dallman, Pecoraro, & la Fleur, 2005; that of understanding the operation of the genome de Kloet, Karst, & Joels, 2008; Heitzer, Wolf, and how the analysis of nucleotide-based variation Sanchez, Witchel, & DeFranco, 2007; Kumar & can contribute to our understanding of individual Thompson, 2005; Munck, 2005; Munck, Guyre, & differences in brain development and function at Holbrook, 1984; Sapolsky, Romero, & Munck, 2000; the level of , , and emotional- Seckl & Holmes, 2007; Walker, 2007). Thus, the gene cognitive states. One obvious way to start is to sim- encoding the glucocorticoid receptor has multiple, ply appreciate the role of DNA in cell biology and tissue-specific functions. The activation of the gluco- realize that DNA directly codes for molecules, more corticoid receptor alters gene expression, the num- specifically for ribonucleic acids (), and not ber and identity of which depends very much upon for higher levels of function. While there are indeed the target cell. When not bound by glucocorticoids, statistical relations between variation in nucleotide the glucocorticoid receptor resides in the cytoplasm sequence and that in complex traits, at the level of of the cell associated with what is termed a chaper- biology there are no for , depres- one protein (Kumar & Thompson, 2005; Lu & Cid- sion, athletic abilities, fashion sense, or any other lowski, 2006; McNally, Mu¨ ller, Walker, Wolford, & such complex trait. Rather, there are certain varia- Hager, 2000). Binding of glucocorticoids induces tions in genomic sequences that can potentially conformational changes in the receptor; dissociation alter either the DNA product (RNA) or the degree from the chaperone proteins; association with to which the DNA is transcribed (i.e., actively pro- another, activated glucocorticoid receptor protein, ducing its molecular product). There are multiple thus forming a dimer (association of two proteins); and complex cellular processes that lie between the translocation into the nuclear compartment; and DNA sequence and the functional outcome associ- binding to specific sites on the DNA (Lu & Cidlow- ated with the gene product. The relation between ski, 2006 and references therein). The activated genotype and phenotype, even at the level of cellu- glucocorticoid receptor can also complex with other lar molecules, is anything but direct. Thus, the proteins, or cofactors, that then modify the effect of intent of this rather brief section is to simply illus- glucocorticoid receptor on the expression of target trate the complexity of the processes that mediate genes. The association of the cofactor may simply genotype–phenotype relations even when consider- modify the ability of the glucocorticoid receptor to ing the most fundamental aspects of gene func- access DNA sites, and thus the magnitude of its tion—the production of RNAs and protein transcriptional effect. However, the presence of the products. cofactor may determine the nature of the affected 46 Meaney genomic target. Thus, an activated glucocorticoid glucocorticoid receptor is to dampen inflammation receptor associated with factor A may influence a (Munck, 2005; Munck et al., 1984). This effect different set of genes than the activated glucocorti- involves, in part, the interaction of the glucocorti- coid receptor associated with factor B. Thus, the coid receptor with another protein, NFKß (e.g., Nis- glucocorticoid receptor can affect two different cel- sen & Yamamoto, 2000), which is a major lular populations in very different ways depending participant in the activation of inflammatory upon the associated cofactor. The function of the responses. The binding of the glucocorticoid recep- glucocorticoid receptor gene product depends upon tor renders NFKß unable to access its DNA targets, the context within which it operates. Indeed, the thus muting the inflammatory reaction. However, presence of other transcriptional factors can actually this glucocorticoid receptor effect is apparent only determine the direction of the effect of the glucocor- if there is an elevation in NFKß. Thus, the effects of ticoid receptor on a common genomic target, either a genomic or environmental influence would increasing transcription in one context and decreas- depend upon the immunological status of the ing it in another (e.g., Diamond, Miner, Yoshinaga, organism, further reflecting the importance of con- & Yamamoto, 1990). For example, activation of the text. The function of the glucocorticoid receptor glucocorticoid receptor promotes cell survival in gene with such protein–protein interactions further liver cells but activates (cell death) in thy- illustrates why notions of main effects of gene or mocytes. Thus, a sequence-based variation in the environment at the level of biological mechanism glucocorticoid receptor could (and indeed does) are untenable. have exactly the opposite effect on two different cel- Figure 1 outlines the organization of the gluco- lular populations. Of course, not all contextual corticoid receptor gene in the rat, which was cloned effects are quite so dramatic. But the influence of in large part by the group of Yamamoto (Miesfield such cofactors reveals a complexity that suggests et al., 1984) with a recent addition from our group that the consequences of even a highly functional in collaboration that of Jonathan Seckl (McCormick sequence variation in the glucocorticoid receptor et al., 2000). The organization of the gene is similar gene, as an example, will be determined by the cel- in rat and (Nobukuni, Smith, Hager, & lular mileu in which the glucocorticoid receptor Detera-Wadleigh, 1995; Turner & Muller, 2005). The gene is operational. Some cells would inevitably be glucocorticoid receptor gene comprised nine more affected than others. The presence and activity (an is a region of DNA that contributes of such cofactors is regulated by a variety of extra- directly to an mRNA molecule). Exons 2–9 are cod- and intracellular signals and forms an important ing regions that produce mRNA for amino acids feature of the context within which glucocorticoid sequences included in the glucocorticoid receptor receptor activity is defined. This rather complicates protein (as in DNA–RNA–protein). Exon 1 is a the study of genotype–phenotype relations and the plausibility of simple main effects of either genome or environment at the level of cell biology. Rather, such processes suggest that the effect of genomic sequence variation is understood only in terms of its interaction with environmental signals repre- sented in the cellular context. Such signals, of course, would also reflect sequence variation in other genomic regions. Since the transcriptional cofactors that influence glucocorticoid receptor activity are constantly regulated by the internal and Figure 1. A schema describing the organization of the rat external environment, the functional implications of glucocorticoid receptor gene including nine exon regions. a sequence-based variation in the glucocorticoid Note. Exons 2–9 participate in the coding of the glucocorticoid receptor protein. Exon 1 comprised multiple regions, each of receptor gene are inexorably linked to context. which is capable of activating gene transcription (i.e., The glucocorticoid receptor can also alter gene sequences). The various exon 1 promoters’ actions are tissue- expression and cellular function through processes specific, with evidence suggesting that certain promoters are that do not involve the interaction of the receptor more active in areas such as liver or thymus, and others more with DNA. Intracellular proteins interact to form active in brain (e.g., exon 17; based on McCormick et al., 2000; see Turner & Muller, 2005, for comparable data in humans). The complexes. In certain cases, such interactions inhi- consensus binding site for NGFI-A lying within the exon 17 bit the action of one of the partners. One of the pri- promoter is highlighted. NGFI-A = nerve growth factor- mary biological effects of the activated inducible factor A. Epigenetics 47 large region containing a number of nucleotide glucocorticoid receptor in one tissue while leaving sequences that are represented in the glucocorticoid it unaltered or even decreased in another. Thus, a receptor mRNA, but do not code for protein. This nucleotide-based variant in a glucocorticoid recep- exon 1 region contains a number of regulatory tor promoter region might lead to altered transcrip- elements, the activation of which alters the level of tion of the glucocorticoid receptor gene in one glucocorticoid receptor gene transcription. This tissue, while the activation of the same gene in a exon 1 region is similarly organized in humans different tissue is completely unaffected. The result- (Turner & Muller, 2005; see Zhang, Haws, & Wu, ing genotype–phenotype interaction would then 2004, for examples of similar genomic organiza- depend upon whether the trait under study is tion). In the rat, each of the individual exon 1 dependent upon a cellular population that employs segments displayed in Figure 1 contains a region the affected promoter sequence. that activates gene transcription (i.e., a transcrip- There is also variation of glucocorticoid receptor tional start site; McCormick et al., 2000) and is gene function at the level of the protein product therefore referred to as a ‘‘promoter.’’ Thus, intra- (Lu & Cidlowski, 2006). Exons 2–9 multiple cellular signals, referred to as transcription factors, variants of the glucocorticoid receptor. Two of the bind to different regions on the exon 1 promoters variants (or isoforms), the a- and b-forms of the to activate or repress the transcription of the gluco- glucocorticoid receptor, are identical for the first corticoid receptor gene. Figure 1 shows a small 727 amino acids, beyond which they differ. Gluco- segment from the exon 17 promoter for the rat glu- corticoid receptor a has an additional 50 amino cocorticoid receptor gene that binds the transcrip- acids not contained in the ß-variant. The ß-variant tion factor nerve growth factor-inducible factor A has 15 distinct amino acids. The classic effects of (NGFI-A; Crosby, Puetz, Simburger, Fahrner, & glucocorticoids on gene expression are mediated Milbrandt, 1991). through glucocorticoid receptor a. Glucocorticoid The exon 1 region contains a number of promot- receptor ß seems to bind to glucocorticoid receptor ers, and this complex level of regulation is essential a and to block its effects on gene transcription (Lu for glucocorticoid receptor function. The glucocorti- & Cidlowski, 2006). Indeed, increased expression of coid receptor gene is potentially active in virtually glucocorticoid receptor ß produces glucocorticoid every cell in the body, with different functions resistance. What is remarkable is that two seem- across various cell types. In the fetal , for ingly antagonistic proteins are actually encoded example, the activation of the glucocorticoid recep- from the same gene! Further complicating the pic- tor induces lung surfactant, a protein essential for ture is the finding that the glucocorticoid receptor the transition to a gaseous environment that marks ß-variant is produced in only a subset of cells. The birth. This effect is the basis for the glucocorticoid production of glucocorticoid receptor variants, and of premature infants. Increasing glucocorti- thus of glucocorticoid receptor gene, is entirely coid receptor function in pulmonary tissues in late dependent upon the cellular context within which fetal life is essential. Not so in the brain. Activation the gene operates. of glucocorticoid receptors in the mammalian fore- The production of glucocorticoid receptor pro- brain is associated with a decrease in tein variants is not limited to the a- and ß-isoforms. (Uno, Tarara, Else, Suleman, & Sapolsky, 1989) and More recent studies reveal at least eight different synaptic plasticity (de Kloet et al., 2008; McEwen, versions of the glucocorticoid receptor a protein, 2007). Similar catabolic effects are apparent in sev- each produced from the same gene and even the eral other tissues (Munck et al., 1984; Sapolsky same mRNA (Lu & Cidlowski, 2006). These gluco- et al., 2000). The ideal scenario is that of increasing corticoid receptor a-isoforms are differentially dis- glucocorticoid receptor expression in the lung while tributed across various tissues and differ in their muting that in and other active sites of transcriptional activity on target genes. Certain growth (Bronnegard & Okret, 1991; Sapolsky & genomic targets are regulated by all of the various Meaney, 1986). Such tissue-specific expression of glucocorticoid receptor a-isoforms, although tran- the glucocorticoid receptor gene is probably accom- scription is regulated to a greater extent by plished through the activation of different tran- some isoforms than by others. But some genes are scriptional promoters in various tissues. Indeed, regulated only by certain isoforms. In no case is this different promoters are associated with the activa- diversity in the genomic product associated with tion of the glucocorticoid receptor in lung, liver, variation in nucleotide sequence. This complexity is thymus, and brain (McCormick et al., 2000). It is an inherent feature of the glucocorticoid receptor therefore possible to increase the expression of the system. 48 Meaney

This brief review considers only the variation in Figure 1 portrays the organization of the gluco- the glucocorticoid receptor protein itself. If we fac- corticoid receptor gene. The schema is actually tor into consideration the potential for such varia- somewhat misleading. For reasons of graphic sim- tion together with the multiple cofactors, each of plicity, we often describe the organization of a gene which can alter glucocorticoid receptor function, or the interactions between transcription factors the potential for diversity in signaling becomes sub- and DNA as if the DNA were a linear molecule to stantially greater. If genotype–phenotype relations which transcription factors gain unimpeded access. are so remarkably intricate at the level of protein, The reality of protein–DNA interactions is quite dif- then imagine the complexity at the level of physiol- ferent. Figure 2 presents the classic crystallographic ogy and behavior! It is critical that we appreciate analysis of the organization of DNA (Luger, Mader, the degree to which the activity of the gene is Richmond, Sargent, & Richmond, 1997). DNA is dependent upon the cellular context within which organized into units referred to as , it functions and to appreciate that this same context each of which contains about 145–150 base pairs is also subject to the influences of sequence varia- wrapped around a core region of proteins tions in other genes that operate within the same (Turner, 2001). The and DNA together are network. Both the form and the function of the referred to as ; the is the protein product are defined by context. The func- organization of chromatin. Under normal condi- tion of the gene can only be fully understood in tions there is a tight physical relation between the terms of the cellular environment in which it oper- histone proteins and its accompanying DNA, ates. And the cellular environment, of course, is resulting in a rather closed nucleosome configura- dynamic, changing constantly as a result of signals tion. This restrictive configuration is maintained, in from other cells, including those that derive from part, by electrostatic bonds between the positively events occurring in the external environment. Ulti- charged histones and the negatively charged DNA. mately, function can only be understood in terms The closed configuration impedes transcription of the interaction between environmental signals factor binding and is associated with a reduced and the genome. level of gene expression. The activation of gene expression commonly requires chemical modifica-

Gene Transcription The most compelling evidence for the predomi- nance of Gene · Environment interactions in cellu- lar function emerges from the study of gene transcription. The transcription of a gene is a highly regulated event. At the heart of this process lies a class of proteins referred to as transcription factors (and see earlier). As the name implies, these pro- teins have the ability to bind to regulatory elements of the gene (e.g., see the exon 1 region of the gluco- corticoid receptor gene) and to activate or repress gene transcription. Importantly, the expression and ⁄ or activation of the transcription factors them- selves are dynamically regulated by environmental signals. Many of the earliest cellular responses to environmental stimuli involve either the activation of preexisting transcriptional signals through, for example, at specific sites, or the rapid synthesis of proteins (e.g., immediate early Figure 2. Crystallographic image of the nucleosome showing gene products) that then serve to regulate the activ- 146 bp wrapped around a histone complex that comprised ity of other genes. Thus, the binding of transcrip- histone 2A, 2B, 3, and 4 proteins (from Luger et al., 1997). tion factors to DNA sites is the biological Note. The tight configuration is maintained, in part, by electrostatic bonds. Modifications, such as , to the machinery of the dynamic Gene · Environment histone regulate binding and occur primarily interactions that result in altered rates of gene tran- at the histone tails protruding out of the nuclesome (pictured is scription. the tail of histone 3). Epigenetics 49 tion of the chromatin that occurs on the histone proteins.

Histone Modifications is required for increased transcription factor binding to regulatory sites on the DNA and the activation of gene expression. The dynamic alteration of chromatin structure is achieved through modifications to the histone pro- teins at the ‘‘tail’’ regions that protrude outside of the nucleosome (Figure 2). This process is achieved through a series of enzymes that bind to the histone tails and modify the local chemical properties of specific amino acids along the tail (Grunstein, 1997; Hake & Allis, 2006; Jenuwein & Allis, 2001). The Figure 3. A summary of in vivo studies with hippocampal tissue histone tails are chains of amino acids that include samples from neonates and in vitro studies using primary and arginines that are primary targets for hippocampal cell cultures. modification. For example, the enzyme histone Note. In vivo, an increased frequency of pup LG associates with acetyltransferase ‘‘transfers’’ an acetyl group onto hippocampal 5-HT turnover, activation of a 5-HT7 receptor positively coupled to cAMP and cyclic nucleotide-dependent specific lysines on the histone tails. The addition of kinases (PKA), and the induction of NGFI-A expression. In vivo, the acetyl group diminishes the positive charge, increased pup LG or artificial tactile stimulation induces NGFI-A loosening the relation between the histones and expression as well as that of the CREB-binding protein, both of DNA, opening the chromatin, and improving the which show greater binding to the exon 17 promoter in the ability of transcription factors to access DNA sites. neonatal offspring High- compared with low-LG mothers. Results of in vitro studies show that blockade of cAMP, PKA, or NGFI-A Thus, histone acetylation at specific sites is abolish the effect of 5-HT of glucocorticoid receptor expression. commonly associated with active gene transcrip- cAMP = concentration of adenosine 3,5-monophosphate; CREB- tion. The functional antagonists of the histone binding protein = cAMP-response element-binding protein; acetyltransfereases are a class of enzymes known as 5-HT = 5-hydroxytryptamine; LG = licking ⁄ grooming; NGFI- histone deacetylases (HDACs). These enzymes A = nerve growth factor-inducible factor A; PKA = protein kinase A. remove acetyl groups and prevent further acetyla- tion, thus serving to maintain a closed chromatin structure, decreasing transcription factor and gene binds NGFI-A abolishes the ability of NGFI-A to expression. The reader should note that there are associate with the exon 17 promoter and eliminates actually multiple modifications to histone tails the effect of NGFI-A on gene transcription (Weaver including (in this case on the histones), et al., 2007). However, the ability of NGFI-A to bind phosphorylation, ubiquitination, and so on. For the to the exon 17 promoter is regulated by another sake on simplicity, the discussion is limited to his- protein, a cofactor, the concentration of adenosine tone acetylation or deacetylation. 3,5-monophosphate (cAMP)-response element-bind- ing (CREB) protein, that is, activated by the same 5-HT–cAMP ⁄ PKA (protein kinase A) signaling Regulation of Glucocorticoid Receptor Expression cascade (Figure 3). The CREB-binding protein is a Figure 3 summarizes the effect of the neurotrans- histone acetyltransferase. The resulting increase in mitter serotonin (5-hydroxytryptamine [5-HT]) on the association of the CREB-binding protein with the glucocorticoid receptor gene transcription in hippo- exon 17 promoter is accompanied by an increase in campal neurons (Mitchell, Betito, Rowe, Boksa, & the acetylation of a specific lysine on the tail of his- Meaney, 1992; Mitchell, Rowe, Boksa, & Meaney, tone 3 of the exon 17 promoter (Weaver et al., 2004; 1990; Weaver et al., 2007). This effect is dependent Weaver et al., 2007). Thus, 5-HT activates both upon the binding of the transcription factor NGFI-A NGFI-A and the CREB-binding protein, which to a specific binding site on the exon 17 GR promoter. appear to associate with one another prior to DNA We can precisely define the importance of this inter- binding. The CREB-binding protein acetylates hi- action. For example, mutating a single nucleotide, in stones associated with the exon 17 promoter, enhanc- this case simply exchanging a for an ade- ing the ability of NGFI-A to bind and activate gene nine, in the region of the promoter that normally transcription. 50 Meaney

Serotonin is a classic neurotransmitter that relates of the 5-HTTP promoter polymorphism as responds dynamically to environmental signals. well as subsequent meta-analyses (Sen, Burmeister, This effect reflects the dependence of gene tran- & Ghosh, 2004; Schinka, Busch, & Robichaux- scription on signals that derive from environmental Keene, 2004; but also see Munafo, Clark, & Flint, events (note the relevant environmental event may 2005) reveal a statistical association with neuroti- be internal or external to the organism; e.g., a cism, which is elevated among individuals that change in the availability of glucose, an electrical carry at least one copy of the S . Since neuroti- impulse, or a social interaction). Thus, the conse- cism predicts an increased risk for affective disor- quences of a variation in DNA sequence for pheno- ders (Hettema, Neale, Myers, Prescott, & Kendler, type will depend upon environmental signals. The 2006; Kendler, Kuhn, & Prescott, 2004), it is not sur- genomic variant is relevant for phenotype only to prising that the 5-HTTP promoter polymorphism is the extent that the gene is actually expressed and at least weakly associated with disorders of emo- participating in cellular function. If a particular tis- tional function, namely, depression and anxiety sue were to reveal a genomic variant in a region of (Canli & Lesch, 2007; Hariri, Emily, Drabant, & the DNA that is transcriptionally inactive, then the Weinberger, 2006). implications for phenotype are minimal. The degree Affective disorders associate with altered activity to which a genomic variant influences phenotype in the amygdala (Davis, 2006; Gorman, Kent, Sulli- will depend upon the cellular context. Likewise, to van, & Coplan, 2000; LeDoux, 2000; Meaney, Le- the extent that an inherited genomic variant influ- Doux, & Leibowitz, 2008; Pe´rez-Edgar & Fox, 2005), ences the capacity for transcription, then the influ- a primary brain structure for emotional states of ence of the environmental signal on phenotype will fear as well as a major target for 5-HT projections be affected by variation in nucleotide sequence. from the raphe´ nuclei. The expression of fear in Such considerations showcase the interdependence rodents is inhibited by 5-HT activity in the hippo- of gene and environment. campus and amygdala (Stutzmann & LeDoux, 1999), and bearing of 5-HT1A receptors show enhanced fearfulness (Gross & Hen, 2004). Aversive stimuli activate the amygdala in Gene · Environment Interactions: The Case of the humans (LaBar et al., 1995; LaBar et al., 1998; Mor- 5-HT Transporter ris, Ohman, & Dolan, 1998) and individual differ- One of the best-studied genomic variations in bio- ences in trait anxiety predict the response of the logical psychiatry is that of the polymorphism in amygdala to threatening stimuli (Etkin, Egner, Per- the promoter region of the SLC6A4 gene that aza, Kande, & Hirsch, 2005; Etkin et al., 2004). Indi- encodes for the 5-HT transporter (5-HTTP). The viduals carrying an S variant of the 5-HTTP synaptic activity of 5-HT is terminated by reuptake promoter polymorphism show increased activation into presynaptic terminals, which occurs via the of the amygdala while processing fearful or angry 5-HTTP protein. Serotonin is a critical neurotrans- faces (Hariri et al., 2002; Hariri et al., 2005; Pezawas mitter in the regulation of emotional states (Lucki, et al., 2005). Similar effects of the 5-HTTP promoter 1998) and a major target for medications designed polymorphism are obtained in subjects during pro- to treat a range of affective disorders (Blier & de cessing of negative relative to neutral pictures Montigny, 1999; Gross & Hen, 2004). The gene for (Heinz et al., 2005) or words (Canli et al., 2006). the 5-HTTP in humans shows a relatively common Moreover, the S allele of the 5-HTTP promoter is polymorphism characterized by a variable repeat associated with increased activity at rest in both the sequence in the promoter region resulting in two amygdala and the hippocampus (Canli & Lesch, common : the short (S) variant comprising 14 2007). copies of a 20–23 base-pair repeat unit, and the long An additional feature of anxiety is that of (L) variant comprising 16 copies (Lesch et al., 1996). decreased inhibitory prefrontal regulation of The 5-HTTP promoter sequence polymorphism evoked activity in the amygdala (Shin et al., 2004, associates with differential transcription of the 5- 2005). Affective disorders associate with decreased HTTP gene, with more efficient expression from the volume in the subgenual region of the anterior cin- L allele (Greenberg et al., 1999; Lesch et al., 1996). gulate cortex (Drevets et al., 1997), which projects More recent studies suggest that further variation to and regulates activity in the amygdala (Maren lying within the L allele also affects 5-HTTP gene & Quirk, 2004). Gray matter volume in the sub- function, resulting in subtypes of the L allele (Hu genual anterior is reduced in et al., 2005). The early studies of the functional cor- carriers of the S variant of the 5-HTTP promoter Epigenetics 51 polymorphism. In addition, individuals bearing the suggesting that the consequences of 5-HT altera- S promoter variant show decreased functional con- tions appear in early life. Here again, however, nectivity between the amygdala and anterior cingu- there is environmental modulation of the geno- late cortex (Pezawas et al., 2005), which could serve type–phenotype relation. Fox et al. (2005) reported as the basis for the increased amygdala response to that the quality of the maternal environment inter- negative stimuli. acted with the 5-HTTP promoter genotype to deter- Since genomic variations are potentially opera- mine behavioral inhibition in children. tive at any time over the lifespan, it is interesting to Fox et al.’s (2005) report is representative of a consider the possibility that the 5-HTTP polymor- growing body of studies revealing Gene · Environ- phism could alter 5-HT activity during periods of ment interactions in emotional and cognitive devel- neural development. Thus, Pezawas et al. (2005) opment. Suomi, Bennett, and colleagues described proposed that the association between the 5-HTTP a brilliant example of such gene–environment inter- promoter polymorphism and vulnerability to dependence in the rhesus monkey (Bennett et al., depression reflects a developmental process that 2002; Champoux et al., 2002; Suomi, 2006). The rhe- affects the structural connectivity and functional sus monkey shows a polymorphism in the 5-HTTP interactions within a neural circuit that regulates promoter that is comparable in form and function emotional reactivity and fear. They further pro- to that in humans such that the S 5-HTTP promoter posed that these functional deficits could be exacer- variant in the monkey is also associated with bated by environmental adverse experiences, decreased 5-HTTP levels in brain (Lesch et al., possibly through an impaired capacity to regulate 1997). Likewise, the 5-HTTP promoter polymor- affective states during periods of adversity. Such an phism affects both 5-HT metabolism and emotional effect would be expected to result in an interaction function in the monkey. Monkeys bearing the S ver- between gene and environment that renders indi- sion of the 5-HTTP promoter polymorphism show viduals vulnerable to affective disorders. There is reduced 5-HT activity, increased , and considerable evidence in support of this develop- are more aggressive than animals bearing the mental hypothesis. Studies with rodent models longer version of the promoter variant (Bennett reveal a profound influence of 5-HT activity on et al., 2002; Champoux et al., 2002). However, the neuronal development, including the sprouting of genotype–phenotype relation is strongly influenced axonal processes. Hen and colleagues (Gross et al., by the rearing environment. The phenotypic differ- 2002) showed that a null of the 5-HT1A in ences described earlier are apparent among animals mice associates with increased fearfulness. This separated from their mothers and reared in peer– knockout model, like that of human sequence poly- peer groups. In contrast, among animals reared by morphisms, reveals effects from conception onward their mothers, the differences in phenotype are to the time of behavioral assessment. It is therefore actually slightly reversed such that animals with unclear as to whether the associated change in phe- the S variant of the 5-HTTP promoter show mod- notype reflects a developmental influence, an effect estly greater 5-HT activity and are less impulsive on neurotransmission at the time of assessment, or and aggressive than those carrying the longer pro- both. Hen’s lab (Gross & Hen, 2004; Gross et al., moter variant. In these studies, the genotype–phe- 2002) resolved this issue by producing a conditional notype relations depended completely upon the knockout model in which the gene that encodes for rearing environment. Similar effects are apparent in the 5-HT1A receptor is rendered inoperative only the mouse where there is an interaction between a during selected periods over the lifespan of the disruption of the 5-HTTP gene and the quality of mouse. The findings reveal that the condition in maternal care on neural systems associated with which the 5-HT1A receptor gene is muted for only fear behavior (Carola et al., 2008). during the first 3 weeks of life completely recapitu- A series of landmark studies (e.g., Caspi & Moffit, lates the increased fearfulness observed in animals 2006; Caspi et al., 2003) provide a parallel set of find- in which the 5-HT1A receptor was knocked out for ings in human populations. In these studies, as the entire life of the animal. These findings suggest shown in previous research, childhood maltreat- that genomic variants that alter 5-HT function can ment increases the probability of an episode of act during early development to regulate neural major depression. However, this effect is only circuitry that mediates the expression of fear. apparent among individuals that bear at least one Moreover, variants of the gene for the 5-HTTP are copy of the S 5-HTTP promoter allele. Individuals associated with negative temperament in infancy homozygous for the L promoter variant show no (Auerbach et al., 2002; Ebstein, 2003, 2006), further increase in the probability of depression despite the 52 Meaney experience of childhood maltreatment (see Caspi & earlier, animals bearing the L variant of the 5-HTTP Moffit, 2006, for a review of other, compelling exam- promoter polymorphism are markedly less sensitive ples of such Gene · Environment interactions). A to peer versus maternal rearing than those carrying more recent study shows evidence for the same the S promoter variant (at least with respect to the Gene · Environment interaction in defining individ- measures used in these studies). The same argument ual differences in neural functions that associate could be made for the Caspi et al. (2003) findings with depression (Canli et al., 2006). There is also evi- that reveal an increased risk for depression associ- dence of a comparable Gene · Environment effect ated with child maltreatment among those bearing examining polymorphisms in other genes impli- S ⁄ S 5-HTTP promoter variants. Indeed, one might cated in 5-HT neurotransmission (Jokela et al., well imagine that certain genomic variants alter 2007). In this case, the polymorphism in question is neuronal function of the child, affecting the a SNP in the 5HTR2A gene that encodes for the response to environmental conditions. Studies 5-HT2A receptor that is also linked to mood disor- examining the children with a polymorphism on ders. The authors found a significant interaction the DRD4 gene that encodes a receptor for the neu- between the 5HTR2A genotype and the quality of rotransmitter dopamine bear on this point. This maternal care in defining the level of depressive genomic variant is associated with differences in symptoms. attachment (Gervai et al., 2007; Lakatos et al., 2002) The Caspi et al. (2003) finding of an interaction as well as infant responses to novelty (Lakatos between the promoter polymorphism and either life et al., 2003). The DRD4 polymorphism interacts stressors or childhood maltreatment has been repli- with the quality of parenting to determine the level cated (Kaufman et al., 2004; Kendler, Kuhn, Vittum, of internalizing and externalizing behavior in chil- Prescott, & Riley, 2005), and other studies suggest dren (Propper, Willoughby, Halpern, Carbone, & further moderation by gender (Eley et al., 2004; Cox, 2007). Moreover, the same polymorphism Grabe et al., 2005) or social support (Kaufman et al., determines the effect of a parent-training interven- 2004). Two studies (Gillespie, Whitfield, Williams, tion targeting maternal sensitivity on the level of Heath, & Martin, 2005; Surtees et al., 2006) failed to externalizing behavior in children (Bakermans- replicate this Gene · Environment interaction effect. Kranenburg, Van IJzendoorn, Pijlman, Mesman, & However, we need to appreciate that an interaction Juffer, 2008). In these studies, the genotype of the between a single genomic variant and a broadly child determined the response to parenting. Like- characterized environmental condition at one stage wise, considering the evidence for the effects of of life is still a rather narrow focus on the complex genomic variants on emotional functions in chil- pathophysiology of mental disorders. Genes encode dren, genomic variation might operate within spe- for RNAs and proteins that operate in complex cific familial settings (e.g., Fox et al., 2005) to networks within pathways. Such influence the behavior of the child and thus the networks form the basis for the interaction of gene quality of the interactions with caregivers. The products that are then revealed as gene–gene inter- 5-HTTP polymorphism associates with differences actions in studies of genomic variations, including in infant temperament and reactions to novelty data sets from genome-wide association studies. A (Auerbach et al., 1999; Ebstein, 2006; Lakatos et al., major advance in genetics is the emergence of com- 2003), which in turn predicts sensitivity to parental putational approaches that apply a biologically influences (Belsky, 1997). Although the point is less informed, statistical analysis of the interactions emphasized in this article, the Gene · Environment between functionally related genes. Such bioinfor- perspective argues no less forcefully that environ- matic approaches feature the description of gene mental effects must be considered within the geno- networks, involving multiple genomic targets and mic context of the individual. The interdependence their variations. We can thus anticipate studies of of gene and environment is bidirectional. Gene Network · Environment interactions. These findings underscore the importance of Arguments for the interdependence of gene and Gene · Environment interactions, and the studies environment should be familiar to those in devel- with the rhesus monkey suggest that the develop- opmental psychology. Belsky (1997) and Rutter mental consequences of such interactions persist (2007) argue from the other point of view that envi- into adulthood. This consideration simply reflects ronmental forces, such as familial influences, will the long-standing interest of on not affect all children equally and that an obvious the potential influence of early experience (Gott- source of differential sensitivity lies in genomic lieb, 1991, 1997), particularly those involving the variation. In the nonhuman primate studies cited interaction of the parent and child, on phenotypic Epigenetics 53 development. In the preceding sections, we consid- considerably greater risk for mental illness, as well ered the molecular mechanisms that underlie Gene · as for , gastrointestinal illness, , and Environment interactions with respect to immediate heart disease. ‘‘Stress diathesis’’ models are pro- and transient changes in gene expression. But what posed as explanations for the relation between early are the mechanisms that might account for the experience and health (e.g., Gorman et al., 2000; enduring effects of environmental conditions on Heim & Nemeroff, 2001; McEwen, 2003; Meaney, gene expression? How might parental influences sta- 2001; Meaney et al., 2007; Repetti et al., 2002; Seckl bly alter phenotype? Could such effects involve the & Meaney, 1994). These models suggest that adver- ‘‘programming’’ of gene expression? How might such sity in early life alters the development of neural stable influences occur? A response to such questions and endocrine responses to stress in a manner that requires an understanding of the physical properties predisposes individuals to disease. Adversity or of Gene · Environment interactions that produce decreased quality of parental investment appears to enduring effects on brain function. directly increase the magnitude of emotional, auto- nomic, and endocrine responses to stress (collec- tively referred to here as defensive responses). Diathesis describes the interaction between devel- Environmental Programming of Gene Expression opment, including the potential influence of genetic Studies in developmental psychobiology and physi- variations at the level of sequence, and the prevail- ology are replete with examples of the environmen- ing level of stress in predicting health outcomes. tal programming of gene expression. Such studies Such models have considerable appeal and could commonly report that a variation in the early envi- identify both the origins of illness and the nature of ronment associates with changes in gene expression the underlying vulnerability (as well as resilience). and function that persist into adulthood and thus Moreover, such models have the virtue of provi- well beyond the duration of the relevant environ- ding a logical bridge between psychosocial mental event. In the rat, for example, prenatal (e.g., family function) and biological (endocrine nutrient deprivation or enhanced exposure to hor- responses to stressors) levels of analyses in under- monal signals associated with stress can stably standing health outcomes. alter, or program, the transcription of genes in the A critical assumption in the stress diathesis mod- liver and other sites that are associated with glu- els is that the increased expression of defensive cose and fat metabolism, including the gene for the responses endangers health. The idea has consider- glucocorticoid receptor (Bateson et al., 2004; able support in clinical physiology (Chrousos & Gluckman & Hanson, 2004, 2007; Meaney, Szyf, & Gold, 1992; Dallman et al., 2005; McEwen, 2007). Seckl, 2007; Seckl & Holmes, 2007). These findings Defensive responses to stress are adaptive. Neural are assumed to represent instances in which the signals associated with the of the stres- operation of a genomic region in adulthood varies sor increase the release of stress hormones into the as a function of early environmental influences. The bloodstream, including glucocorticoids from the results of recent studies suggest that such ‘‘pro- adrenal gland and catecholamines, particularly nor- gramming’’ effects can derive from Gene · Envi- epinephrine, from the sympathetic nervous system. ronment interactions in early life that lead to a The combined actions of these hormones increase structural alteration of the DNA, which in turn the availability of energy substrates, such as those mediates the effects on gene expression as well as derived from fat and glucose metabolism. Such more complex levels of phenotype (Meaney & Szyf, effects maintain the normal cellular function and 2005; Meaney et al., 2007). These studies have been organ efficiency. These actions protect against performed in rodents but were inspired by the vast catastrophes such as hypotensive shock. The release literature reporting the pervasive effects of family of catecholamines in the brain increases vigilance environment on health outcomes in humans. and provokes states of fear, and enhances avoid- Epidemiological studies reveal the importance of ance and fear conditioning, which can family function and early life events as predictors serve to reduce the chances of further encounters of health in adulthood (Felitti et al., 1998; Leserman with the same conditions. et al., 1996; Lissau & Sorensen, 1994; McCauley Defensive responses are the logical outcome of et al., 1997; Repetti, Taylor, & Seeman, 2002; Russak the stress-induced changes in activity in the brain & Schwartz, 1997; Seckl, 2001). As adults, victims of and endocrine organs, and are defined by increases childhood physical or sexual abuse, emotional in the synthesis and release of glucocorticoids and neglect and harsh, inconsistent discipline are at catecholamines (Dallman et al., 1995; Munck et al., 54 Meaney

1984; Rosen & Schulkin, 1998; Sapolsky et al., 2000). natal handling of rat pups does indeed increase the However, there is a cost associated with persistent licking ⁄ grooming (LG) of pups by the mother (e.g., activation of these same responses: chronically Lee & Williams, 1977; Liu et al., 1997). Pup LG is a enhanced emotional arousal, sustained increases in major source of tactile stimulation for the neonatal blood sugars and fats, possible hyperinflammation, rat that regulates endocrine and cardiovascular disruption of sleep and normal cognitive function, function in the pup (Hofer, 2005; Levine, 1994; among others (Chrousos & Gold, 1992; Dallman Schanberg, Evoniuk, & Kuhn, 1984). The question et al., 2005; McEwen, 2007; Sapolsky et al., 2000; then was whether such variations in pup LG might Walker, 2007). Thus, chronic activation of defensive directly alter the development of individual differ- responses can indeed predispose individuals to ences in defensive responses. illnesses such as diabetes, heart disease, mood dis- Subsequent findings revealed considerable evi- orders, and so on. Not surprisingly, studies in dence for the effect of maternal care on the behav- behavioral reveal that individuals with ioral and endocrine responses to stress in the enhanced stress reactivity are at greater risk for offspring. One approach was to simply examine the such forms of chronic illness. However, it is also consequences of naturally occurring variations in important to note that insufficient activation of pup LG among lactating rats independent of any defensive responses under conditions of threat also experimental manipulation. Among lactating rats compromises health and is associated with chronic there are considerable individual differences in the fatigue, chronic pain, posttraumatic stress disorder, frequency of pup LG that are stable over the repro- and hyperinflammation (Munck et al., 1984; Raison ductive lifetime of the female (Champagne, Francis, & Miller, 2003; Yehuda & Bierer, 2008). We walk a Mar, & Meaney, 2003). The results of longitudinal fine line here. And this underscores the importance studies are consistent with the Levine (1970) of an appropriate level of stress reactivity for the hypothesis. The male or female adult offspring of individual, one sufficient to ensure the maintenance mothers that naturally exhibit increased levels of of function during adversity, but not so excessive pup LG (i.e., high-LG mothers) show more modest as to promote chronic illness. What is appropriate, behavioral and endocrine responses to stress com- of course, will vary depending upon the prevailing pared to animals reared by low-LG mothers (Caldji level of environmental demand. There is no single, et al., 1998; Francis, Diorio, Liu, & Meaney, 1999; ideal level of stress reactivity across all populations. Liu et al., 1997; Menard, Champagne, & Meaney, 2004; Toki et al., 2007; Weaver et al., 2004; Zhang et al., 2006). Specifically, the offspring of high-LG Individual Differences in Defensive Responses mothers show reduced fearfulness and more mod- In the late 1950s and early 1960s, psychologists est HPA responses to stress. Cross-fostering stud- Gig Levine and Victor Denenberg (Denenberg, 1964; ies, where pups born to high-LG mothers are Levine, 1970) reported that postnatal handling (or fostered at birth to low-LG mothers (and vice infantile stimulation) of infant rodents decreased versa), suggest a direct relation between maternal the magnitude of both behavioral and hypotha- care and the postnatal development of individual lamic–pituitary–adrenal (HPA) responses to stress differences in behavioral and HPA responses to in adulthood. These findings demonstrated the stress (Caldji, Diorio, & Meaney, 2003; Caldji, Fran- influence of the early environment over the develop- cis, Sharma, Plotsky, & Meaney, 2000; Francis et al., ment of rudimentary defensive responses to threat. 1999; Weaver et al., 2004). In these studies, the rear- The importance of such findings in this era should ing mother determined the phenotype of the off- not be underestimated. This was a period when spring. Thus, variations within a normal range of defensive responses were considered as ‘‘innate,’’ parental care can dramatically alter phenotypic developing well outside the realm of experiential development in the rat. influence. Levine and others later suggested that the The effects of maternal care on the development effects of handling were actually mediated by of defensive responses to stress in the rat involve changes in maternal care. Thus, handling of the alterations in the function of the corticotrophin- pups was to alter the behavior of the releasing factor (CRF) systems in selected brain mother, which was then critical for the ‘‘handling’’ regions (Figure 4). The CRF system furnishes the effects. The handling paradigm involves exposure critical signal for the activation of behavioral, of the neonate to a complex set of stimuli, including emotional, autonomic, and endocrine responses that of a novel physical environment (Tang, Akers, to stressors (Bale & Vale, 2004; Koob, Heinrichs, Reeb, Romeo, & McEwen, 2006). Nevertheless, post- Menzaghi, Pich, & Britton, 1994; Plotsky, Cunning- Epigenetics 55

Figure 4. A schema outlining the function of the hypothalamic–pituitary–adrenal axis, the nexus of which are the corticotropin- releasing factor (CRF) neurons of the paraventricular nucleus of the hypothalamus. Note. CRF is released into the portal system of the anterior pituitary stimulating the synthesis and release of adrenocorticotropin (ACTH), which then stimulates adrenal glucocorticoid release. Glucocorticoids act on glucocorticoid receptors in multiple brain regions, including the hippocampus, to inhibit the synthesis and release of CRF (i.e., glucocroticoid negative feedback). The adult offspring of high-LG mothers, by comparison to those of low-LG dams, show (a) increased glucocorticoid receptor expression, (b) enhanced negative-feedback sensitivity to glucocorticoids, (c) reduced CRF expression in the hypothalamus, and (d) more modest pituitary–adrenal responses to stress. LG = licking ⁄ grooming. ham, & Widmaier, 1989). As adults, the offspring of the pup with a brush, provide direct evidence for high-LG mothers show decreased CRF expression the importance of tactile stimulation derived from in the hypothalamus, as well as reduced plasma pup LG. Thus, stroking pups over the 1st week of adrenocorticotropin (ACTH) and glucocorticoid life increases hippocampal glucocorticoid receptor responses to acute stress by comparison to the adult expression (Jutapakdeegul, Casalotti, Govitrapong, offspring of low-LG mothers (Francis et al., 1999; & Kotchabhakdi, 2003) and dampens behavioral Liu et al., 1997; Weaver et al., 2004; Weaver et al., and HPA responses to stress (Burton et al., 2007; 2005). Circulating glucocorticoids act at glucocorti- Gonzalez, Lovic, Ward, Wainwright, & Fleming, coid receptor sites in corticolimbic structures, such 2001). Likewise, manipulations of lactating mothers as the hippocampus, to regulate HPA activity (Fig- that directly increase the frequency of pup LG also ure 4). Such feedback effects commonly inhibit increase hippocampal glucocorticoid receptor hypothalamic CRF expression. The high-LG off- expression and decrease HPA responses to stress spring showed significantly increased hippocampal (Francis et al., 1999; Toki et al., 2007). glucocorticoid receptor expression, enhanced gluco- The offspring of the high- and low-LG mothers corticoid negative-feedback sensitivity, and also differed in behavioral responses to stress decreased hypothalamic CRF levels. Indeed, the (Caldji et al., 1998; Caldji et al., 2003; Francis et al., magnitude of the glucocorticoid response to acute 1999). As adults, the offspring of the high-LG stress was significantly correlated with the fre- mothers showed decreased startle responses, quency of pup LG during the 1st week of life, as substantially less fearfulness in the presence of was the level of both hippocampal glucocorticoid stressors, such as novel environments. Moreover, receptor and hypothalamic CRF expression (all active defensive responses, such the burying of rs > .70; Liu et al., 1997). Importantly, pharmaco- threatening stimuli, are more conspicuous among logical manipulations that block the effect of the the adult offspring of low-LG mothers (Menard glucocorticoid receptor eliminate the maternal et al., 2004). While such pervasive effects are no effect on the HPA response to stress, suggesting doubt associated with alterations in multiple neural that the differences in hippocampal glucocorticoid systems, there are robust effects of maternal care on receptor expression are directly related to those at the central CRF systems, including those that lie the level of HPA function. outside of the hypothalamus. The activation of fear Pup LG is a major source of tactile stimulation behavior corresponds to an increase in CRF release for the neonate. Experimental models that directly from the amygdala and bed nucleus of the stria apply tactile stimulation, through the stroking of terminalis onto catecholaminergic cell bodies in the 56 Meaney coeruleus. CRF acts at CRF1 receptors in the the inhibitory function of the GABAA receptor. The locus coeruleus to stimulate the release of norepi- adult offspring of high-LG mothers show signifi- nephrine in a variety of corticolimbic structures cantly increased expression of both the a1 and g2 (Bale & Vale, 2004; Valentino, Curtis, Page, Pavco- subunits. The effect is almost unique to the amyg- vich, & Florin-Lechner, 1998). Gamma-aminobutyric dala and is reversed with cross-fostering (Caldji acid type A (GABA), which is the primary source et al., 2003). of neural inhibition in the adult mammalian brain, The results of these studies suggest that the dampens the activation of this CRF–catecholamine behavior of the mother toward her offspring can connection during stress. The offspring of the high- ‘‘program’’ stable changes in gene expression that LG mothers show decreased CRF1 receptor levels in then serve as the basis for individual differences in the locus coeruleus and increased GABAA ⁄ benzodi- behavioral and neuroendocrine responses to stress azepine receptor levels in the basolateral and cen- in adulthood. The maternal effects on phenotype tral nucleus of the amygdala, as well as in the locus are associated with sustained changes in the coeruleus and decreased CRF expression in the expression of genes in brain regions that mediate amygdala (Caldji et al., 1998; Caldji et al., 2003). responses to stress, and form the basis for stable The adult offspring of high-LG mothers are more individual differences in stress reactivity. These sensitive to the inhibitory effects of benzodiaze- findings provide a potential mechanism for the pines on fear behavior (Fries, Moragues, Caldji, influence of parental care on vulnerability ⁄ resis- Hellhammer, & Meaney, 2004). Receptors for the tance to stress-induced illness over the lifespan. benzodiazepines (e.g., diazepam) are part of the However appealing, this hypothesis has yet to be GABAA receptor complex, and the anxiolytic effects directly confirmed. But the critical issue for this of these compounds occur through an ability to article is simply that of how maternal care might enhance the effect of GABA at the GABAA receptor stably affect gene expression. How are variations in site. GABAA, or benzodiazepine receptor agonists, the social interactions between the mother and her suppress CRF expression in the amygdala, thus offspring ‘‘biologically embedded’’ so as to stably reducing activity within the amygdala–locus coeru- alter the activity of specific regions of the genome? leus and decreasing norepinephrine responses to The answers to these questions appear to involve stress. Predictably, stress-induced increases in nor- the ability of social interactions in early develop- epinephrine that are normally stimulated by CRF ment to actually modify the structure of the rele- are significantly higher in the offspring of the low- vant genomic regions. LG offspring. The increased release of norepineph- rine is consistent with the enhanced fearfulness of these animals. Thus, increased pup LG is associated Epigenetic Regulation of the Genome with the enhanced efficacy of systems that normally serve to inhibit the expression and actions of CRF. When we think of genomic influences we most com- These systems include the hippocampal glucocorti- monly imagine effects associated with variation in coid receptor and the GABAA receptor in the amyg- nucleotide sequence—the genetic code. Yet, this is dala. Both effects involve sustained alterations in only one form of information contained on the gene expression as a function of maternal care. DNA. Despite the reverence afforded DNA, a gene The complexity of such maternal effects on gene is basically like any other molecule in the cell; it is expression is apparent in the alterations in GABAA subject to physical modifications. These modifica- receptor function. The GABAA receptor is a multi- tions alter the structure and chemical properties of protein complex that comprised five individual the DNA, and thus gene expression. Collectively, subunits, each of which is a product of a distinct the modifications to the DNA and its chromatin genomic region. An interesting feature to this sys- environment can be considered as an additional tem is that there are at least 19 different subunits layer of information that is contained within the that can be employed in the formation of a GABAA genome. This information is thus epigenetic in nature receptor. The activity of the receptor is determined (the name derives from the Greek epi meaning by its subunit composition. For example, the inclu- ‘‘upon’’ and genetics). The acetylation of the histone sion of an a1 subunit confers an increased affinity proteins referred to earlier is one example of an epi- of the receptor for GABA, enhancing GABAergic genetic modification. Epigenetic modifications do activity at the GABAA receptor. Others, such not alter the sequence composition of the genome. as the g2 subunit, will define the presence of a Instead, epigenetic marks on the DNA and the other benzodiazepine binding site that further increases features of the chromatin regulate the operation of Epigenetics 57 the genome. Thus, epigenetics has been defined as a while that emanating from the father is silenced—a functional modification to the DNA that does not ‘‘maternally imprinted gene.’’ In other cases, it is involve an alteration of sequence. While this defini- the reverse; the copy of the gene inherited from the tion has recently been subjected to revision (Bird, father that is active, while that from the mother is 2007; Hake & Allis, 2006), the essential features of silenced—a ‘‘paternally imprinted gene.’’ The silent epigenetic mechanisms are (a) structural modifica- copy is methylated in DNA regions that regulate tions to chromatin either at the level of the histone gene expression and thus inactive. Again, the epi- proteins (Figure 2) or the DNA, (b) regulation of the genetic marks associated with gene imprinting are structure and function of chromatin, (c) affects on established very early in life. These marks, as well gene expression, and (d) that these effects occur in as those associated with X- inactiva- the absence of any change in nucleotide sequence. tion, are stable, leaving researchers in the field with The functional byproduct of the epigenetic modifica- the impression that under normal conditions DNA tions is that of a change in gene transcription. methylation occurs early in embryonic life and is The classic epigenetic alteration is that of DNA largely irreversible. Indeed, it was commonly methylation, which involves the addition of a thought that DNA methylation was an actively methyl group onto in the DNA (Bird, dynamic process only during periods of cell divi- 1986; Holliday, 1989; Razin & Cedar, 1993; Razin & sion such that in mature, postmitotic cells further Riggs, 1980). The methylation of DNA is an active alteration of methylation patterns was improbable. biochemical modification that in selec- Moreover, the loss of cytosine methylation in such tively targets cytosines and is achieved through the models is associated with profound . This actions of a class of enzymes, DNA methyltransfe- perspective was further reinforced by findings rases, which transfer the methyl groups from showing that an alteration of DNA methylation at methyl donors. There are two critical features to critical genomic targets (i.e., tumor suppressors) is DNA methylation: First, it is a stable chemical mod- associated with (Eden, Gaudet, Waghmare, & ification, and second, it is associated with the Jaenisch, 2003; Feinberg, 2007; Laird, 2005). silencing of gene transcription (Bestor, 1998; Bird, While these assumptions concerning DNA meth- 2002; Bird & Wolffe, 1999; Razin, 1998). ylation appear valid for the examples cited earlier, Until recently, DNA methylation patterns on the recent studies reveal that DNA methylation pat- genome were thought to be overlaid upon the gen- terns are actively modified in mature (i.e., fully ome only during early periods in embryonic devel- differentiated) cells, including, and perhaps opment. This belief was derived in part from the especially, neurons and that such modifications can experimental models commonly used to study occur in animals in response to cellular signals DNA methylation. DNA methylation-induced gene driven by environmental events (Bird, 2007; Jirtle & silencing mediates two of the most commonly Skinner, 2007; Meaney & Szyf, 2005). For example, studied examples of epigenetic silencing, namely, variations in the diet of mice during gestation or X-chromosome inactivation and gene imprinting. later in development, such as the early postwea- Mammalian females bear two copies of the X-chro- ning period, can stably alter the methylation status mosome. The inactivation of one copy of the of the DNA (Cooney, Dave, & Wolff, 2002; X-chromosome occurs in all mammalian females Waterland & Jirtle, 2003; Waterland, Lin, Smith, & and is essential for normal function (i.e., maintain- Jirtle, 2006; Whitelaw & Whitelaw, 2006). Likewise, ing a constant gene dosage in males and females). both mature lymphocytes (Bruniquel & Schwartz, The silencing of the X-chromosome is associated 2003; Murayama et al., 2006) and neurons (e.g., with DNA methylation (Mohandas, Sparkes, & Champagne, 2008; Champagne et al., 2006; Lubin, Shapiro, 1981; Riggs & Pfeifer, 1992; see Hellman & Roth, & Sweatt, 2008; Martinowich et al., 2003; Chess, 2007, for a more current update). The second Sweatt, 2009) show changes in the DNA methyla- example of epigenetic-mediated is tion patterns at critical genomic regions in response that of gene imprinting (da Rocha & Ferguson- to environmental stimuli that stably alter cellular Smith, 2004; Reik, 2001), a remarkable subject in its function. The ability of environmental signals to own right, and one with considerable implications actively remodel epigenetic marks that regulate for growth and development (Charalambous, da gene expression is a rather radical change in our Rocha, & Ferguson-Smith, 2007). For humans and understanding of the environmental regulation of other mammals, the expression-specific genes are gene expression. Such epigenetic modifications are determined by the parent of origin. For certain thus a candidate mechanism for the environmental genes, the copy derived from the mother is active, ‘‘programming’’ of gene expression. 58 Meaney

ences in the frequency of pup LG between high- DNA Methylation and Gene Transcription and low-LG mothers are limited to the 1st week of DNA methylation is associated with the silencing postnatal life. And yet the differences in gene of gene transcription. This effect appears to be med- expression and neural function are apparent in iated in one of two ways (Bird, 2002). First, wide adulthood. How might the effects of an essentially swaths of DNA can be methylated and the shear social interaction stably alter the expression of the density of methylation precludes transcription fac- glucocorticoid receptor gene? tor binding to DNA sites, thus silencing gene The focus of the epigenetic studies is the NGFI-A expression. The second manner is subtler, and consensus sequence in the exon 17 promoter probably far more prevalent, in regions with more (Figure 1) that activates glucocorticoid receptor dynamic variations in gene transcription, such as expression in hippocampal neurons. The tactile the brain. In this case, selected cytosines are methy- stimulation associated with pup LG increases 5-HT lated and the presence of the methyl group attracts metabolism in the hippocampus. As described ear- a class of proteins know as methylated-DNA bind- lier (Figure 3), in vitro studies show that 5-HT acts ing proteins (Klose & Bird, 2006). These proteins, in on 5-HT7 receptors to initiate a series of intracellular turn, attract an entire cluster of proteins, known as signals that culminate with an increase in the the complexes that are the active media- expression of NGFI-A as well as in the CREB-bind- tors of the gene silencing. The HDACs are a critical ing protein. Comparable effects occur in vivo. component of the repressor complex. HDACs pre- Manipulations that increase pup LG by lactating rats vent histone acetylation and favor a closed chroma- result in an increased level of cAMP as well as tin state that constrains transcription factor binding NGFI-A (Meaney et al., 2000). Pups reared by and gene expression (Figure 2 and see earlier). high-LG mothers show increased NGFI-A expres- Compounds that inhibit HDACs can thus increase sion in hippocampal neurons as well as an increased transcription from methylated DNA. binding of NGFI-A to the exon 17 promoter sequence (Hellstrom , Zhang, Diorio, & Meany, 2009; Weaver et al., 2007). Moreover, the binding of NGFI-A to the Epigenetics and the Social Environment exon 17 promoter sequence is actively regulated by The following section describes studies of the mother–pup interactions, such that there is increased molecular basis for the effects of maternal care on NGFI-A bound to the exon 17 promoter immediately the development of individual differences in gene following a nursing bout but not at a period that expression and stress responses. The mechanism follows 25 min without mother–pup contact for this interaction is epigenetic, involving altera- (Hellstrom et al., 2009). tions in DNA methylation at specific sites in the Nerve growth factor-inducible factor A and the genome. In summary, variations in mother–infant CREB-binding protein form a complex that binds interactions in the rat alter the extra- and intracellu- directly to the exon 17 promoter sequence and lar environment of neurons in selected brain actively redesigns the methylation pattern at this regions. Such alterations directly modify the epige- region of the genome (Weaver et al., 2004; Weaver netic marks on regions of the DNA that regulate et al., 2007). Thus, as adults the offspring reared by the transcription of the glucocorticoid receptor, high-LG mothers show very modest levels of meth- which in turns regulates the HPA response to ylation at the 5¢-CpG of the NGFI-A consensus stress. These epigenetic marks are stable, enduring sequence (Figure 5). This effect on methylation is well beyond the period of maternal care, and thus very precise. Lying only a few removed provide a molecular basis for a stable maternal from this site is 3¢-CpG site (Figure 5), the methyla- effect on the phenotype of the offspring. Thus, the tion status of which is unaffected by maternal care. behavior of the mother directly alters cellular sig- A rather novel aspect to the effect of maternal nals that then actively sculpt the epigenetic land- care on DNA methylation was apparent in the scape of the offspring, influencing the activity of results of a simple developmental study examining specific regions of the genome and the phenotype the methylation status of the 5¢- and 3¢-CpG sites of the offspring. from late in fetal life to adulthood (Weaver et al., 2004). Neither the 5¢- nor the 3¢-CpG site is methy- lated in hippocampal neurons from fetal rats, Epigenetic Effects of Variations in Maternal Care whereas both sites are heavily methylated on the The critical feature of the maternal effects day following birth, with no difference as a func- described earlier is that of persistence. The differ- tion of maternal care. These findings reflect what is Epigenetics 59

neurons with a virus containing a construct that was engineered to express high levels of NGFI-A produces demethylation of the 5¢-CpG of the exon 17 promoter sequence and increased glucocorticoid receptor expression. But there is a complication. If DNA methylation blocks transcription factor binding and the 5¢-CpG site of the exon 17 promoter is heavily methylated in neonates, then how might maternally activated NGFI-A bind to and remodel the exon 17 region? And why is the effect apparent at the 5¢- but not the 3¢-CpG? The answer to these questions appears Figure 5. A working hypothesis for the experience (maternal to involve other transcriptional signals that are care) driven remodeling of the epigenetic state of the NGFI-A affected by maternal care. Levels of the transcrip- consensus binding sequence over the 1st week of postnatal life tion factor specific protein-1 (SP-1) and the in the offspring of high-LG mothers. CREB-binding protein are also increased in the Note. The binding of a NGFI-A ⁄ CBP complex actively targets the hippocampus of pups reared by high-LG mothers association of a putative (MBD2) resulting in the removal of the methyl group from the 5¢-CpG site of the NGFI-A (Hellstrom et al., 2009; Weaver et al., 2007). The binding site (Meaney & Szyf, 2005). CBP = CREB-binding exon 17 promoter contains a DNA sequence that protein; LG = licking ⁄ grooming; NGFI-A = nerve growth factor- binds SP-1 and this region overlaps with that for inducible factor A. NGFI-A. SP-1 can actively target both methylation and demethylation of CpG sites (Brandeis et al., referred to as de novo methylation, whereby a methyl 1994). The 5¢-CpG site is the region of overlap in group is applied to previously unmethylated sites. the binding sites. The CREB-binding protein, on the However, between the day following birth and the other hand, acts as a histone acetyltransferase, an end of the 1st week of life, the 5¢-CpG is ‘‘demethy- enzyme capable of acetylating histone tails, includ- lated’’ in pups reared by high-, but not low-LG ing the exon 17 region, opening chromatin and per- mothers. The difference then persists into adult- mitting the binding of transcription factors such as hood. Importantly, the period over which the NGFI-A and SP-1. Increasing histone acetylation demethylation occurs is precisely that during which can lead to transcription factor binding at previ- high- and low-LG mothers differ in the frequency ously methylated sites, and the subsequent deme- of pup LG; the difference in pup LG between high- thylation of these regions (Szyf, Weaver, and low-LG mothers is not apparent in the 2nd Champagne, Diorio, & Meaney, 2005). week of postnatal life (Caldji et al., 1998; Cham- The NGFI-A transcription factor binds to multi- pagne, Francis, et al., 2003). ple sites across the genome. If NGFI-A-related com- The demethylation of the 5¢-CpG site occurs as a plexes target demethylation, then one might function of the same 5-HT-activated signals that assume that other NGFI-A-sensitive regions should regulate glucocorticoid receptor gene expression in show a comparable to that observed cultured hippocampal neurons (Weaver et al., with the glucocorticoid receptor. Zhang, Hellstrom, 2007). Thus, when hippocampal neurons of embry- Wei, and Meaney (2009) showed that the hippo- onic origin are placed in culture and treated with campal expression of the GAD1 gene that encodes 5-HT, which mimics the extracellular signal associ- for glutamic acid decarboxylase, a rate limiting ated with maternal LG, the 5¢-CpG site is demethy- enzyme in the production of GABA, is increased in lated; there is no effect at the 3¢-CpG site. The the adult offspring of high-LG mothers. This effect binding of NGFI-A to the exon 17 site is critical. is associated with increased cytosine methylation of Hippocampal neurons that are rendered incapable an NGFI-A response element. Moreover, as with of increasing NGFI-A expression through antisense the effect on the glucocorticoid receptor, an in vitro or siRNA treatment show neither the demethyla- increase in NGFI-A expression mimics the effects of tion of the 5¢-CpG site nor the increase in glucocor- increased pup LG. ticoid receptor expression (Weaver et al., 2007). In summary, the maternally induced changes in Likewise, a mutation of the NGFI-A site that specific intracellular signals in hippocampal neurons completely abolishes the binding of NGFI-A to the can physically remodel the genome. The increased exon 17 promoter also prevents the demethylation binding of NGFI-A that derives from pup LG of the 5¢-CpG. Finally, the of hippocampal appears critical for the demethylation of the exon 17 60 Meaney promoter. We suggest that this process involves HDACs. The HDACs deactylate histone tails, thus accompanying increases in SP-1 and the CREB- favoring a closed chromatin configuration. Indeed, binding protein, and that the combination of these the exon 17 promoter is more prominently acety- factors results in the active demethylation of the lated in hippocampi from adult offspring of high- exon 17 promoter. It should be noted that there are compared with low-LG mothers (Weaver et al., important features of this model that remain to be 2004, 2005). This finding is consistent with the clearly defined, including the identification of the increased transcription of the glucocorticoid recep- enzyme that is directly responsible for the deme- tor gene in animals reared by high- versus low-LG thylation. Nevertheless, the events described to mothers. A subsequent study (Weaver et al., 2004) date represent a model by which the biological examined the effects of directly blocking the actions pathways activated by a social event may become of the HDACs in the adult offspring of high- and imprinted onto the genome. This imprint is then low-LG mothers. An HDAC inhibitor was infused physically apparent in the adult genome, resulting directly into the hippocampus daily for 4 consecu- in stable alterations (or programming) of gene tive days. The treatment with the HDAC inhibitor expression. produces a series of predictable results that reflect a cause–effect relation between DNA methylation and gene expression. First, HDAC blockade elimi- nates the differences in the acetylation of the his- The Functional Importance of the Social Imprint tone tails of the exon 17 promoter in hippocampal The presence of a methyl group on the 5¢-CpG of the samples from high- and low-LG mothers. Second, NGFI-A binding site is functionally related to gluco- the increased histone acetylation of the exon 17 pro- corticoid receptor gene expression in adult animals. moter in the offspring of low-LG mothers is associ- In vitro studies reveal that the methylation of the ated with an increase in the binding of NGFI-A to 5¢-CpG site reduces the ability of NGFI-A to bind to the exon 17 promoter in the offspring of low-LG the exon 17 promoter and activate glucocorticoid mothers, eliminating the maternal effect on NGFI-A receptor transcription (Weaver et al., 2007). These binding to the exon 17 promoter. Comparable levels findings are consistent with the model described of NGFI-A binding to the exon 17 promoter then earlier, whereby DNA methylation impedes tran- eliminate the maternal effect on hippocampal glu- scription factor binding and thus the activation of cocorticoid receptor expression, such that glucocor- gene expression. The question concerns the in vivo ticoid receptor levels in the adult offspring of low- situation. In contrast to the situation with neonates, LG mothers treated with the HDAC inhibitor are there is no difference in NGFI-A expression as a comparable to those in animals reared by high-LG function of maternal care among adult animals. mothers. And most importantly, the infusion of the However, the altered methylation of the exon 17 HDAC inhibitor reversed the differences in the promoter would suggest differences in the access of HPA response to stress. NGFI-A to its binding site on the promoter. Chroma- inhibition increases NGFI-A tin-immunoprecipitation assays, which permit binding to the exon 17 promoter in the offspring of measurement of the direct interaction between a low-LG mothers. The studies with neonates reveal specific protein and a defined region of the DNA, that increased NGFI-A binding results in the deme- reveal increased NGFI-A association with the exon thylation of the 5¢-CpG. In vitro, the introduction of 17 promoter in hippocampi from adult offspring of a viral tool that leads to the increased expression of high- compared to low-LG mothers (Weaver et al., NGFI-A is sufficient to demethylate the exon 17 2004; Weaver et al., 2005). These findings show that promoter. We (Weaver et al., 2007) argue that the in the living animal, under normal conditions, there binding of NGFI-A is critical for the demethylation is more NGFI-A associated with the exon 17 of the 5¢-CpG site. The same effect is apparent promoter in hippocampal neurons of adult animals in vivo and even with the adult animals used in the reared by high- compared with low-LG mothers. studies described earlier. HDAC infusion into There is also evidence that directly links the the hippocampus increases NGFI-A binding to the maternal effect on the epigenetic state of the exon exon 17 promoter in the adult offspring of low-LG 17 promoter to the changes in glucocorticoid recep- mothers and decreases the level of methylation of tor expression and thus HPA responses to stress. the 5¢-CpG site on the exon 17 promoter. A subse- Recall that the methylation of specific CpG sites can quent study (Weaver et al., 2005) showed that the diminish transcription factor binding through the reverse pattern of results could be obtained in recruitment of repressor complexes that include response to a dietary manipulation (methionine): Epigenetics 61

Greater methylation of the 5¢-CpG in the offspring 1F promoter of the glucocorticoid receptor, which of high-LG mothers decreased NGFI-A binding and corresponds to the exon 17 promoter in the rat GR expression and increased HPA responses to (Turner & Muller, 2005). The results showed stress (Weaver et al., 2005). increased DNA methylation of the exon 1F promoter While these studies employ rather crude phar- in hippocampal samples from victims com- macological manipulations, the results are critical pared with controls, but only if suicide was accom- as they suggest that fully mature neurons in an panied with a developmental history of child adult animal express the necessary enzymatic maltreatment. Child maltreatment, independent of machinery to demethylate or remethylate DNA. psychiatric state, predicted the DNA methylation Thus, it is possible that environmentally driven status of the exon 1F promoter. As in the previous changes in neuronal transcriptional signals could rodent studies, the methylation state of the exon 1F potentially remodel the methylation state of specific promoter also determined the ability of NGFI-A to regions of the DNA (Meaney & Szyf, 2005). The bind to the promoter and activate gene transcrip- cytosine methylation state of a promoter for the tion. While such studies are obviously correlational, brain-derived neurotrophic factor (bdnf) gene is also and limited by postmortem approaches, the results influenced by maternal care in early life (Roth, are nevertheless consistent with the hypothesis that Lubin, Funk, & Sweatt, 2009). Indeed, Lubin et al. variations in parental care can modify the epigenetic (2008) provided evidence for an alteration of the state of selected sites of the . More- methylation state of the same bdnf promoter follow- over, the findings are also consistent with studies ing contextual fear conditioning in adult rats (also that link childhood abuse to individual differences see Bredy et al., 2007). These effects are consistent in stress responses (Heim et al., 2000). Childhood with previous reports of activity-dependent altera- abuse was associated with an increase in pituitary tions in the methylation of the same bdnf promoter ACTH responses to stress among individuals with (Martinowich et al., 2003) and suggest that epige- or without concurrent major depression. Heim netic states might be altered by a wide range of et al.’s (2000) findings are particularly relevant since biologically relevant events that result in synaptic pituitary ACTH directly reflects central activation of remodeling (Meaney & Szyf, 2005; Renthal & the HPA stress response and hippocampal GR acti- Nestler, 2008; Sweatt, 2009). Such epigenetic modifi- vation dampens HPA activity. These findings are cations might therefore underlie a wide range of consistent the rodent studies cited earlier investigat- stable changes in neural function following expo- ing epigenetic regulation of the glucocorticoid sure to highly salient events (e.g., chronic stress, receptor gene and with the hypothesis that early life drugs of abuse, reproductive phases such as parent- events can alter the epigenetic state of relevant ing, etc.), and are thus logical mechanisms for envi- genomic regions, the expression of which may con- ronmentally induced alterations in mental health tribute to individual differences in the risk for psy- (Akbarian & Huang, 2009; Jiang et al., 2008; Tsank- chopathology (Holsboer, 2000; Neigh & Nemeroff, ova, Renthal, Kumar, & Nestler, 2007). While such 2006; Schatzberg, Rothschild, Langlais, Bird, & Cole, effects have yet to be reported for DNA methyla- 1985). tion, modifications of histone proteins are associ- ated with exposure to drugs of abuse and stressors in rodent models (Renthal & Nestler, 2008; Renthal Transgenerational Effects et al., 2007). A set of recent studies (McGowan et al., 2009) Individual differences in stress responses in the suggests that comparable epigenetic modifications adult rat are associated with naturally occurring might occur in humans in response to variations in variations in maternal care during infancy. Manipu- parent–offspring interactions. DNA was extracted lations that alter mother–pup interactions in the rat from hippocampal samples obtained from victims alter patterns of gene expression and stress of suicide or from individuals that had died sud- response in the offspring (e.g., Francis et al., 1999; denly from other causes (auto accidents, heart Meaney, 2001; Roth et al., 2009; Toki et al., 2007). attacks, etc.). The samples were obtained from the Such effects are certainly familiar to child psycholo- Que´bec Suicide Brain Bank, which conducts forensic gists working with parenting training programs. phenotyping that includes a validated assessment of Moreover, these maternal effects might also serve psychiatric status and developmental history (e.g., as a possible nongenomic mechanism by which McGirr, Renaud, Seguin, Alda, & Turecki, 2008). The selected traits could be transmitted from one gener- studies examined the methylation status of the exon ation to another through variations in mother–pup 62 Meaney interactions. Interestingly, low-LG mothers are in the medial preoptic area (MPOA) of the hypo- more fearful than are high-LG dams (Francis, thalamus, a region that is critical for maternal Champagne, & Meaney, 2000). Individual differ- behavior in the rat (Fleming et al., 1999; Numan & ences in stress reactivity are apparently transmitted Insel, 2003). There is increased receptor a across generations: Fearful mothers beget more expression in female offspring of high-LG mothers stress reactive offspring. The obvious question is (Champagne, Weaver, Diorio, Sharma, & Meaney, whether the transmission of these traits occurs only 2003). Estrogen acts during late gestation to as a function of genomic-based inheritance. If this increase oxytocin receptor levels in the MPOA is the case, then the differences in maternal behav- (Fahrbach & Pfaff, 1986; Pedersen, 1997), and this ior may be simply be an epiphenomenon and not effect is greater in the female offspring of high- causally related to the development of individual compared to low-LG mothers (Champagne et al., differences in stress responses. The issue is not one 2001), reflecting the increased sensitivity to estro- of inheritance but rather the mode of inheritance. gen. Oxytocin appears to act at oxytocin receptor However, the results of the previously cited cross- levels in the MPOA to facilitate the release of dopa- fostering studies indicate that individual differ- mine from neurons in the ventral tegmental ences at the level of gene expression or complex nucleus, and the increased dopamine release then phenotype can be directly altered during the post- activates pup LG in lactating female rats (Cham- natal period by maternal behavior. This interpreta- pagne, Stevenson, Gratton, & Meaney, 2004; Shah- tion is buttressed from studies showing that tactile rohk et al., in press). This effect is abolished with stimulation of pups by the experimenter with a the infusion of an oxytocin receptor antagonist into brush, so-called stroking, increases hippocampal the ventral tegmental area (Shahrohk et al., in NGFI-A (Hellstrom et al., 2009) increases hippo- press). Likewise, the same oxytocin receptor antag- campal GR expression (Jutapakdeegul et al., 2003), onist completely eliminates the differences in and dampens HPA responses to stress in adulthood maternal behavior between high- and low-LG (Burton et al., 2007; Gonzalez et al., 2001). mothers (Champagne et al., 2001). Drugs that block The cross-fostering studies also reveal that indi- the reuptake of synaptic dopamine, thus increasing vidual differences in maternal behavior are trans- the overall dopaminergic signal, increase pup LG mitted from mother to the female offspring. Hence, in low-LG mothers and eliminate the difference in as adults, the female offspring of more fearful, maternal behavior between high- and low-LG low-LG mothers are also more fearful low-LG mothers. In summary, the increased estrogen recep- mothers (Francis et al., 1999; also see Fleming, tor a expression in the MPOA of high-LG mothers O’Day, & Kraemer, 1999). Regardless of their bio- leads to greater sensitivity to estrogen, an increased logical origins, females that are reared by high-LG level of oxytocin levels in MPOA neurons that pro- mothers are less fearful and show an increased fre- ject directly to the dopamine neurons in the ventral quency of pup LG. And the mechanism for the in- tegemental regions, and greater activation of dopa- tergenerational transmission of such individual mine release during nursing bouts. And the differ- differences appears to be the difference in the fre- ences in estrogen receptor a expression, like those quency of pup LG during early postnatal life. Flem- in pup LG, are reversed with cross-fostering ing and colleagues found that as adults, females (Champagne, Francis, et al., 2003), suggesting that rats deprived of maternal showed reduced frequen- maternal care regulates the activity of the estrogen cies of multiple forms of maternal behavior, and receptor in the MPOA, which then forms the basis such effects were reversed if animals were provided for subsequent ‘‘inherited’’ differences in maternal with tactile stimulation by stroking the pups with a behavior. brush over the first weeks of life (Gonzalez et al., This finding reveals another example of a mater- 2001; Lovic, Gonzalez, & Fleming, 2001; Melo et al., nal effect on gene expression, and there is evidence 2006). Moreover, the stroking in infancy also for epigenetic mediation. The activation of the increased the expression of maternal behaviors in estrogen receptor a-gene in the rat brain occurs response to ovarian hormones in adulthood (Nova- through the estrogen receptor 1B promoter. This kov & Fleming, 2005), an effect that is comparable promoter contains multiple cytosine sites that are to that associated with increased pup LG (Cham- potential targets for DNA methylation. Champagne pagne, Diorio, Sharma, & Meaney, 2001; see also et al. (2006) found increased cytosine methylation next). across the exon 1B promoter in the offspring of The individual differences in pup LG involve low-LG mothers. Activation of the exon 1B differences in estrogen receptor a-gene expression promoter occurs in response to the binding of the Epigenetics 63 transcription factor Stat5, and the adult offspring of low-LG mothers show decreased Stat5 association with the exon 1B promoter in the MPOA. These findings suggest that differences in DNA methyla- tion may mediate the effect of maternal care on the expression of estrogen receptor a in the MPOA and thus serve as the molecular basis for the nongenom- ic transmission of individual differences in maternal behavior from the mother to her female offspring.

Figure 6. Summary of the literature from Adaptive Value of Epigenetic Maternal Effects on ‘‘maternal effects.’’ Note. The consistent theme reflected in these studies is that Variations in parental signals alter gene expression various environmental signals can alter multiple phenotypic and thus the development of individual differences outcomes through effects on parent–offspring interactions, in complex . But why bother? Why broadly referred to as parental investment. The relevant form of the variation in parent–offspring interaction (investment) will would nature configure such a process? Why trans- vary depending upon the species. The principal idea is that of mit individual differences in stress reactivity or parental mediation and of coordinated effects on multiple maternal behavior across generations through a pro- phenotypic outcomes. cess that is driven by parental care and mediated by the complex cellular machinery necessary to rear- most relevant for mammalian species. Nevertheless, range epigenetic markings on the DNA? Why not the stability of the relation between the quality of simply leave such issues of inheritance in the hands the prevailing environment, parental investment, of classic genetic transmission? The answer may lie and phenotypic development in the offspring led in the simple fact that unlike nucleotide sequence, Hinde (1986) to suggest that may have epigenetic marks are dynamic and indeed revers- actually shaped the offspring to ‘‘use’’ parental sig- ible. The more flexible epigenetic mechanism would nals, such as variations in parental care behaviors, provide the basis for an adaptive parental effect: In to forecast the quality of the environment in which the rat at least, parents can actively remodel epige- they must function (also see Bateson, 1994; Bateson netic marks and thus affect patterns of gene expres- et al., 2004; Rossiter, 1998). Evolution should come sion in the offspring. Such effects do not occur at to favor offspring that are able to accurately ‘‘read’’ the level of nucleotide sequence. variations in parental behavior as forecasts of envi- Studies in the fields of evolutionary biology and ronmental conditions and thus as useful signposts report maternal effects on phenotype in the to guide developmental outcomes (Hinde, 1986). By offspring across a wide range of species (see definition, such responses should occur in reaction Figure 6; Badyaev, 2008; Cameron et al., 2005; to variation within the normal range. Why evolve Meaney, 2007; Mousseau & Fox, 1998; Rossiter, responses to forms of parental care that are unlikely 1998). Despite the fact that these studies have been to occur? Parents (or parent in some cases) are a performed largely with simpler , such as logical source of such information since they are plants, insects, and reptiles, the emerging theme is the one ‘‘environmentally informed’’ and constant much the same as that described earlier in studies experience for the offspring. Moreover, since par- with human families: Environmental adversity ents are invested in the adaptive success of their decreases parental investment in the offspring and biological offspring, one would expect that the thus alters phenotypic development. And there is fidelity of signals emanating from a parent would evidence that these phenotypic effects are adaptive be greater than that of other adult conspecifics. within adverse settings (Mousseau & Fox, 1998). Thus, to the extent that the parent and offspring Obviously, the form of ‘‘parental investment’’ share a common interest in the adaptive value of varies across species. In plants, the variation may such , selection may also act involve seed quality. Among certain insects on the signaling capacity of the parent. Indeed, ‘‘investment’’ would include the nutrient value of phenotypic plasticity in response to parental signals the ‘‘propulgate,’’ the food source left by the may also be thought of as a parental strategy, mother for the hatching offspring. While some although it is clear that the benefits work in both insects do exercise postnatal behavioral care of the directions as would be expected among organisms offspring, variations in this form of investment are that share genes. In either case, the sensitivity of 64 Meaney the offspring to parental signals during critical Stress and Parenting phases of development may be thought of as a strategy that favors a highly predictable relation Perhaps, the most compelling evidence for a between environmental conditions, parental input, direct effect of environmental adversity on parent– and phenotypic variation in defensive responses. offspring interactions emerges from the studies of For the offspring, parents should matter. Rosenblum, Coplan, and colleagues with nonhu- Hinde’s (1986) formulation implies an adaptive man primates (Coplan, Andrews, Rosenblum, & function for parental effects in preparing the off- Nemeroff, 1996). Bonnet macaque mother–infant spring for the environmental demands that are dyads were maintained under typical lab condi- likely to prevail during development and into tions, with free, unhindered access to food (a low- adulthood (also see Bateson et al., 2004; Gluckman foraging-demand condition) or one in which the & Hanson, 2004; Meaney, 2001). Indeed, it is proba- amount of available food varied and required long bly true that the environmental conditions of the periods of searching (high-foraging-demand condi- adult are a reasonable predictor of those facing the tion). The high-foraging-demand condition severely offspring (with modern, urbanized human popula- disrupted mother–infant interactions, producing tions as a rare exception; Gluckman & Hanson, significant conflict. Infants of mothers housed 2004; West & King, 1987). However, environmental under these conditions were more timid and fear- conditions can vary with the migration of popula- ful, with some revealing signs of depression that tions. Moreover, even within a similar environ- are commonly observed in maternally separated ment, the ability of individuals to access resources macaque infants. Such reactions were even appar- can vary from one individual to another as a func- ent among infants that were in contact with their tion of, for example, social hierarchies. mothers. As adolescents, the infants reared in the Success requires the ability to adapt, to fine-tune high-foraging-demand conditions were more fear- phenotypic development in relation to the prevail- ful, submissive, and showed less social play behav- ing environmental demands. Learning, of course, ior. As expected, these conditions also affected would provide a mechanism for such phenotypic the development of neural systems that mediate plasticity. But learning requires trial and, unfortu- behavioral and endocrine response to stress. Adult nately, error. Nature does not always permit such monkeys reared under variable foraging demand indulgences. Epigenetic variation may provide a (VFD) conditions showed increased central levels of mechanism for the dynamic regulation of the geno- CRF and increased noradrenergic responses to mic machinery in response to variation in prevail- stress. It would be fascinating to see if such traits ing environmental conditions during development would then be transmitted to the next generation. (Jablonka & Lamb, 2005). Such epigenetic adapta- In some rather remarkable cross-fostering studies, tions could then occur directly as a result of the rel- Maestripieri (2005) has shown that individual dif- evant environment condition (e.g., nutrient ferences in maternal behavior are transmitted availability) or indirectly through parental media- across generations in the rhesus monkey. tion. Such effects could, for example, explain the The critical issue here is that of a direct effect of enduring effects of childhood socioeconomic status environmental adversity on maternal behavior. on health outcomes in humans. Stress during decreases maternal respon- We suggest that parental signals over the perina- sivity in lactating rats (Fride, Dan, Gavish, & Wein- tal period serve as an important catalyst for epige- stock, 1985; Kinsley, Mann, & Bridges, 1998; Moore netic remodeling of the genome. If such effects & Power, 1986). Gestational stress also eliminates support an adaptive function, then there should be the differences in pup LG between high- and low- a high degree of fidelity between the quality of the LG mothers (Champagne & Meaney, 2006). Gesta- environment and parental care. This argument rests tional stress decreases the frequency of maternal LG upon at least three assumptions. First, there is a in the high-, but not in low-LG mothers (also see consistent relation between the quality of the envi- Smith, Seckl, Evans, Costall, & Smythe, 2004), and ronment and that of parental care. Second, varia- the effect is apparent even with subsequent litters in tions in specific developmental outcomes should be the absence of further exposure to stress. As adaptive. This includes an adaptive value to traits expected, the effects on maternal behavior are that appear to increase vulnerability to illness, such apparent in the development of the offspring. As as increased stress reactivity. And third, the ecol- adults, the offspring of high-LG ⁄ gestationally ogy of the neonate reliably predicts the conditions stressed mothers were comparable to those of low- of adulthood. LG dams on measures of maternal behavior, as well Epigenetics 65 as in fear behavior and hippocampal glucocorticoid could represent an adaptive approach to develop- receptor gene expression. These effects can be dis- ment. Since the offspring usually inhabit a niche tinguished from those associated with prenatal that is similar to their parents, the transmission of stress in animals that were in utero during the individual differences in traits from parent to off- imposition of the stressor by simply examining the spring could serve to be adaptive with respect to offspring of subsequent litters in which the behavior survival. The conceptually challenging feature of of the mother remains affected, but in the absence of this argument is that it requires that we identify an the stressor (Champagne & Meaney, 2006). adaptive virtue to a phenotypic profile, such as As mentioned earlier, there is considerable, albeit increased HPA stress responses, that predict greater correlational, evidence for a relation between envi- vulnerability to chronic illness. ronmental adversity and parental care in humans. Adverse environmental conditions such as those Environmental adversity influences emotional well- associated with poverty over the adult life of the being in parents and these effects are reflected in parent have historically predicted more of the same alterations in parental care, commonly reflecting a for the offspring. The critical challenge over the decreased level of investment (Fleming, 1988; Repet- course of development is to mold specific features ti et al., 2002). Increased maternal stress is associated of phenotype in a manner that is most appropriate with less sensitive child care (Dix, 1991; Goldstein, to the level of environmental demand. This capacity Diener, & Mangelsdorf, 1996). The children of for phenotypic plasticity (Agrawal, 2001; Mousseau highly stressed primary caregivers tend to develop & Fox, 1998) is a product of evolutionary forces. As insecure parental attachment (Goldstein et al., 1996; such, it is critical to note that the ultimate measure Vaughn, Egeland, Sroufe, & Waters, 1979). Parents of success for any phenotypic trait is the degree to in conditions of poverty experience more negative which it enhances the probability for reproductive emotions, irritability, depressed, and anxious success. The health of the individual is a relevant moods, which lead to more punitive parenting (Bel- consideration only to the degree that it influences the sky, 1993; Conger, McCarty, Yang, Lahey, & Kropp, ability to successfully reproduce. Health will influence 1984; Grolnick, Gurland, DeCourcey, & Jacob, 2002). survival and the ability to attract mating opportuni- The resulting patterns of parental behavior can ties, and may therefore be of relevance, at least dur- affect the development of HPA responses to stress ing the period of active . However, the in the offspring (Heim et al., 2000; Pruessner, Cham- incidence of chronic illnesses that commonly occur pagne, Meaney, & Dagher, 2004), an effect that may in later life, once the prospects for reproductive associate with altered glucocorticoid receptor activity have declined, is not a major consideration expression (McGowan et al., 2009). Hane and Fox in evaluating the adaptive value of any specific trait (2006) found a direct relation between the quality of in the evolutionary sense. maternal care and behavioral inhibition in children. There is a very real divide between the way in There is rather compelling evidence for parental which the success of development is evaluated by mediation in the effects of impoverished environ- those working in the health sciences as compared ments on phenotypic development. Thus, the effects to those in biology. Optimal development for of poverty on emotional and cognitive development those in the health sciences is judged by the qual- are, in part, mediated by variations in parent–off- ity of life and the absence of disease. For biolo- spring interactions. If parental care factors are statis- gists, success is measured in the terms of the tically controlled, there no longer remains any currency of —reproductive fit- discernible effect of poverty of child development ness. For , there are no ideal pheno- (Conger, Ge, Elder, Lorenz, & Simons, 1994; McL- types. Rather, the adaptive merits of phenotype loyd, 1998). Moreover, parents and parenting style are apparent only in relation to success within a are highly effective targets for intervention studies particular set of environmental conditions: One aimed at development outcomes in children living single phenotype does not fit all. For example, an in adverse environmental conditions (Belsky, 1997; increase in the responsivity of an individual to Fisher et al., 2000; Offord et al., 1992; Olds et al., threat (i.e., greater stress reactivity) is realistically 1998; Van den Boom, 1994). considered as a risk factor for multiple forms of chronic illness. Thus, the catabolic effects associ- ated with adrenal glucocorticoids or sympathetic Adaptive Value of Increased Stress Responses catecholamines tend to be vilified, especially in Intergenerational transmission of individual dif- psychology and psychiatry. Indeed, chronic eleva- ferences in stress reactivity via parental behavior tions in the levels of these ‘‘stress mediators’’ can 66 Meaney directly promote illness (Chrousos & Gold, 1992; Maternal Care and Neurocognitive Development Dallman et al., 2005; McEwen, 2007; Sapolsky et al., 2000; Walker, 2007). However, the bitter Pup LG in the rat dynamically alters the activity truth of adaptation and survival in the face of endocrine systems that favor of growth adversity is that such hormonal stress responses (Levine, 1994; Schanberg et al., 1984). Comparable are essential for continued life. The survival inter- effects are apparent in the hippocampus, a brain ests of an individual during periods of increased region intimately associated with learning and environmental demand are well served by behav- . The neonatal offspring of high-LG moth- ioral (e.g., vigilance, fearfulness) and endocrine ers show increased hippocampal expression of (HPA and metabolic ⁄ cardiovascular) responses to genes that encode for neurotrophic factors, which stress. These responses promote detection of supports the sprouting and survival of synapses, potential threat, fear conditioning to stimuli asso- this effect includes the expression glutamate recep- ciated with threat and avoidance learning. More- tor subunits. The activation of glutamate receptors over, the hormonal effectors of sympathoadrenal in the hippocampus stimulates synaptogenesis and HPA stress responses mobilize energy (Kirkwood, Dudek, Gold, Aizenman, & Bear, 1993; reserves through effects of lipolysis, glycolysis, Schatz, 1990). Thus, as adults, the offspring of high- and gluconeogenesis. These effects are the hall- LG mothers show increased hippocampal synaptic mark of the shift to catabolism that occurs during density (Bredy, Humpartzoomian, Cain, & Meaney, periods of stress and are essential for animals 2003; Champagne et al., 2008; Liu et al., 2000) asso- exposed to chronic stress, particularly if the stres- ciated with more extensive dendritic arborization of sor is coupled to conditions of famine. Indeed, both pyramidal and granule cell populations (Bagot the ability to survive sustained periods of nutri- et al., 2009; Champagne et al., 2008). The effect of ent deprivation depends upon the capacity to maternal care on synaptic development suggests an increase circulating levels of glucocorticoids and increased capacity for synaptic plasticity in the off- catecholamines. Impoverished environments are spring of high-LG mothers. Long-term potentiation also commonly associated with multiple sources (LTP) is an electrophysiological model of synaptic of infection. Under such conditions adrenal gluco- plasticity at specific neural sites that is thought to corticoids serve as a potent defense against septic mimic the structural remodeling of synaptic con- shock (Munck et al., 1984; Sapolsky et al., 2000). nections that underlie learning and memory. Hip- Among rats, animals with increased HPA pocampal LTP formation is indeed stronger in the responses to agents such as bacterial endotoxins adult offspring of High compared to Low LG moth- are at reduced risk for sepsis. Interestingly, adults ers (Bagot et al., 2009; Bredy et al., 2003; Cham- exposed to a bacterial endotoxin during the 1st pagne et al., 2008). And not surprisingly, the adult week of life exhibit increased HPA responses to offspring of high-LG mothers show improved per- stress as well increased resistance to sepsis formance hippocampal-dependent tests of learning upon subsequent exposure to bacterial infection and memory, such as spatial learning (Bredy, (Shanks, Larocque, & Meaney, 1995; Shanks Zhang, Grant, Diorio, & Meaney, 2004; Liu et al., et al., 2000). Conversely, postnatal conditions 2000) and object recognition (Bredy et al., 2004). that increase maternal LG and dampen HPA These findings suggest a rather predictable effect responses to stress increase vulnerability to endo- of maternal care on hippocampal development, toxin-induced sepsis. These findings underscore involving enhanced synaptic development in neo- the potentially adaptive value of increased HPA natal life and an increased capacity for synaptic and sympathetic responses to stress, especially for plasticity in adulthood, revealed in electrophysio- individuals living under conditions of impoverish- logical as well as behavioral studies. Such findings ment and infection. Thus, the decreased parental would seem to reflect an advantage for the off- investment associated with more stressful condi- spring of high-LG mothers. But these findings fail tions may in fact be of benefit to the offspring, if to consider the issue of context. Studies of in vitro indeed such conditions remain stable over time. LTP are performed with hippocampal slices Decreased parental investment might enhance the obtained from animals in the resting state. Studies stress responses of the offspring, which could of in vivo LTP or behavioral tests of learning and serve to protect against repeated periods of nutri- memory are performed in animals habituated to ent deprivation and infection. This theme is the testing conditions. What if animals were tested apparent in the effects of maternal care on neuro- under more demanding, stressful conditions? cognitive development in the rat. Champagne et al. (2008) and Bagot et al. (2009) Epigenetics 67 examined LTP in hippocampal slices obtained from The results of studies on the effects of maternal the adult offspring of high- or low-LG mothers care on hippocampal development in the rat reflect under basal conditions, or in the presence of stress- an important point. Although it is tempting to like levels of glucocorticoids. Glucocorticoids are an assume that increased pup LG enhances synaptic endocrine signature of the stress response and plasticity and cognitive performance, such effects known to diminish hippocampal LTP (Bodnoff are apparent only under conditions that are mini- et al., 1995; Diamond & Rose, 1994; Joels, Karst, mally stressful. Under stressful conditions, such as Krugers, & Lucassen, 2007; Kim & Yoon, 1998; those typified by contextual fear conditioning, it is Shors, Foy, Levine, & Thompson, 1990). This effect the offspring of low-LG mothers that show is readily apparent in sections obtained from improved synaptic plasticity and learning. We sug- high-LG mothers. LTP under basal conditions is gest that such findings represent a situation in markedly greater in hippocampal slices from high- which parental care shapes adaptive phenotypes compared to low-LG offspring, and the magnitude (Hinde, 1986; Meaney, 2001; Bateson et al., 2004; of LTP in slices from high-LG offspring is markedly Zhang et al., 2004; Gluckman & Hanson, 2007). In reduced by glucocorticoids. The exact opposite is the rat, environmental stressors decrease parental observed in hippocampal slices prepared from the investment and enhance behavioral and endocrine offspring of low-LG mothers: Glucocorticoids sig- responses to stress, as well as learning and mem- nificantly enhance LTP to the levels observed in ory under stressful conditions. If indeed the hippocampal slices from high-LG mothers under decreased parental investment accurately reflects basal conditions. This is true for LTP obtained from an increased level of environmental demand for multiple regions within the hippocampus (i.e., for the offspring, then such effects could be highly aficionados of hippocampal , this included adaptive. Regardless of the merits of this proposal, neurons within both the Ammon’s Horn and the these findings attest to the simple fact that the dentate gyrus). And an analogous effect is apparent adaptive value of any phenotypic profile depends at the level of learning and memory. When the upon the environmental context: There is no uni- adult offspring of high- and low-LG mothers are versally ‘‘ideal’’ phenotype. tested not under the benign conditions described There is also evidence for the potentially adap- earlier but in tests on contextual fear conditioning, tive effects of increased stress reactivity that the performance of the offspring of the low-LG more directly bears on the interests of child and mothers is significantly improved (i.e., increased adolescent psychology. The research of Farring- learning of the context–shock contingency) over ton, Gallagher, Morley, St Ledger, and West that of high-LG dams (Bagot et al., 2009; Cham- (1988) and Tremblay (e.g., Haapasap & Tremblay, pagne et al., 2008). Contextual fear conditioning is 1994) on young males growing up in impover- also a hippocampal-dependent test of learning and ished, high-crime urban settings illustrates the memory (Fanselow, 2000; Maren & Quirk, 2004), in potential advantages of increased emotional stress this case one that involves the association of a con- reactivity. Both studies show that shy and more text with an aversive event (i.e., a mild shock). timid males are most successful in avoiding the Thus, both hippocampal synaptic plasticity and pitfalls associated with ‘‘criminogenic’’ environ- hippocampal-dependent learning and memory are ments. Under such conditions, behavioral inhibi- enhanced in the offspring of low-LG mothers under tion emerges as a protective factor (Haapasap & stressful conditions. The effects of maternal care on Tremblay, 1994), despite the fact that this same hippocampal function are context dependent. profile is associated with an increased risk for mood disorders in later life (Pe´rez-Edgar & Fox, 2005). Moreover, under such adverse conditions a parental rearing style that favored the develop- Implications ment of a greater level of stress reactivity to All cellular processes derive from a constant dialo- threat could be viewed as adaptive. If indeed gue between the genome and environmental sig- there is no single ideal phenotype, then it should nals. Thus, genotype–phenotype relations are follow that there is no single ideal form of parenting. defined by the context within which the genome If this conclusion has worth, then it leads us to operates. Likewise, the consequences for question the wisdom of establishing parenting Gene · Environment interactions at the level of programs that foster parental skills based on function are defined by the broader context, includ- studies of families rearing children under more ing the demands of the prevailing environment. favorable conditions. 68 Meaney

Likewise, the presence of histone modifications Caveats (H3K4me) that associate with transcriptionally The magnitude and scope of the maternal effects on active genes appears to be determined not only by rodent development should not be surprising. DNA methylation but also by the underlying geno- Hinde (1986), Bateson (1994), and others have mic sequence. Such findings should dissuade us described the logic for the importance of maternal from assuming that genomic sequence is merely signals for the developing mammal. Indeed, even passive player in the definition of epigenetic states. the gametes include parental signals apart from The underlying genomic sequence influences the genomic DNA, such as RNAs and proteins, which nature of potential epigenetic states as well as their influence . Nevertheless, it importance for gene transcription. Finally, many is important to understand that the rodent studies transcriptionally inactive genes show relatively un- described earlier are designed to examine how vari- methylated promoters. While this finding does not ations in maternal can affect development; they do preclude the possibility that specific modifications not address the issue of how and to what degree in the methylation status at critical CpG sites may maternal care influences development under nor- define changes in gene transcription, such as occurs mal circumstances in the rat. Thus, the model for the IL2 gene (Murayama et al., 2006), it is described earlier examines maternal effects on neu- important to bear in that DNA methylation is ral development in animals that are then housed only one level of influence in the very complex bio- under highly standardized conditions following the chemical machinery that regulates gene transcrip- weaning period, eliminating the possibility that tion. These are still early days in the study of the maternal effects might be subsequently altered by molecular mechanisms for the environmental pro- variations in the postweaning environment. Indeed, gramming of gene expression. But the door has there is evidence that postweaning environments clearly been opened. can reverse the effects of preweaning variations in maternal care in the rat (Bredy et al., 2003; Cham- pagne et al., 2006). These studies are perhaps best General Conclusions considered as a model for the mechanisms by which social experience in early life can influence The effects of maternal care on gene expression and the structure and operation of the genome with neural function in the rat provide an understanding respect to specific phenotypic outcomes. Indeed, as of how environmental events, including variations suggested by Belsky (1997), Rutter (2007), and oth- in parent–offspring interactions at the level of ers, the effects of environmental influences, such as behavior, can become physically imprinted upon maternal care, will vary across individuals, and the the genome. Maternal care can directly alter intra- studies of Fox et al. (2005) in humans and Suomi cellular signals that, in turn, structurally alter the and colleagues (Bennett et al., 2002; Champoux DNA and its operation. These structural modifica- et al., 2002; Suomi, 2006) in the nonhuman primate, tions involve DNA methylation, a classic epigenetic suggest that variation in genomic sequence may be mark that regulates gene transcription. More recent a critical factor in determining sensitivity to paren- studies suggest that comparable epigenetic modifi- tal signals. cations associate with learning and memory (Lubin The study of epigenetics provides a molecular et al., 2008; Miller & Sweatt, 2007), chronic expo- mechanism for environmental effects of gene sure to drugs of abuse (Renthal & Nestler, 2008), expression. But the ultimate effect on gene tran- and psychiatric illness (Grayson et al., 2005; Ptak & scription is best thought of as an emergent property Petronis, 2008; Tsankova et al., 2007). The dynamic of the interaction between the epigenetic state and genome is probably a slightly foreign concept to the underlying genome. Thus, genomic sequence those who imagine the DNA as simply the reposi- determines the effects of epigenetic states on gene tory of the sequence information that forms what is transcription. While DNA methylation is linked to commonly referred to as the ‘‘genetic code.’’ But transcriptional silencing, such relations are not uni- the research of the postgenomic era, with its focus versal and the strength of the relation may be deter- on the operation rather than simply the composi- mined by the underlying genomic sequence (Weber tion of the genome, reveals that the DNA is an et al., 2007). For example, among promoters with a active target for remodeling by cellular signals that lower percentage of CpG dinucleotides, there are activated by environmental events. The reality appears to be a weaker relation between the overall of the functional genome does not admit to main level of methylation and that of gene transcription. effects of either gene or environment, but rather to Epigenetics 69 a constant interaction between the DNA and its very physical interaction of the genome with envi- environment. ronmentally regulated, intracellular signals that The study of epigenetic mechanisms reveals the directly alter chromatin structure. Thus, function at importance of DNA remodeling for gene transcrip- any level of biology emerges as a function of the tion, which is certainly the most fundamental opera- continuous dialogue between the genome and its tion of the genome. The transcription of the genome environment. Attempts to parse the influence of requires dynamic alterations to the chromatin struc- genomic and environmental influences on the tures within which the DNA operates. Such events expression of complex traits are inconsistent with involve transient modifications of the histone pro- even the most rudimentary understanding of gene teins that determine the accessibility of the DNA. function. Indeed, cellular signals that increase or decrease These issues are critical for the study of child gene transcription often function within complexes development. Research in this field guides both the that include enzymes that directly operate on chro- creation and the evaluation of intervention studies matin. Potentially more stable signals, such as DNA (e.g., Fisher et al., 2000; Olds et al., 1998). The science methylation, are also regulated by environmental and technology of the genomic era offer remarkable events and regulate chromatin activity over longer opportunities to enhance the sophistication of our periods. Such epigenetic marks are thus a candidate approach to these challenges. We are now positioned mechanism for the environmental programming of to study directly the origins of individual differences gene expression. This level of complexity is an in sensitivity ⁄ resistance to treatment within the con- essential feature of biology. The ability of an organ- text of intervention studies (Belsky, 1997; Rutter, ism to adapt to variations in environmental condi- 2007). Indeed, the many years of remarkable tions over the course of its lifespan demands research in child psychology and psychiatry forms plasticity in genotype–phenotype relations. the basis for hypothesis-driven studies of candidate The emergence of an integrative developmental Gene · Environment interactions of responses to perspective has been constrained by historical mis- treatment within the context of intervention studies. understandings of the processes by which variations Thus, the response of a child to psychosocial inter- in genomic sequence and nongenomic factors con- ventions, such as those targeting mother–infant tribute to the development of individual differences attachment, is influenced by the temperament or in any specific traits. While quantitative behavioral reactivity of the child (Belsky, 1997). More reactive genetics is a valid approach in establishing patterns children also seem to be more sensitive to the quality of familial transmission (i.e., heritability) that might of the rearing environment (Boyce & Ellis, 2005). The imply a causal role for genomic variation, such stud- temperament of the child is influenced by the pres- ies are commonly subject to misinterpretation of ence of specific genomic variations in genes that genomic influences. The heritability of individual encode for proteins involved in the function of the 5- differences could result as a function of nongenomic HT and dopamine systems (Auerbach et al., 2002; biological signals such as RNA and protein emanat- Ebstein, 2003; Fox et al., 2005; Lakatos et al., 2003). ing from the parent as well as from instances where Could such genomic variations mediate the epigenetic modifications enter the germline (Anway, observed effect of child temperament on the Cupp, Uzumcu, & Skinner, 2005; Jablonka & Lamb, response to specific interventions? What might be a 2005; Vardhman et al., 2003; Whitelaw & Whitelaw, more appropriate intervention for ‘‘treatment-resis- 2006). There are multiple processes by which indi- tant’’populations? A more integrated Gene · Envi- vidual differences in complex traits might be trans- ronment approach offers the opportunity to mitted from parent to offspring. Moreover, understand clearly the nature of individual differ- estimates of the contribution of either genetic or envi- ences in vulnerability or resistance for psychopathol- ronmental influences at the level of individual varia- ogy and to more effectively target interventions. tion are complicated by the reality of Gene · Environment interactions (Lewontin, 1974). Most troublesome is the obvious disjunction References between the definition of genomic influences that derives from statistical models and the actual bio- Agrawal, A. A. (2001). Phenotypic plasticity in the inter- logical reality of the genome. It is the operation of actions and evolution of species. Science, 294, 321–326. the genome that directly influences phenotype. The Akbarian, S., & Huang, H. S. (2009). Epigenetic regulation operation of the genome at any phase of the life in -focus on histone lysine methylation. cycle is an emergent property of the constant and , 65, 198–203. 70 Meaney

Anway, M. D., Cupp, A. S., Uzumcu, M., & Skinner, M. Bird, A. P. (1986). CpG-rich islands and the function of K. (2005). Epigenetic transgenerational actions of endo- DNA methylation. Nature, 321, 209–213. crine disruptors and mate fertility. Science, 308, 1466– Bird, A. P. (2002). DNA methylation patterns and epige- 1469. netic memory. Genes Development, 16, 6–21. Auerbach, J., Geller, V., Lezer, S., Shinwell, E., Belmaker, Bird, A. (2007). of epigenetics. Nature, 447, R. H., Levine, J., et al. (1999). Dopamine D4 receptor 396–398. (D4DR) and serotonin transporter promoter (5- Bird, A. P., & Wolffe, A. P. (1999). Methylation-induced HTTLPR) polymorphisms in the determination of tem- repression—Belts, braces, and chromatin. Cell, 99, 451– perament in 2-month-old infants. Molecular Psychiatry, 454. 4, 369–373. Blier, P., & de Montigny, C. (1999). Serotonin and drug- Auerbach, J., Geller, V., Lezer, S., Shinwell, E., Belmaker, induced therapeutic responses in major depression, R. H., Levine, J., et al. (2002). Dopamine D4 receptor obsessive-compulsive and panic disorders. Neuropsy- (D4DR) and serotonin transporter promoter (5- chopharmacology, 21, 91S–98S. HTTLPR) polymorphisms in the determination of tem- Bodnoff, S. R., Humphreys, A. G., Lehman, J. C., Dia- perament in 2-month-old infants. Molecular Psychiatry, mond, D. M., Rose, G. M., & Meaney, M. J. (1995). 4, 369–373. Enduring effects of chronic corticosterone treatment on Badyaev, A. V. (2008). Maternal effects as generators of spatial learning, synaptic plasticity, and hippocampal evolutionary change. Annals of the New York Academy of neuropathology in young and mid-aged rats. Journal of Sciences, 1133, 151–161. , 15, 61–69. Bagot, R. C., van Hasselt, F. N., Champagne, D. L., Mea- Boulinier, T., & Staszewski, V. (2008). Maternal transfer ney, M. J., Krugers, H. J., & Joe¨ls, M. (2009). Maternal of antibodies; raising immuno-ecology issues. Trends in care determines rapid effects of stress mediators on Ecology and Evolution, 23, 282–288. synaptic plasticity in adult rat hippocampal dentate Boyce, W. T, & Ellis, B. J. (2005). Biological sensitivity to gyrus. Neurobiology of Learning & Memory, 92, 292–300. context: I. An evolutionary-developmental theory of Bakermans-Kranenburg, M. J., Van IJzendoorn, M. H., the origins and functions of stress reactivity. Develop- Pijlman, F. T., Mesman, J., & Juffer, F. (2008). ment and Psychopathology, 17, 271–301. Experimental evidence for differential susceptibility: Brandeis, M., Frank, D., Keshet, I., Siegfried, Z., Mendel- Dopamine D4 receptor polymorphism (DRD4 VNTR) sohn, M., Nemes, A., et al. (1994). Sp1 elements protect moderates intervention effects on toddlers’ exter- a CpG island from de novo methylation. Nature, 371, nalizing behavior in a randomized controlled trial. 435–438. Developmental Psychology, 44, 293–300. Bredy, T. W., Humpartzoomian, R. A., Cain, D. P., & Bale, T. L., & Vale, W. W. (2004). CRF and CRF receptors: Meaney, M. J. (2003). The influence of maternal care Role in stress responsivity and other behaviors. Annual and environmental enrichment on hippocampal devel- Review of & , 44, 525–557. opment and function in the rat. Neuroscience, 118, 571– Bateson, P. (1994). The dynamics of parent–offspring rela- 576. tionships in mammals. Trends in Ecology and Evolution, Bredy, T. W., Wu, H., Crego, C., Zellhoefer, J., Sun, Y. E., 9, 399–403. & Barad, M. (2007). Histone modifications around indi- Bateson, P., Barker, D., Clutton-Brock, T., Deb, D., vidual BDNF gene promoters in are D’Udine, B., Foley, R. A., et al. (2004). Developmental associated with extinction of conditioned fear. Learning plasticity and human health. Nature, 430, 419–421. & Memory, 14, 268–276. Belsky, J. (1993). Etiology of child maltreatment: A devel- Bredy, T. W., Zhang, T.-Y., Grant, R. J., Diorio, J., & Mea- opmental-ecological analysis. Psychological Bulletin, 114, ney, M. J. (2004). Peripubertal environmental enrich- 413–434. ment reverses the effects of maternal care on Belsky, J. (1997). Theory testing, effect-size evaluation, hippocampal development and glutamate receptor sub- and differential susceptibility to rearing influence: The unit expression. European Journal of Neuroscience, 20, case of mothering and attachment. Child Development, 1355–1362. 64, 598–600. Bronnegard, M., & Okret, S. (1991). Regulation of the glu- Bennett, A. J., Lesch, K.-P., Heils, A., Long, J. C., Lorenz, cocorticoid receptor in fetal rat lung during develop- J. G., Shoaf, S. E., et al. (2002). Early experience and ment. Journal of Steroid Biochemistry & , serotonin transporter gene variation interact to influ- 39, 13–17. ence primate CNS function. Molecular Psychiatry, 7, Bruniquel, D., & Schwartz, R. H. (2003). Selective, stable 118–122. demethylation of the interleukin-2 gene enhances tran- Bestor, T. H. (1998). Gene silencing. Methylation meets scription by an active process. Nature , 4, acetylation. Nature, 393, 311–312. 235–240. Bettegowda, A., & Smith, G. W. (2007). Mechanisms of Burton, C. L., Chatterjee, D., Chatterjee-Chakraborty, M., maternal mRNA regulation: Implications for mamma- Lovic, V., Grella, S. L., Steiner, M., et al. (2007). Prenatal lian early embryonic development. Frontiers in Biosci- restraint stress and motherless rearing disrupts expres- ence, 12, 3713–3726. sion of plasticity markers and stress-induced corticoste- Epigenetics 71

rone release in adult female Sprague–Dawley rats. development of the offspring in a rodent model. Biolog- Brain Research, 1158, 28–38. ical Psychiatry, 59, 1227–1235. Caldji, C., Diorio, J., & Meaney, M. J. (2003). Variations in Champagne, F. A., Stevenson, C., Gratton, A., & Meaney, maternal care alter GABAA receptor subunit expres- M. J. (2004). Individual differences in maternal behav- sion in brain regions associated with fear. Neuropsycho- ior are mediated by dopamine release in the nucleus pharmacology, 28, 150–159. accumbens. Journal of Neuroscience, 24, 4113–4123. Caldji, C., Francis, D. D., Sharma, S., Plotsky, P. M., & Mea- Champagne, D. L., van Hasselt, F., Ramakers, G., Mea- ney, M. J. (2000). The effects of early rearing environment ney, M. J., de Kloet, E. R., Joels, M., et al. (2008). Mater- on the development of GABAA and central benzodiaze- nal care alters dendritic length, spine density and pine receptor levels and novelty-induced fearfulness in synaptic potentiation in adulthood. Journal of Neurosci- the rat. Neuropsychopharmacology, 22, 219–229. ence, 28, 6037–6045. Caldji, C., Tannenbaum, B., Sharma, S., Francis, D. D., Champagne, F. A., Weaver, I. C. G., Diorio, J., Dymov, S., Plotsky, P. M., & Meaney, M. J. (1998). Maternal care Szyf, M., & Meaney, M. J. (2006). Maternal care regu- during infancy regulates the development of neural lates methylation of the estrogen receptor alpha 1b pro- systems mediating the expression of behavioral fearful- moter and estrogen receptor alpha expression in the ness in adulthood in the rat. Proceedings of the National medial preoptic area of female offspring. Endocrinology, Academy of Sciences, USA, 95, 5335–5340. 147, 2909–2915. Cameron, N., Parent, C., Champagne, F. A., Fish, E., Champagne, F. A., Weaver, I. C. G., Diorio, J., Sharma, S., Ozaki-Kuroda, K., & Meaney, M. J. (2005). The pro- & Meaney, M. J. (2003). Natural variations in maternal gramming of individual differences in defensive care are associated with estrogen receptor alpha responses and reproductive strategies in the rat expression and estrogen sensitivity in the MPOA. Endo- through variations in maternal care. Neuroscience & Bio- crinology, 144, 4720–4724. behavioral Reviews, 29, 843–865. Champoux, M., Bennett, A., Shannon, C., Higley, J. D., Canli, T., & Lesch, K.-P. (2007). Long story short: The Lesch, K. P., & Suomi, S. J. (2002). Serotonin transporter serotonin transporter in emotion regulation and social gene pollymorphism, differential early rearing, and cognition. Nature Neuroscience, 10, 1103–1109. behavior in rhesus monkey neonates. Molecular Psychia- Canli, T., Qiu, M., Omura, K., Congdon, E., Haas, B. W., try, 7, 1058–1063. Amin, Z., et al. (2006). Neural correlates of . Charalambous, M., da Rocha, S. T., & Ferguson-Smith, A. Proceedings of the National Academy of Sciences, USA, 103, C. (2007). , growth control and the 16033–16038. allocation of nutritional resources: Consequences for Carola, V., Frazzetto, G., Pascucci, T., Audero, E., Puglisi- postnatal life. Current Opinion in Endocrinology, Diabetes Allegra, S., Cabib, S., et al. (2008). Identifying molecular & Obesity, 14, 3–12. substrates in a mouse model of the serotonin trans- Chong, S., & Whitelaw, E. (2004). Epigenetic germline porter · environment risk factor for anxiety and inheritance. Current Opinion in Genetics and Develop- depression. Biological Psychiatry, 63, 840–846. ment, 14, 692–696. Caspi, A., & Moffit, T. E. (2006). Gene–environment inter- Chrousos, G. P., & Gold, P. W. (1992). The concepts of actions in psychiatry: Joining forces with neuroscience. stress and stress system disorders. Journal of the Ameri- Nature Reviews Neuroscience, 7, 583–590. can Medical Association, 267, 1244–1252. Caspi, A., Sugden, K., Moffitt, T. E., Taylor, A., Craig, I. Conger, R., Ge, X., Elder, G., Lorenz, F., & Simons, R. W., Harrington, H., et al. (2003). Influence of life stress (1994). Economic stress, coercive family process and on depression: Moderation by a polymorphism in the developmental problems of adolescents. Child Develop- 5-HTT gene. Science, 301, 386–389. ment, 65, 541–561. Champagne, F. A. (2008). Epigenetic mechanisms and the Conger, R. D., McCarty, J. A., Yang, R. K., Lahey, B. B., & transgenerational effects of maternal care. Frontiers in Kropp, J. P. (1984). Perception of child, childrearing Neuroendocrinology, 29, 386–397. values, and emotional distress as mediating links Champagne, F. A., Diorio, J., Sharma, S., & Meaney, M. J. between environmental stressors and observed mater- (2001). Variations in maternal care in the rat are associ- nal behavior. Child Development, 55, 2234–2247. ated with differences in estrogen-related changes in Cooney, C. A., Dave, A. A., & Wolff, G. L. (2002). Mater- oxytocin receptor levels. Proceedings of the National nal methyl supplements in mice affect epigenetic varia- Academy of Sciences, USA, 98, 12736–12741. tion and DNA methylation of offspring. Journal of Champagne, F. A., Francis, D. D., Mar, A., & Meaney, , 132, 2393–2400. M. J. (2003). Naturally-occurring variations in maternal Coplan, J. D., Andrews, M. W., Rosenblum, L. A., & care in the rat as a mediating influence for the effects Nemeroff, C. B. (1996). Persistent elevations of cerebro- of environment on the development of individual dif- spinal fluid concentrations of corticotropin-releasing ferences in stress reactivity. Physiology & Behavior, 79, factor in adult nonhuman primates exposed to early- 359–371. life stressors: Implications for the pathophysiology of Champagne, F. A., & Meaney, M. J. (2006). Stress during mood and anxiety disorders. Proceedings of the National gestation alters postpartum maternal care and the Academy of Sciences, USA, 93, 1619–1623. 72 Meaney

Crosby, S. D., Puetz, J. J., Simburger, K. S., Fahrner, T. J., Etkin, A., Egner, T., Peraza, D. M., Kande, E. R., & Hir- & Milbrandt, J. (1991). The early response gene NGFI-C sch, J. (2005). Resolving emotional conflict: A role for encodes a zinc finger transcriptional activator and is a the rostral anterior cingulate cortex in modulating member of the GCGGGGGCG (GSG) element-binding activity in the amygdala. , 51, 871–882. protein family. Molecular and Cellular Biology, 11, 3835– Etkin, A., Klemenhagen, K. C., Dudman, J. T., Rogan, M. 3841. T., Hen, R., Kandel, E. R., et al. (2004). Individual dif- da Rocha, S. T., & Ferguson-Smith, A. C. (2004). Genomic ferences in trait anxiety predict the response of the ba- imprinting. Current Biology, 14, R646–R649. solateral amygdala to unconsciously processed fearful Dallman, M. F., Akana, S. F., Strack, A. M., Hanson, E. S., faces. Neuron, 44, 1043–1055. & Sebastian, R. J. (1995). The neural network that regu- Fahrbach, S. E., & Pfaff, D. W. (1986). Effect of preoptic lates energy balance is responsive to glucocorticoids region implants of dilute estradiol on the maternal and insulin and also regulates HPA axis responsivity at behavior of ovariectomized, nulliparous rats. Hormones a site proximal to CRF neurons. Annals of the New York & Behavior, 20, 354–363. Academy of Sciences, 771, 730–742. Falk, R. (1986). What is a gene? Studies in the History and Dallman, M. F., Pecoraro, N. C., & la Fleur, S. E. (2005). Philosophy of Science, 17, 133–173. Chronic stress and comfort foods: Self-medication and Fanselow, M. S. (2000). Contextual fear, gestalt , abdominal obesity. Brain Behavior and Immunity, 19, and the hippocampus. Behavioural Brain Research, 110, 275–280. 73–81. Davis, M. (2006). Neural systems involved in fear and Farrington, D. A., Gallagher, B., Morley, L., St Ledger, R. anxiety measured with fear-potentiated startle. Ameri- J., & West, D. J. (1988). Are there any successful men can , 61, 741–756. from criminogenic backgrounds? Psychiatry, 51, 116– de Kloet, E. R., Karst, H., & Joels, M. (2008). Corticoste- 130. roid hormones in the central stress response: Quick- Feinberg, A. P. (2007). Phenotypic plasticity and the epi- and-slow. Frontiers in Neuroendocrinology, 29, 268–272. genetics of human disease. Nature, 447, 433–440. Denenberg, V. H. (1964). Critical periods, stimulus input, Felitti, V. J., Anda, R. F., Nordenberg, D., Williamson, D. and emotional reactivity: A theory of infantile stimula- F., Spitz, A. M., Edwards, V., et al. (1998). Relationship tion. Psychological Reviews, 71, 335–351. of childhood abuse and household dysfunction to Diamond, M. I., Miner, J. N., Yoshinaga, S. K., & Yamam- many of the leading causes of death in adults. American oto, K. R. (1990). Transcription factor interactions: Journal of Preventive Medicine, 14, 245–258. Selectors of positive or negative regulation from a sin- Fisher, P. A., Gunnar, M. R., Chamberlain, P., & Reid, J. gle DNA element. Science, 249, 1266–1270. B. (2000). Preventive intervention for maltreated pre- Diamond, D. M., & Rose, G. M. (1994). Stress impairs school children: Impact on children’s behavior, neuro- LTP and hippocampal-dependent memory. Annals of endocrine activity and foster parent functioning. the New York Academy of Sciences, 746, 411–414. Journal of the American Academy of Child & Adolescent Dix, T. (1991). The affective organization of parenting Psychiatry, 39, 1356–1364. adaptive and maladaptive processes. Psychological Bul- Fleming, A. S. (1988). Factors influencing maternal letin, 110, 3–25. responsiveness in humans: Usefulness of an animal Donaldson, Z. R., & Young, L. J. (2008). Oxytocin, vaso- model. Psychoneuroendocrinology, 13, 189–212. pressin, and the of sociality. Science, 322, Fleming, A. S., O’Day, D. H., & Kraemer, G. W. (1999). 900–904. Neurobiology of mother-infant interactions: Experience Drevets, W. C., Price, J. L., Simpson, J. R. J., Todd, R. D., and central nervous system plasticity across develop- Reich, T., Vannier, M., et al. (1997). Subgenual prefron- ment and generations. Neuroscience and Biobehavioral tal cortex abnormalities in mood disorders. Nature, 386, Reviews, 23, 673–685. 824–827. Fox, N. A., Nichols, K. E., Henderson, H. A., Rubin, K., Ebstein, R. P. (2003). Relation of shyness in grade school Schmidt, L., Hamer, D., et al. (2005). Evidence for a children to the genotype for the long form of the sero- gene-environment interaction in predicting behavioral tonin transporter promoter region polymorphism. inhibition in middle childhood. Psychological Science, 16, American Journal of Psychiatry, 160, 671–676. 921–926. Ebstein, R. P. (2006). The molecular Francis, D. D., Champagne, F., & Meaney, M. J. (2000). of human personality: Beyond self-report question- Variations in maternal behaviour are associated with naires. Molecular Psychiatry, 11, 427–445. differences in oxytocin receptor levels in the rat. Journal Eden, A., Gaudet, F., Waghmare, A., & Jaenisch, R. of Neuroendocrinology, 12, 1145–1149. (2003). Chromosomal instability and tumors promoted Francis, D. D., Diorio, J., Liu, D., & Meaney, M. J. (1999). by DNA hypomethylation. Science, 300, 455. Nongenomic transmission across generations in mater- Eley, T. C., Sugden, K., Corsico, A., Gregory, A. M., nal behavior and stress responses in the rat. Science, Sham, P., McGuffin, P., et al. (2004). Gene–environment 286, 1155–1158. interaction analysis of serotonin system markers with Fride, E., Dan, Y., Gavish, M., & Weinstock, M. (1985). adolescent depression. Molecular Psychiatry, 9, 908–915. Prenatal stress impairs maternal behavior in a conflict Epigenetics 73

situation and reduces hippocampal benzodiazepine tonin uptake in human blood platelets. American Journal receptors. Life Sciences, 36, 2103–2109. of , 88, 83–87. Fries, E., Moragues, N., Caldji, C., Hellhammer, H. H., & Griffiths, P. E., & Stotz, K. (2007). Gene. In M. Ruse & D. Meaney, M. J. (2004). Preliminary evidence of altered Hull (Eds.), Cambridge companion to sensitivity to benzodiazepines as a function of maternal (pp. 85–102). Cambridge, UK: Cambridge University care in the rat. Annals of the New York Academy of Sci- Press. ences, 1032, 320–324. Griffiths, P. E., & Tabery, J. (2008). Behavioral genetics Gervai, J., Novak, A., Lakatos, K., Toth, I., Danis, I., Ro- and development: Historical and conceptual causes of nai, Z., et al. (2007). Infant genotype may moderate sen- controversy. New Ideas in Psychology, 26, 332–352. sitivity to maternal affective communications: Grindstaff, J. L., Brodie, E. D., & Ketterson, E. D. (2003). Attachment disorganization, quality of care, and the Immune function across generations: Integrating mech- DRD4 polymorphism. Social Neuroscience, 2, 307–319. anism and evolutionary process in maternal antibody Gillespie, N. A., Whitfield, J. B., Williams, B., Heath, A. transmission. Proceedings of the Royal Society B: Biological C., & Martin, N. G. (2005). The relationship between Sciences, 270, 2309–2319. stressful life events, the serotonin transporter (5- Grolnick, W. S., Gurland, S. T., DeCourcey, W., & Jacob, HTTLPR) genotype and major depression. Psychological K. (2002). Antecedents and consequences of mothers’ Medicine, 35, 101–111. autonomy support: An experimental investigation. Gluckman, P. D., & Hanson, M. A. (2004). Living with Developmental Psychology, 38, 143–155. the past: Evolution, development, and patterns of dis- Gross, C., & Hen, R. (2004). The developmental origins of ease. Science, 305, 1733–1736. anxiety. Nature Reviews Neuroscience, 5, 545–552. Gluckman, P. D., & Hanson, M. A. (2007). Developmental Gross, C., Zhuang, X., Stark, K., Ramboz, S., Oosting, R., plasticity and human disease: Research directions. Jour- Kirby, L., et al. (2002). Serotonin1A receptor acts during nal of Internal Medicine, 261, 461–471. development to establish normal anxiety-like behaviour Goldstein, L. H., Diener, M. L., & Mangelsdorf, S. C. (1996). in the adult. Nature, 416, 396–400. Maternal characteristics and social support across the Grunstein, M. (1997). Histone acetylation in chromatin transition to motherhood: Associations with maternal structure and trasncription. Nature, 389, 349–352. behavior. Journal of Family Psychology, 10, 60–71. Haapasap, J., & Tremblay, R. E. (1994). Physically aggres- Gonzalez, A., Lovic, V., Ward, G. R., Wainwright, P. E., & sive boys from ages 6 to 12: Family background. Par- Fleming, A. S. (2001). Intergenerational effects of com- enting behavior, and prediction of delinquency. Journal plete maternal deprivation and replacement stimula- of Consulting & , 62, 1044–1052. tion on maternal behavior and emotionality in female Hake, S. B., & Allis, C. D. (2006). variants and rats. Developmental Psychobiology, 38, 11–32. their potential role in indexing mammalian : Gorman, J. M., Kent, J. M., Sullivan, G. M., & Coplan, The ‘‘H3 barcode hypothesis.’’ Proceedings of the J. D. (2000). Neuroanatomical hypothesis of panic dis- National Academy of Sciences, USA, 103, 6428–6435. order revised. American Journal of Psychiatry, 157, 493– Hammock, E. A. D., & Young, L. J. (2005). 505. instability generates diversity in brain and sociobehav- Gottlieb, G. (1991). Experiential canalization of behavioral ioral traits. Science, 308, 1630–1634. development: Theory. Developmental Psychology, 27, Hane, A. A., & Fox, N. A. (2006). Ordinary variations in 4–13. maternal caregiving influence human infants’ stress Gottlieb, G. (1997). Synthesizing nature–nurture. Mahwah, reactivity. Psychological Sciences, 17, 550–556. NJ: Erlbaum. Hariri, A. R., Drabant, E. M., Munoz, K. E., Kolachana, B. Gottlieb, G. (1998). Normally occurring environmental S., Mattay, V. S., Egan, M. F., et al. (2005). A suscepti- and behavioral influences on gene activity: From cen- bility gene for affective disorders and the response of tral dogma to probabilistic epigenesis. Psychological the human amygdala. Archives of General Psychiatry, 62, Reviews, 105, 792–892. 146–152. Grabe, H. J., Lange, M., Wolff, B., Vo¨lzke, H., Lucht, M., Hariri, A. R., Drabant, E. M., & Weinberger, D. R. (2006). Freyberger, J. H., et al. (2005). Mental and physical dis- Imaging genetics: Perspectives from studies of geneti- tress is modulated by a polymorphism in the 5-HT cally driven variation in serotonin function and cortico- transporter gene interacting with social stressors and limbic affective processing. Biological Psychiatry, 59, chronic disease burden. Molecular Psychiatry, 10, 220– 888–897. 224. Hariri, A. R., Mattay, V. S., Tessitore, A., Kolachana, B., Grayson, D. R., Jia, X., Chen, Y., Sharma, R. P., Mitchell, Fera, F., Goldman, D., et al. (2002). Serotonin trans- C. P., Guidotti, A., et al. (2005). promoter hyper- porter and the response of the human methylation in . Proceedings of the National amygdala. Science, 297, 400–403. Academy of Sciences, USA, 102, 9341–9346. Hasselquist, D., & Nilsson, J.-A. (2009). Maternal transfer Greenberg, B. D., Tolliver, T. J., Huang, S. J., Li, Q., Ben- of antibodies in vertebrates: Trans-generational effects gel, D., & Murphy, D. L. (1999). Genetic variation in on offspring immunity. Philosophical Transactions of the the serotonin transporter promoter region affects sero- Royal Society B: Biological Sciences, 364, 51–60. 74 Meaney

Heim, C., & Nemeroff, C. B. (2001). The role of childhood Joels, M., Karst, H., Krugers, H. J., & &. Lucassen, P. J. trauma in the neurobiology of mood and anxiety disor- (2007). Chronic stress: Implications for neuronal mor- ders: Preclinical and clinical studies. Biological Psychia- phology, function and neurogenesis. Frontiers in Neuro- try, 49, 1023–1039. endocrinology, 28, 72–96. Heim, C., Newport, D. J., Heit, S., Graham, Y. P., Wilcox, Johannsen, W. (1911). The genotype conception of hered- M., Bonsall, R., et al. (2000). Pituitary–adrenal and ity. The American Naturalist, 45, 129–159. autonomic responses to stress in women after sexual Jokela, M., Keltikangas-Jarvinen, L., Kivimaki, M., Putto- and physical abuse in childhood. Journal of the American nen, S., Elovainio, M., Rontu, R., et al. (2007). Serotonin Medical Association, 284, 592–597. receptor 2A gene and the influence of childhood mater- Heinz, A., Braus, D. F., Smolka, M. N., Wrase, J., Puls, I., nal nurturance on adulthood depressive symptoms. Hermann, D., et al. (2005). Amygdala-prefrontal cou- Archives of General Psychiatry, 64, 356–360. pling depends on a genetic variation of the serotonin Jutapakdeegul, N., Casalotti, S. O., Govitrapong, P., & transporter. Nature Neuroscience, 8, 20–21. Kotchabhakdi, N. (2003). Postnatal touch stimulation Heitzer, M. D., Wolf, I. M., Sanchez, E. R., Witchel, S. F., acutely alters corticosterone levels and glucocorticoid & DeFranco, D. B. (2007). Glucocorticoid receptor phys- receptor gene expression in the neonatal rat. Develop- iology. Reviews in Endocrine & Metabolic Disorders, 8, mental Neuroscience, 25, 26–33. 321–330. Kaufman, J., Yang, B. Z., Douglas-Palumberi, H., Housh- Hellman, A., & Chess, A. (2007). Gene body-specific yar, S., Lipschitz, D., Krystal, J. H., et al. (2004). Social methylation on the active X chromosome. Science, 315, supports and serotonin transporter gene moderate 1141–1143. depression in maltreated children. Proceedings of the Hellstrom, I. C., Zhang, T. Y., Diorio, J., & Meaney, M. J. National Academy of Sciences, USA, 101, 17316–17321. (2009). Maternal regulation of glucocorticoid receptor Kendler, K. S., Kuhn, J., & Prescott, C. A. (2004). The expression through serotonin-induced nerve growth factor- interrelationship of neuroticism, , and stressful life inducible factor A signaling. Unpublished manuscript. events in the prediction of episodes of major depres- Hettema, J. M., Neale, M. C., Myers, J. M., Prescott, C. A., sion. American Journal of Psychiatry, 161, 631–636. & Kendler, K. S. (2006). A population-based Kendler, K. S., Kuhn, J. W., Vittum, J., Prescott, C. A., & of the relationship between neuroticism and internaliz- Riley, B. (2005). The interaction of stressful life events ing disorders. American Journal of Psychiatry, 163, 857– and a serotonin transporter polymorphism in the pre- 864. diction of episodes of major depression: A replication. Hinde, R. A. (1986). Some implications of evolutionary Archives of General Psychiatry, 62, 529–535. theory and comparative data for the study of human Kim, J. J., & Yoon, K. S. (1998). Stress: Metaplastic effects prosocial and aggressive behaviour. In D. Olweus, J. in the hippocampus. Trends in , 21, 505– Block, & M. Radke-Yarrow (Eds.), Development of anti- 509. social and prosocial behaviour (pp. 13–32). Orlando, FL: Kinsley, C. H., Mann, P. E., & Bridges, R. S. (1998). Prena- Academic Press. tal stress alters - and stress induced analgesia Hofer, M. A. (2005). The psychobiology of early attach- in male and female rats. Pharmacology, Biochemistry and ment. Clinical Neuroscience Research, 4, 291–300. Behavior, 50, 413–419. Holliday, R. (1989). DNA methylation and epigenetic Kirkwood, A., Dudek, S. M., Gold, J. T., Aizenman, C. D., mechanisms. Cell , 15, 15–20. & Bear, M. F. (1993). Common forms of synaptic plas- Holsboer, F. (2000). The corticosteroid receptor hypothe- ticity in the hippocampus and neocortex in vitro. sis of depression. Neuropsychopharmacology, 23, 477–501. Science, 260, 1518–1521. Hu, X., Oroszi, G., Chun, J., Smith, T. L., Goldman, D., & Klose, R. J., & Bird, A. P. (2006). Genomic DNA methyla- Schuckit, M. A. (2005). An expanded evaluation of the tion: The mark and its mediators. Trends in Biochemical relationship of four alleles to the level of response to Sciences, 31, 89–97. alcohol and the risk. Alcohol Clinical Experi- Koob, G. F., Heinrichs, S. C., Menzaghi, F., Pich, E. M., & mental Research, 29, 8–16. Britton, K. T. (1994). Corticotropin-releasing factor, Jablonka, E., & Lamb, M. J. (2005). Evolution in four dimen- stress and behavior. Seminars in Neuroscience, 6, 221– sions: Genetic, epigenetic, behavioral, and symbolic variation 229. in the . Cambridge, MA: MIT Press. Kumar, R., & Thompson, E. B. (2005). Gene regulation by Jenuwein, T., & Allis, C. D. (2001). Translating the histone the glucocorticoid receptor: Structure–function relation- code. Science, 293, 1074–1080. ship. Journal of Steroid Biochemistry & Molecular Biology, Jiang, Y., Langley, B., Lubin, F. D., Renthal, W., Wood, 94, 383–394. M. A., Yasui, D. H., et al. (2008). Epigenetics in the LaBar, K. S., Gatenby, J. C., LeDoux, J. E., et al. (1998). nervous system. Journal of Neuroscience, 28, 11753– Human amygdala activation during conditioned fear 11759. acquisition and extinction—A mixed-trial fMRI study. Jirtle, R. L., & Skinner, M. K. (2007). Environmental epige- Neuron, 205, 937–945. nomics and disease susceptibility. Nature Reviews Genet- LaBar, K. S., LeDoux, J. E., Spencer, D. D., et al. (1995). ics, 8, 253–262. Impaired fear conditioning following unilateral tempo- Epigenetics 75

ral lobectomy in humans. Journal of Neuroscience, 15, Lu, N. Z., & Cidlowski, J. A. (2006). Glucocorticoid recep- 6846–6855. tor isoforms generate transcription specificity. Trends in Laird, P. W. (2005). . Human Molecular Cell Biology, 16, 301–307. Genetics, 14(Spec. No. 1), R65–R76. Lubin, F. D., Roth, T. L., & Sweatt, J. D. (2008). Epigenetic Lakatos, K., Nemoda, Z., Birkas, E., Ronai, Z., Kovacs, E., regulation of bdnf gene transcription in the consolida- Ney, K., et al. (2003). Association of D4 dopamine tion of fear memory. Journal of Neuroscience, 28, 10576– receptor gene and serotonin transporter promoter 10586. polymorphisms with infants’ response to novelty. Lucki, I. (1998). The spectrum of behaviors influenced by Molecular Psychiatry, 8, 90–97. serotonin. Biological Psychiatry, 44, 151–162. Lakatos, K., Nemoda, Z., Toth, I., Ronai, Z., Ney, K., Sas- Luger, K., Mader, A. W., Richmond, R. K., Sargent, D. F., vari- Szekely, M., et al. (2002). Further evidence for the & Richmond, T. J. (1997). Crystal structure of the nucle- role of the dopamine D4 receptor (DRD4) gene in osome coreparticle at 2.8 A resolution. Nature, 389, 251– attachment disorganization: Interaction of the exon III 260. 48-bp repeat and the -521 C ⁄ T promoter polymor- Maestripieri, D. (2005). Early experience affects the inter- phisms. Molecular Psychiatry, 7, 27–31. generational transmission of infant abuse in rhesus LeDoux, J. E. (2000). Emotion circuits in the brain. Annual monkeys. Proceedings of the National Academy of Sciences, Review of Neuroscience, 23, 155–184. USA, 102, 9726–9729. Lee, M. H., & Williams, D. I. (1977). A longitudinal study Maren, S., & Quirk, G. J. (2004). Neuronal signalling of of mother-young interaction in the rat: The effects of fear memory. Nature Reviews Neuroscience, 5, 844–852. infantile stimulation, diurnal rhythms, and pup matu- Martinowich, K., Hattori, D., Wu, H., Fouse, S., He, F., ration. Behaviour, 63, 241–261. Hu, Y., et al. (2003). DNA methylation-related chroma- Lesch, K., Bengel, D., Heils, A., Sabol, S., Greenberg, B., tin remodeling in activity-dependent BDNF gene regu- Petri, S., et al. (1996). Association of anxiety-related lation. Science, 302, 890–893. traits with a polymorphism in the serotonin transporter McCauley, J., Kern, D. E., Kolodner, K., Dill, L., Schroe- gene regulatory region. Science, 274, 1527–1531. der, A. F., DeChant, H. K., et al. (1997). Clinical charac- Lesch, K. P., Meyer, J., Glatz, K., Flu¨ gge, G., Hinney, A., teristics of women with a history of childhood abuse: Hebebrand, J., et al. (1997). The 5-HT transporter gene- Unhealed wounds. Journal of the American Medical Asso- linked polymorphic region (5-HTTLPR) in evolutionary ciation, 277, 1362–1368. perspective: Alternative biallelic variation in rhesus McCormick, J. A., Lyons, V., Jacobson, M. D., Noble, J., monkeys. Journal of Neural Transmission, 104, 1259–1266. Diorio, J., Nyirenda, M., et al. (2000). 5¢Heterogeneity Leserman, J., Drosman, D. A., Li, Z., Toomey, T. C., of glucocorticoid receptor messenger ribonucleic acid is Nachman, G., & Glogau, L. (1996). Sexual and physical tissue-specific: Differential regulation of variant tran- abuse history in gasteroenterology practice: How types scripts by early life events. Molecular Endocrinology, 14, of abuse impact health status. Psychosomatic Medicine, 506–517. 58, 4–15. McEwen, B. S. (2003). Early life influences on life-long Levine, S. (1970). The pituitary–adrenal system and the patterns of behavior and health. Mental Retardation & developing brain. Progress in Brain Research, 32, 79–85. Developmental Disabilities Research Reviews, 9, 149–154. Levine, S. (1994). The of the hypothalamic–pitu- McEwen, B. S. (2007). Physiology and neurobiology of itary–adrenal axis. The influence of maternal factors. stress and adaptation: Central role of the brain. Physio- Annals of the New York Academy of Sciences, 746, 275– logical Reviews, 87, 873–904. 288. McGirr, A., Renaud, J., Seguin, M., Alda, M., & Turecki, Lewontin, R. C. (1974). The analysis of variance and the G. (2008). Course of major depressive disorder and sui- analysis of causes. American Journal of Human Genetics, cide outcome: A psychological autopsy study. Journal of 26, 400–411. Clinical Psychiatry, 69, 966–970. Lissau, J., & Sorensen, T. I. A. (1994). Parental neglect McGowan, P. O., Sasaki, A., Dymov, S., LaBoonte´, B., during childhood and increased risk of obesity in Turecki, G., Szyf, M., et al. (2009). Epigenetic regulation young adulthood. Lancet, 343, 324–327. of the glucocorticoid receptor in human brain associ- Liu, D., Diorio, J., Day, J. C., Francis, D. D., Mar, A., & ates with childhood abuse. Nature Neuroscience, 12, 342– Meaney, M. J. (2000). Maternal care, hippocampal syna- 348. ptogenesis and cognitive development in the rat. Nat- McLloyd, V. C. (1998). Socioeconomic disadvantage and ure Neuroscience, 3, 799–806. child development. American Psychologist, 53, 185–204. Liu, D., Diorio, J., Tannenbaum, B., Caldji, C., Francis, D. McNally, J. G., Mu¨ ller, W. G., Walker, D., Wolford, R., & D., Freedman, A., et al. (1997). Maternal care, hippo- Hager, G. L. (2000). The glucocorticoid receptor: Rapid campal glucocorticoid receptors and HPA responses to exchange with regulatory sites in living cells. Science, stress. Science, 277, 1659–1662. 287, 1262–1265. Lovic, V., Gonzalez, A., & Fleming, A. S. (2001). Mater- Meaney, M. J. (2001). The development of individual dif- nally separated rats show deficits in maternal care in ferences in behavioral and endocrine responses to adulthood. Developmental Psychobiology, 39, 19–33. stress. Annual Review of Neuroscience, 24, 1161–1192. 76 Meaney

Meaney, M. J. (2007). Social influences on sexual differ- Mousseau, T. A., & Fox, C. W. (1998). The adaptive sig- entiation in mammals. Advances in Genetics, 59, 173– nificance of maternal effects. Trends in Ecology and Evo- 215. lution, 13, 403–407. Meaney, M. J., Diorio, J., Donaldson, L., Yau, J., Chap- Munafo, M. R., Clark, T., & Flint, J. (2005). Does measure- man, K., & Seckl, J. R. (2000). Handling alters the ment instrument moderate the association between the expression of messenger RNAs for AP-2, NGFI-A and serotonin transporter gene and anxiety related person- NGFI-B in the hippocampus of neonatal rats. Journal of ality traits? A meta-analysis. Molecular Psychiatry, 10, Neuroscience, 20, 3936–3945. 415–419. Meaney, M. J., LeDoux, J. E., & Leibowitz, M. L. (2008). Munck, A. (2005). Glucocorticoid receptors and physiol- Neurobiology of anxiety disorders. In A. Tasman, J. ogy: A personal history. Steroids, 70, 335–344. Kay, J. A. Lieberman, M. B. First, & M. Maj (Eds.). Psy- Munck, A., Guyre, P. M., & Holbrook, N. J. (1984). Physi- chiatry, 1, 317–338. ological functions of glucocorticoids in stress and their Meaney, M. J., & Szyf, M. (2005). Maternal effects as a relation to pharmacological actions. Endocrine Reviews, model for environmentally-dependent chromatin plas- 5, 25–44. ticity. Trends in Neuroscience, 28, 456–463. Murayama, A., Sakura, K., Nakama, M., Yasuzawa- Meaney, M. J., Szyf, M., & Seckl, J. R. (2007). Epigenetic Tanaka, K., Fujita, E., Tateishi, Y., et al. (2006). A mechanisms of perinatal programming of hypotha- specific CpG site demethylation in the human interleu- lamic–pituitary–adrenal function and health. Trends in kin 2 gene promoter is an epigenetic memory. EMBO Molecular Medicine, 13, 269–277. Journal, 25, 1081–1092. Melo, A. I., Lovic, V., Gonzalez, A., Madden, M., Sinopo- Neigh, G. N., & Nemeroff, C. B. (2006). Reduced gluco- li, K., & Fleming, A. S. (2006). Maternal and littermate corticoid receptors: Consequence or cause of depres- deprivation disrupts maternal behavior and social- sion? Trends in Endocrinology & Metabolism, 17, 124–125. learning of food preference in adulthood: Tactile stimu- Nissen, R. M., & Yamamoto, K. R. (2000). The glucocorti- lation, nest odor, and social rearing prevent these coid receptor inhibits NFkappaB by interfering with effects. Developmental Psychobiology, 48, 209–219. serine-2 phosphorylation of the RNA polymerase II car- Menard, J., Champagne, D., & Meaney, M. J. (2004). boxy-terminal domain. Genes & Development, 14, 2314– Maternal care alters behavioral and neural activity pat- 2329. terns in the defensive burying paradigm. Neuroscience, Nobukuni, Y., Smith, C. L., Hager, G. L., & Detera- 129, 297–308. Wadleigh, S. D. (1995). Characterization of the human Miesfield, R., Okret, S., Wikstrom, A.-C., Wrange, O., glucocorticoid receptorpromoter. Biochemistry, 34, 8207– Gustafsson, A. J., & Yamamoto, K. R. (1984). Charac- 8214. terization of a steroid hormone receptor gene and Novakov, M., & Fleming, A. S. (2005). The effects of early mRNA in wild-type and cells. Nature, 312, rearing environment on the hormonal induction of 779–781. maternal behavior in virgin rats. Hormones & Behavior, Miller, C. A., & Sweatt, J. D. (2007). Covalent modifica- 48, 528–536. tion of DNA regulates memory formation. Neuron, 53, Numan, M. J., & Insel, T. R. (2003). The neurobiology of 857–869. parental behavior (3rd ed.). New York: Springer. Mitchell, J. B., Betito, K., Rowe, W., Boksa, P., & Meaney, Offord, D. R., Boyle, M. H., Racine, Y. A., Fleming, J. E., M. J. (1992). regulation of type II cortico- Cadman, D. T., Blum, H. M., et al. (1992). Outcome, steroid receptor binding in cultured hippocampal cells: prognosis, and risk in a longitudinal follow-up study. The role of serotonin-induced increases in cAMP levels. Journal of the American Academy of Child and Adolescent Neuroscience, 48, 631–639. Psychiatry, 31, 916–923. Mitchell, J. B., Rowe, W., Boksa, P., & Meaney, M. J. Olds, D., Henderson, C. R., Cole, R., Eckenrode, J., Kitz- (1990). Serotonin regulates type II, corticosteroid recep- man, H., Luckey, D., et al. (1998). Long-term effects of tor binding in cultured hippocampal cells. Journal of nurse home visitation on children’s criminal and anti- Neuroscience, 10, 1745–1752. social behavior: 15-year follow-up of a randomized Mohandas, T., Sparkes, R. S., & Shapiro, L. J. (1981). Reac- controlled trial. Journal of the American Medical Associa- tivation of an inactive human X chromosome: Evidence tion, 280, 1238–1244. for X inactivation by DNA methylation. Science, 211, Pedersen, C. A. (1997). Oxytocin control of maternal 393–396. behavior. Regulation by sex steriods and offspring Moore, C. L., & Power, K. L. (1986). Prenatal stress affects stimuli. Annals of the New York Academy of Sciences, 614, mother-infant interaction in Norway rats. Developmental 126–145. Psychobiology, 19, 235–245. Pe´rez-Edgar, K., & Fox, N. A. (2005). Temperament and Morris, J. S., Ohman, A., & Dolan, R. J. (1998). Conscious anxiety disorders. Child and Adolescent Psychiatric Clinics and unconscious emotional learning in the human of North America, 14, 681–706. amygdala. Nature, 393, 467–470. Pezawas, L., Meyer-Lindenberg, A., Drabant, E. M., Ver- Moss, L. (2005). What genes can’t do. Cambridge, MA: MIT chinski, B. A., Munoz, K. E., Kolachana, B. S., et al. Press. (2005). 5-HTTLPR polymorphism impacts human cin- Epigenetics 77

gulate-amygdala interactions: A genetic susceptibility Robinson, G. E., Fernald, R. D., & Clayton, D. F. (2008). mechanism for depression. Nature Neuroscience, 8, 828– Genes and social behavior. Science, 322, 896–900. 834. Rosen, J. B., & Schulkin, J. (1998). From normal fear to Plomin, R., DeFries, J. C., & McClearn, G. E. (1990). Behav- pathological anxiety. Psychological Review, 105, 325–350. ioral genetics. A primer. New York: Freeman. Rossiter, M. C. (1998). The role of environmental varia- Plomin, R., & Rutter, M. (1998). Child development, tion in parental effects expression. In T. A. Mousseau & , and what to do with genes once C. W. Fox (Eds.), Maternal effects as (pp. 112– they are found. Child Development, 69, 1223–1242. 136). Oxford, UK: Oxford University Press. Plotsky, P. M., Cunningham, E. T., Jr., & Widmaier, E. P. Roth, T. L., Lubin, F. D., Funk, A. J., & Sweatt, J. D. (2009). (1989). Catecholaminergic modulation of corticotropin- Lasting epigenetic influence of early-life adversity on releasing factor and adrenocorticotropin . the BDNF gene. Biological Psychiatry, 65, 760–769. Endocrine Reviews, 10, 437–458. Russak, L. G., & Schwartz, G. E. (1997). Feelings of paren- Propper, C., Willoughby, M., Halpern, C. T., Carbone, M. tal care predict health status in midlife: A 35 year fol- A., & Cox, M. (2007). Parenting quality, DRD4, and the low-up of the Harvard Mastery of Stress Study. Journal prediction of externalizing and internalizing behaviors of Behavioral Medicine, 20, 1–11. in early childhood. Developmental Psychobiology, 49, Rutter, M. (2007). Gene–environment interdependence. 619–632. Developmental Science, 10, 12–18. Pruessner, J. L., Champagne, F. A., Meaney, M. J., & Da- Sapolsky, R. M., & Meaney, M. J. (1986). The maturation gher, A. (2004). Parental care and neuroendocrine and of the adrenocortical stress response in the rat. Brain dopamine responses to stress in humans: A PET imag- Research Reviews, 11, 65–76. ing study. Journal of Neuroscience, 24, 2825–2831. Sapolsky, R. M., Romero, L. M., & Munck, A. U. (2000). Ptak, C., & Petronis, A. (2008). Epigenetics and complex How do glucocorticoids influence stress responses? disease: From etiology to new therapeutics. Annual Integrating permissive, suppressive, stimulatory, and Review of Pharmacology & Toxicology, 48, 257–276. preparative actions. Endocrine Reviews, 21, 55–89. Raison, C. L., & Miller, A. H. (2003). When not Schanberg, S. M., Evoniuk, G., & Kuhn, C. M. (1984). Tac- enough is too much: The role of insufficient gluco- tile and nutritional aspects of maternal care: Specific corticoid signaling in the pathophysiology of stress- regulators of neuroendocrine function and cellular related disorders. American Journal of Psychiatry, 160, development. Proceedings of the Society for Experimental 1554–1565. Biology & Medicine, 175, 135–146. Rassoulzadegan, M., Grandjean, V., Gounon, P., Vincent, Schatz, C. J. (1990). Impulse activity and the patterning of S., Gillot, I., & Cuzin, F. (2006). RNA-mediated non- connections during CNS development. Neuron, 5, 745– mendeliaqn inheritance of an epigenetic change in the 756. mouse. Nature, 441, 469–474. Schatzberg, A. F., Rothschild, A. J., Langlais, P. J., Bird, E. Razin, A. (1998). CpG methylation, chromatin structure D., & Cole, J. O. (1985). A corticosteroid ⁄ dopamine and gene silencing-a three-way connection. EMBO Jour- hypothesis for psychotic depression and related states. nal, 17, 4905–4908. Journal of Psychiatric Research, 19, 57–64. Razin, A., & Cedar, H. (1993). DNA methylation: Molecu- Schinka, J. A., Busch, R. M., & Robichaux-Keene, N. lar biology and biological significance. In J. P. Jost & H. (2004). A meta-analysis of the association between the P. Saluz (Eds.), DNA methylation: Molecular biology and serotonin transporter gene polymorphism (5-HTTLPR) biological significance (pp. 343–359). Basel, Switzerland: and trait anxiety. Molecular Psychiatry, 9, 197–202. Birkha¨user Verlag. Schneirla, T. C. (1966). Behavioral development and com- Razin, A., & Riggs, A. D. (1980). DNA methylation and parative psychology. Quarterly Review of Biology, 41, gene function. Science, 210, 604–610. 283–302. Reik, W. J. (2001). Genomic imprinting: Parental influence Seckl, J. R. (2001). Glucocorticoid programming of the on the genome. Nature, 293, 1089–1096. fetus adult phenotypes and molecular mechanisms. Renthal, W., Maze, I., Krishnan, V., Covington, H. E., III, Molecular and Cellular Endocrinology, 185, 61–71. Xiao, G., Kumar, A., et al. (2007). Histone deacetylase 5 Seckl, J. R., & Holmes, M. C. (2007). Mechanisms of dis- epigenetically controls behavioral adaptations to ease: Glucocorticoids, their placental metabolism and chronic emotional stimuli. Neuron, 56, 517–529. fetal ‘‘programming’’ of adult pathophysiology. Nature Renthal, W., & Nestler, E. J. (2008). Epigenetic mecha- Clinical Practice Endocrinology & Metabolism, 3, 479–488. nisms in drug . Trends in Molecular Medicine, Seckl, J. R., & Meaney, M. J. (1994). Early life events and 14, 341–350. later development of ischaemic heart disease. Lancet, Repetti, R. L., Taylor, S. E., & Seeman, T. E. (2002). Risky 342, 1236. families: Family social environments and the mental Sen, S., Burmeister, M., & Ghosh, D. (2004). Meta-analysis and physical health of offspring. Psychological Bulletin, of the association between a serotonin transporter pro- 128, 330–366. moter polymorphism (5-HTTLPR) and anxiety-related Riggs, A. D., & Pfeifer, G. P. (1992). X-chromosome inacti- personality traits. American Journal of Medical Genetics, vation and cell memory. Trends in Genetics, 8, 169–174. 127B, 85–89. 78 Meaney

Shahrohk, D., Cameron, N. M., Zhang, T.-Y., Gratton, A., Szyf, M., Weaver, I. C., Champagne, F. A., Diorio, J., & & Meaney, M. J. (in press). Oxytocin–dopamine interac- Meaney, M. J. (2005). Maternal programming of steroid tions in the ventral tegmental area mediate variations receptor expression and phenotype through DNA in maternal behaviour in the rat. methylation in the rat. Frontiers in Neuroendocrinology, Shanks, N., Larocque, S., & Meaney, M. J. (1995). Neona- 26, 139–162. tal endotoxin exposure alters the development of the Tang, A. C., Akers, K. G., Reeb, B. C., Romeo, R. D., & hypothalamic–pituitary–adrenal axis: Early illness and McEwen, B. S. (2006). Programming social, cognitive, later responsivity to stress. Journal of Neuroscience, 15, and neuroendocrine development by early exposure to 376–384. novelty. Proceedings of the National Academy of Sciences, Shanks, N., Windle, R. J., Perks, P. A., Harbuz, M. S., Jes- USA, 103, 15716–15721. sop, D. S., Ingram, C. D., et al. (2000). Early-life expo- Toki, S., Morinobu, S., Imanaka, A., Yamamoto, S., Yam- sure to endotoxin alters hypothalamic-pituitary-adrenal awaki, S., & Honma, K. (2007). Importance of early function and predisposition to inflammation. Proceed- lighting conditions in maternal care by dam as well as ings of the National Academy of Sciences, USA, 97, 5645– anxiety and memory later in life of offspring. European 5650. Journal of Neuroscience, 25, 815–829. Shin, L. M., Orr, S. P., Carson, M. A., Rauch, S. L., Mack- Tsankova, N., Renthal, W., Kumar, A., & Nestler, E. J. lin, M. L., Lasko, N. B., et al. (2004). Regional cerebral (2007). Epigenetic regulation in psychiatric disorders. blood flow in the amygdala and medial prefrontal cor- Nature Reviews Neuroscience, 8, 355–367. tex during traumatic imagery in male and female Viet- Turner, B. (2001). Chromatin structure and the regulation of nam veterans with PTSD. Archives of General Psychiatry, gene expression. Oxford, UK: Blackwell. 61, 168–176. Turner, J. D., & Muller, C. P. (2005). Structure of the glu- Shin, L. M., Wright, C. I., Cannistraro, P. A., Wedig, M. cocorticoid receptor (NR3C1) gene 5¢untranslated M., McMullin, K., Martis, B., et al. (2005). A functional region: Identification and tissue distribution of multiple magnetic resonance imaging study of amygdala and new human exon 1. Journal of Molecular Endocrinology, medial prefrontal cortex responses to overtly presented 35, 283–292. fearful faces in posttraumatic stress disorder. Archives Uno, H., Tarara, R., Else, G., Suleman, M. A., & Sapolsky, of General Psychiatry, 62, 273–281. R. M. (1989). Hippocampal damage associated with Shors, T. J., Foy, M. R., Levine, S., & Thompson, R. F. prenatal glucocorticoid exposure. Developmental Brain (1990). Unpredictable and uncontrollable stress impairs Research, 9, 1705–1711. neuronal plasticity in the rat hippocampus. Brain Valentino, R. J., Curtis, A. L., Page, M. E., Pavcovich, L. Research Bulletin, 24, 663–667. A., & Florin-Lechner, S. M. (1998). Activation of the Shorter, J., & Lindquist, S. (2005). as adaptive con- locus cereulus brain noradrenergic system during duits of memory and inheritance. Nature Reviews Genet- stress: Circuitry, consequences, and regulation. ics, 6, 435–450. Advances in Pharmacology, 42, 781–784. Smith, J. W., Seckl, J. R., Evans, A. T., Costall, B., & Smy- Van den Boom, D. (1994). The influence of temperament the, J. W. (2004). Gestational stress induces post-partum and mothering on attachment and exploration: An depression-like behaviour and alters maternal care in experimental manipulation of sensitive responsiveness rats. Psychoneuroendocrinology, 29, 227–244. among lower-class mothers and irritable infants. Child Sokolowski, M. B., & Wahlsten, D. (2001). Gene–environ- Development, 65, 1457–1477. ment interaction and complex behavior. In S. O. Mol- Vardhman, K., Rakyan, V. K., Suyinn Chong, S., Marnie, din (Ed.), Methods in genomic neuroscience (pp. 3–27). E., Champ, M. E., Cuthbert, P. C., et al. (2003). Transge- Boca Raton, FL: CRC Press. nerational inheritance of epigenetic states at the murine Stutzmann, G. E., & LeDoux, J. E. (1999). GABAergic AxinFu allele occurs after maternal and paternal trans- antagonists block the inhibitory effects of serotonin in mission. Proceedings of the National Academy of Sciences, the lateral amygdala: A mechanism for modulation of USA, 100, 2538–2543. sensory inputs related to fear conditioning. Journal of Vaughn, B., Egeland, B., Sroufe, L. A., & Waters, E. Neuroscience, 19, RC8. (1979). Individual differences in infant-mother attach- Suomi, S. J. (2006). Risk, resilience, and gene · environ- ment at twelve and eighteen months: Stability and ment interactions in rhesus monkeys. Annals of the New change in families under stress. Child Development, 50, York Academy of Sciences, 1094, 83–104. 971–975. Surtees, P. G., Wainwright, N. W. J., Saffron, A. G., Lube- Wahlsten, D. (1990). Insensitivity of the analysis of vari- na, R., Day, N. E., & Flint, J. (2006). Social adversity, ance to heredity–environment interaction. Behavioral the serotonin transporter (5-HTTLPR) polymorphism and Brain Sciences, 13, 109–161. and major depressive disorder. Biological Psychiatry, 59, Wahlsten, D. (2003). Airbrushing heritability. Genes, Brain, 224–229. & Behavior, 2, 327–329. Sweatt, J. D. (2009). Experience-dependent epigenetic Walker, B. R. (2007). Glucocorticoids and cardiovascular modifications in the central nervous system. Biological disease. European Journal of Endocrinology, 157, 545– Psychiatry, 65, 191–197. 559. Epigenetics 79

Waterland, R. A., & Jirtle, R. L. (2003). Transposable ele- potential and evolutionary impact of promoter DNA ments: Targets for early nutritional effects on epigenetic methylation in the human genome. Nature Genetics, 39, gene regulation. Molecular Cell Biology, 23, 5293–5300. 457–466. Waterland, R. A., Lin, J. R., Smith, C. A., & Jirtle, R. L. West, M. J., & King, A. P. (1987). Settling nature and nur- (2006). Post-weaning diet affects genomic imprinting at ture into an ontogenetic niche. Developmental Psychobiol- the insulin-like growth factor 2 (Igf2) locus. Human ogy, 20, 549–562. Molecular Genetics, 15, 705–716. Whitelaw, N. C., & Whitelaw, E. (2006). How lifetimes Weaver, I. C. G., Cervoni, N., D’Alessio, A. C., Cham- shape epigeneotpye within and across generations. pagne, F. A., Seckl, J. R., Szyf, M., et al. (2004). Epige- Human Molecular Genetics, 15, R131–R137. netic programming through maternal behavior. Nature Yehuda, R., & Bierer, L. M. (2008). Transgenerational (Neuroscience), 7, 847–854. transmission of and PTSD risk. Progress in Brain Weaver, I. C. G., Champagne, F. A., Brown, S. E., Dymov, Research, 167, 121–135. S., Sharma, S., Meaney, M. J., et al. (2005). Reversal of Zhang, T.-Y., Bagot, R., Parent, C., Nesbitt, C., Bredy, T. maternal programming of stress responses in adult off- W., Caldji, C., et al. (2006). Maternal programming of spring through methyl supplementation: Altering epi- defensive responses through sustained effects on gene genetic marking later in life. Journal of Neuroscience, 25, expression. Biological Psychiatry, 73, 72–89. 11045–11054. Zhang, T., Haws, P., & Wu, Q. (2004). Multiple variable Weaver, I. C. G., DiAlessio, A. C., Brown, S. E., Hellstrom, first exons: A mechanism for cell- and tissue-specific I. C., Dymov, S., Diorio, J., et al. (2007). The transcrip- gene regulation. Genome Research, 14, 79–89. tion factor NGFI-A mediates epigenetic programming: Zhang, T. Y., Hellstrom, I. C., Wei, X. L., & Meaney, M. J. Altering epigenetic marking through immediate early (2009). Maternal regulation of glutamic acid decarboxylase 1 genes. Journal of Neuroscience, 27, 1756–1768. gene expression through DNA methylation. Manuscript Weber, M., Hellmann, I., Stadler, M. B., Ramos, L., Paabo, submitted for publication. S., Rebhan, M., et al. (2007). Distribution, silencing