You are on page 1of 12

Hormones and Behavior 77 (2016) 272–283

Contents lists available at ScienceDirect

Hormones and Behavior

journal homepage: www.elsevier.com/locate/yhbeh

Variable postpartum responsiveness among humans and other primates


with “cooperative breeding”: A comparative and
evolutionary perspective
Sarah B. Hrdy
Citrona Farms, 21440 County Road 87, Winters, CA 95694, USA

a r t i c l e i n f o a b s t r a c t

Article history: This article is part of a Special Issue “Parental Care”.


Received 7 August 2015
Revised 21 October 2015 Until recently, evolutionists reconstructing mother–infant bonding among human ancestors relied on nonhuman
Accepted 26 October 2015
primate models characterized by exclusively maternal care, overlooking the highly variable responsiveness ex-
Available online 28 October 2015
hibited by mothers in species with obligate reliance on allomaternal care and provisioning. It is now increasingly
Keywords:
recognized that apes as large-brained, slow maturing, and nutritionally dependent for so long as early humans
Allomaternal care were, could not have evolved unless “alloparents” (group members other than genetic parents), in addition to
Alloparental provisioning in primates parents, had helped mothers to care for and provision offspring, a rearing system known as “cooperative breed-
Cooperative breeding ing.” Here I review situation-dependent maternal responses ranging from highly possessive to permissive, tem-
Postpartum responsiveness porarily distancing, rejecting, or infanticidal, documented for a small subset of cooperatively breeding primates.
Infanticide As in many mammals, primate maternal responsiveness is influenced by physical condition, endocrinological
Abandonment priming, prior experience and local environments (especially related to security). But mothers among primates
Social support
who evolved as cooperative breeders also appear unusually sensitive to cues of social support. In addition to
more “sapient” or rational decision-making, humankind's deep history of cooperative breeding must be consid-
ered when trying to understand the extremely variable responsiveness of human mothers.
© 2015 Elsevier Inc. All rights reserved.

Adding allomothers to the maternal mix This central role of mothers was consistent with long-standing
evolutionary assumptions about early humankind's division of labor
In the second half of the twentieth century, Jay Rosenblatt, a psycho- between man-the-hunter-provider and mother-the-nurturer (Darwin,
therapist interested in human emotions, turned his attention to the physio- 1874; Kaplan et al., 2000; Lovejoy, 1981). And why not? Across traditional
logical underpinnings of maternal responses. His experiments with rats and societies, the continued presence of the mother during the first two years is
other laboratory animals ushered in motherhood's Age of Psychobiological the single best predictor of infant survival (e.g., Sear and Mace, 2008).
Enlightenment. Ensuing research in his and other laboratories illuminated Meanwhile, the model organisms selected for study—such as the rats in
the roles of hormones acting on the brain during gestation, at birth, and Rosenblatt's lab, or the macaques, baboons, chimpanzees and gorillas in
following the onset of lactation priming mothers to nurture newborns. the laboratory and field studies that Bowlby selected as templates for
Over time, Rosenblatt and colleagues also documented the roles of infantile early human parenting (1971, 228–229)—fit matrifocal presumptions. So
cues, local environments and past experiences in eliciting and modulating did caretaking in other model organisms, including mice or marmosets, or
maternal behaviors most likely to ensure infant survival and well-being for that matter, humans, provided these mothers were singly housed in
(reviewed in Lonstein et al., in press; Fleming et al., 2016-in this issue; cages, or in the human case spent their days isolated with their infant with-
Barrett and Fleming, 2011; Numan and Insel, 2003). Their findings in walled dwellings rather than in social groups. By century's end, however,
complemented the writings of another psychotherapist and psychiatrist, it was increasingly clear that among hominins living by hunting and
John Bowlby, who relied on animal (especially nonhuman primate) models gathering in African contexts more typical of what Bowlby referred to as
to inform his ideas about infant-to-mother attachments and the part these humankind's Environment of Evolutionary Adaptedness, a wider cast of
powerful relationships play in future psychological development. These characters, would have been required to insure survival of offspring.
early research programs tended to focus on mothers as the main—and as Given the roughly 13 million calories needed to provision human
is true for most mammals—the exclusive, nurturers of young. offspring between birth and age 18 or so, when children in foraging
societies begin to produce as much as they consume (Kaplan, 1994),
children had to be provisioned long past weaning and required more
E-mail address: sbh@citrona.com. than a mother by herself could provide. Given the vagaries of hunting,

http://dx.doi.org/10.1016/j.yhbeh.2015.10.016
0018-506X/© 2015 Elsevier Inc. All rights reserved.
S.B. Hrdy / Hormones and Behavior 77 (2016) 272–283 273

a father by himself also could not provide calories as reliably as growing underway even earlier, perhaps with australopithecines some three
children need to be fed (O'Connell et al., 1999). Such constraints led to million years ago. The evolutionary chronology remains murky. But so
customary sharing such that hunters not only provisioned their own far as sequence goes, logic dictates that shared care and provisioning
families, but shared with other group members, providing insurance of young must have begun to emerge before (or co-evolved with?),
against recurrent shortages critical for long-term survival (Cashdan, the doubling of brain sizes between australopiths and H. erectus (from
1990; Marlowe, 2010). Mothers received help not only from fathers ca. 450 to ~900 cm3). That is, cooperative breeding must have preceded
but from other male and female allomothers as well (Hrdy, 1999, the evolution of hyper-long childhoods and the tripling of brain size to
2009). Note that the term “alloparent” is only used when paternity is 1350 cm3 of gray matter that characterize “anatomically modern
known so researchers can ascertain that a given allomother is “other humans” (at least by 200,000 years ago) and before “behaviorally
than” a parent. modern” humans' capacities for symbolic thought and sophisticated
Among Africans still living by hunting and gathering when they language over the past 150,000 years (Fig. 1).
were first studied, meat and honey brought in by men accounted Breeding systems in which alloparents in addition to parents help to
for some 40% of calories consumed. The majority of the diet came care for and provision offspring are uncommon but scarcely unique.
from plant-foods collected by women, with older women past repro- Cooperative breeding has evolved many times across insect, avian and
ductive age contributing disproportionate shares (Hawkes et al., 1989; mammalian taxa. Among vertebrates, cooperative breeding has been
Hawkes et al., 1998; Marlowe, 2010:Fig. 5.11). Older siblings and studied longest among birds. Well known for biparental nest-building,
cousins sometimes provided gathered food such as berries, but more egg tending and care of nestlings (Lynn, 2016-in this issue; Angelier
often they helped with childcare (Crittenden and Marlowe, 2013; et al., 2015-in this issue), alloparents in addition to parents are observed
Henry, 2000). Even though mothers remained key attachment figures, to tend and provision chicks in some nine percent of 10,000 avian spe-
co-sleeping with their infants at night as in all primates, human infants cies (Cockburn, 2006). Longitudinal studies of cooperatively breeding
may be cared for by male and female group members other than birds with known individuals and data on lifetime reproductive success
their mother 30% or more of the day time as well as sometimes being have yielded rich troves of data against which to test Hamilton's Rule
briefly suckled by other lactating women (Hewlett and Winn, 2014; and other theories that explain why individuals might evolve so as to
Hrdy, 2009). In this way, infants become attached to multiple help care for someone else's offspring (Koenig and Dickinson, 2004).
allomothers rendering standard measures of attachment security such Since such theories can be applied to cooperation and conflict within
as Ainsworth's Strange Situation procedure unusable in hunter–gather- families generally, the term has been extended to humans as well
er contexts. Where familiar allomothers are almost always nearby and (Emlen, 1995). It is important to note that the descriptor “cooperative”
infants have several attachment figures, they do not object to temporary notwithstanding “cooperative breeding” does not imply absence of con-
disappearances by their mother the same way as do infants in the flict. There can be considerable competition between mothers for
Western samples typically studied by developmental psychologists allomaternal assistance; between sibs for resources; between males
(Meehan and Hawks, 2013; cf. prescient early critique by Van for mates, even indirect competition between infants for succor. Nor
IJzendoorn et al., 1992). does shared care and provisioning of young by themselves explain
cooperation in other domains.
Cooperative breeding followed by (even) longer childhoods, Cooperative breeding evolves through various routes. Typically, high
bigger brains within-group genetic relatedness is required to get it started but the
threshold for shared care of infants is already set low among primates.
New evidence and reassessments of older studies are leading to a Typically new infants are magnetically attractive to at least some
paradigm shift among evolutionary anthropologists. It is increasingly other group members and rudimentary forms of alloparental caretaking
assumed that alloparental care and especially provisioning from males are documented for a slim majority of species of the several hundred
and females other than parents, in addition to parents, was essential to species in this highly social order (Hrdy, 2009).
keep Pleistocene youngsters safe and fed (Hewlett and Lamb, 2005; Even in primate species without full-fledged cooperative breeding
Hrdy, 1999, 2009; Burkart et al., 2009; Konner, 2010; Meehan and (i.e. extensive alloparental care plus provisioning), mothers may allow
Crittenden, in press; Trevathen and Rosenberg, in press). Nutritional other group members to hold and carry their young resulting in consid-
subsidies provided by allomothers would have been particularly critical erable “parenting practice” and sensitization prior to actual parenthood.
around the time of weaning when primates are most vulnerable to For example among infant-sharing langur monkeys (Semnopithecus
malnutrition. entellus) females remain for life in natal groups composed of overlap-
The importance of allomaternal provisioning was magnified in ping generations of matrilineal kin accompanied by one or more
the case of young hominins simultaneously surviving, growing, males who enter from outside the group. Since diets are composed
and also building and maintaining increasingly large brains (Isler and largely of leaves, there is little inter-individual competition for food;
van Schaik, 2012). The significance of provisioning past weaning is dominance relations among females on average related as closely as
underscored by scans of developing human brains indicating that ener- second cousins, tend to be relaxed. Thus pre- and post-reproductive
gy demands to support synapse formation peaks between 4 and 5 years, females are available to protect and/or carry young born to other
after most hunter–gatherer children begin to be, or are, weaned and females and mothers permit them to. Mothers need not (as among
when, in many cases, their mother would be pregnant with or nursing extremely hierarchical macaques) fear that their infant will fail to be
a subsequent infant (Kuzawa et al., 2014). All apes take a long time to returned, or as among chimpanzees fear that an unrelated female will
mature, but escalating demands from this unusually greedy organ harm (even cannibalize) her infant. Langur mothers tolerate other
consuming glucose at a rate up to 66% of resting metabolism, required females taking their infants as early as the first day of life and
even more extreme trade-offs with somatic growth. Thus youngsters allomothers carry them up to 50% of daytime (Hrdy, 1977: Ch. 7).
in the line leading to the genus Homo began growing up even more Nulliparous juvenile and subadult females are especially attracted to
slowly than other apes, further extending “childhood” and increasing new babies, take them most frequently, try hardest to pacify them so as
the need for nutritional subsidies from male and female group members to keep them from complaining and attracting competing caretakers,
other than the mother (Kuzawa et al., 2014). and keep hold of borrowed babies for the most minutes (Fig. 2).
I have conservatively estimated that cooperative breeding began to Nulliparae are presumably gaining valuable caretaking experience.
emerge with Homo erectus around two million years ago (Hrdy, 2009). Allomothers may also be experientially priming themselves for subse-
Paleontologists such as Jeremy De Silva (2011, in press) hypothesize quent parenthood (cf. Storey and Ziegler, 2015-in this issue). In addition
that selection pressures favoring shared care and provisioning were to juvenile and subadult females, the second most motivated category
274 S.B. Hrdy / Hormones and Behavior 77 (2016) 272–283

Fig. 1. Artist Viktor Deak's imaginative reconstruction of a H. erectus mother passing her infant to a juvenile allomother 1.8 million years ago (copyright SBH Lit, reproduced with
permission).

of langur allomother includes adult females in the late stages of preg- allomother. Brutal as parous allomothers appear (Fig. 3), no injury to
nancy, followed by mothers who have just given birth and who briefly borrowed charges has been observed. Even an infant left on the ground
take and inspect a second infant in addition to their own. Like will be picked up by either the mother or another allomother within
Rosenblatt's rats, these langurs are presumably hormonally primed to two minutes.
be especially attracted to babies (Hrdy, 1977: Tables 7.5 and 7.9). Langur infants do not like being pulled from their mother and com-
plain loudly. Nevertheless, infant-sharing is typical for all but a few spe-
Difficulties defining solicitous vs. “abusive” care cies in the sub-family Colobinae and over half of the several hundred
species in the order Primates exhibit at least some allomaternal care.
Depending on their own reproductive state, langur allomothers Mothers who allow their infants to be taken are neither neglectful nor
exhibit varying levels of interest and solicitude. In contrast to either abusive. Rather, they are preoccupied with staying fed and benefit
inexperienced or very pregnant females, experienced parous females from having group members whose intentions are sufficiently benign
will take and inspect infants but soon tire of them. Desperately clinging that a mother can temporarily entrust care of her infant to others. In
infants may be forcibly pushed away, held upside down to inspect their this respect it is worth noting that the handful of non-infant-sharing
genitals, even pressed against the ground, or sat upon, resulting in pite- exceptions within the sub-family colobinae are all species where
ous complaints prompting retrieval by either the mother or another females move between groups and so (like chimpanzees) rarely have

Fig. 2. A pregnant nullipara takes a langur infant from a juvenile female who has been carrying it.
S.B. Hrdy / Hormones and Behavior 77 (2016) 272–283 275

or at weaning, when tantrums may be accompanied by rough but


non-injurious punishments (e.g., slapping).
Primate mothers closely resemble other mammals among whom
steroid hormones during gestation, peptide hormones after birth, and
neurological systems all collaborate to render it rewarding for mothers
to taste, smell, and remain close to their babies and bond to them
(Feldman et al., 2010; Jenkins et al., 2016-in this issue; Lonstein et al.,
in press; Numan and Insel, 2003; Pryce et al., 1993). Under natural con-
ditions in the wild it takes dire circumstances for an experienced an-
thropoid mother to abandon her infant and they are even less likely to
do so than parous mothers among many nonprimate mammalian spe-
cies. Rare exceptions include a mother gorilla with twins who became
ill, or a langur mother who found herself subject to chronic stalking by
an infanticidal male (reviewed in Hrdy, 1999). Ape mothers in particu-
lar stand out for single-minded, highly protective possessiveness, re-
maining in continuous skin-to-skin contact with their baby for months
after birth. It is against this primate background that the extremely
variable responsiveness of human mothers seems unusual. For compa-
rable variability we must turn to other primates with shared care and
provisioning.

From shared care to full-fledged cooperative breeding

Since langurs are leaf-eaters, there would be little advantage to


allomaternal provisioning and they do not share food. Among primates
generally, shared care is fairly common, but provisioning the offspring
of other primates either via allomaternal suckling or actual delivery of
food to nearly weaned immatures is very unusual. Animal prey in
particular is highly prized among those apes that consume it, and it is
difficult to explain how voluntary delivery and offering of prey to others
ever became established in the hominin line (Hrdy, in press-b). Never-
theless, such provisioning must have been essential to permit the evolu-
tion of the larger brains and longer childhoods in the line of apes leading
to the genus Homo.
Today, the best documented cases of provisioning among still extant
primates are found among phylogenetically very distant New World
monkeys, monogamous species in the genus Aotus (Fernandez-Duque
Fig. 3. A middle-aged, multiparous allomother who has taken and infant but is tired of et al., 2009) and the many species in the subfamily Callitrichidae
carrying it, presses it to the ground. The infant was quickly retrieved, uninjured, by anoth-
er allomother (S.B. Hrdy/Anthro-photo).
where the majority of calories consumed by just-weaned young are
provided by fathers and other allomothers (Burkart and Finkenwirth,
2015; Rapaport, 2006) as is also the case among human hunter–gath-
erers. Once underway, with alloparental provisioning as well as obligate
matrilineal kin nearby (Hrdy, 2009). It takes two conditions for shared care embedded in the life history of a particular species, cooperative
care to emerge. The normal possessiveness and extreme protectiveness breeding plays out in various ways. In the case of cooperative breeders
of a new mother has to be relaxed, and she has to feel sufficiently characterized by rigid female dominance hierarchies as marmosets
trusting or tolerant of nearby group members to allow another individ- and tamarins are, millions of years of cooperative breeding have led to
ual access to her new infant. Obviously it would be interesting to know the evolution of some highly derived and quite surprising, physiological
more about neuroendocrinological status of postpartum primates adaptations. Dominant female have been selected so as to eliminate
with and without trusted associates nearby, but I know of no relevant infants born to subordinates thereby increasing the number of helpers
research. available for their own offspring.
As Drury et al. (2016-in this issue) emphasize, and the langur cases As predicted by Rosenblatt's work, there tends to be a correlation be-
discussed above illustrate, caution is called for when applying a cultur- tween urinary estradiol levels during pregnancy and post-partum
ally laden term like “abuse” to nonhuman animals. A lot depends on infant-directed behaviors (such as licking and carrying) among
who is doing what to whom, why, and especially on whether or not callitrichids (e.g. red-bellied tamarins, Saguinus laibatus, Pryce et al.,
the “abused” infant is harmed. One species' “negative early caregiving” 1993). Nevertheless, there are circumstances when pregnant callitrichids
is another creature's normal. Furthermore, in environments where it respond quite differently. Instead of extremely pregnant females finding
is as challenging to avoid being eaten as it is to stay fed, mothers babies especially attractive and seeking to nurture them, some marmoset
who reject their infants or leave a vocalizing infant unattended, would and tamarin females in the late stages of pregnancy actually find infants
be unlikely to remain mothers for long. Unless that rejecting mother aversive. Even experienced multiparous marmosets who previously suc-
learned from that experience to be more attentive with subsequent in- cessfully reared young exhibit longer latencies to approach during late
fants, heritable traits leading to maternal neglect would be selected pregnancy (Saltzman and Abbott, 2005).
against and disappear form the population's gene pool. When a mother Some dominant females in the later stages of pregnancy even attack
forcibly restrains her infant, she is less likely to be abusing it than and bite to death infants born to subordinates. This risk probably
protecting her infant from some perceived threat. Nor can infant explains why subordinate females avoid reproduction in the vicinity
complaints be taken as reliable indicators of maternal “abuse.” Infants of potentially infanticidal alphas—sometimes their own mother
protest when mothers deem it too dangerous for them to move about (Digby, 2001; Digby and Saltzman, 2009; Saltzman and Abbott, 2005,
276 S.B. Hrdy / Hormones and Behavior 77 (2016) 272–283

see also Clutton-Brock, in press for similar cases among other coopera- infants. Poachers are well aware that to capture a baby gorilla or orang-
tively breeding mammals). In spite of such risks, subordinate female utan to sell in local markets, they first must shoot the highly protective
common marmosets sometimes do risk giving birth, as for example mother. Otherwise she is not going to let go. Theirs is a reflexive protec-
after ovulating and soliciting matings from males belonging to another tiveness. Similarly, Old World monkey mothers carry and suckle infants
group during inter-troop encounters (Lazaro-Perea, 2001). Such no matter what, even those born with palsy or missing limbs (Berkson,
inconsistencies underscore just how situation-dependent and variable 1974; Turner et al., 2005). Both Old World monkey and ape mothers
reproductive and parenting behaviors among cooperative breeders are. continue to carry for days the desiccating corpses of infants that die
Across vertebrates, alloparental care combined with provisioning (Fig. 4). Exceptions to such reflexive protectiveness among wild primate
have most often evolved among highly social species already character- mothers almost all fall among species with obligate allomaternal care.
ized by altricial, fairly slow maturing, young under ecological conditions For example, among cooperatively breeding callitrichids, even experi-
which make it especially challenging to keep recently fledged or enced mothers who have already successfully reared infants may not
weaned offspring provided with palatable foods. Such conditions only ignore infants who cease to move, but sometimes abandon still vo-
include extreme fluctuations in rainfall or difficult-to-extract or process calizing infants.
food requiring adult provisioning (Hrdy, 2009; Jetz and Rubenstein, Infanticide is widespread among primates and a major source of
2011). As it happens, such conditions pertained in areas of early infant mortality. However, perpetrators virtually always attack the
Pleistocene Africa 1.8 million or so years ago, areas where early hominin offspring of others, not their own infants (Hausfater and Hrdy, 1984;
remains are found today (Magill et al., 2013; O'Connell et al., 2002; Parmigiani and vom Saal, 1994; van Schaik and Janson, 2001). Yes,
Potts, 1996). inexperienced first-time monkey and ape mothers may prove in-
Although relatively recent by the standards of cooperative breed- competent or neglectful, and may even abandon infants altogether,
ing's far longer evolutionary history among marmosets and tamarins, especially in captivity or if a primipara lacked opportunities for
two to three million years is sufficient for Darwinian selection to have prior caretaking experiences. Cases of mothers dragging or
begun to shape responses of all involved, mothers, fathers, alloparents neglecting infants such as those reported among free-ranging
and infants. Elsewhere I have explored cooperative breeding's implica- Japanese macaques (e.g., Hiraiwa, 1981; Schino and Troisi, 2005)
tions for the unusually variable expression of paternal commitment typically involve first-time young mothers or females who had lost
within and between human societies (Hrdy, 2008, 2009). I have also their own mother and so presumably had less exposure to younger
examined how the need to appeal to, and elicit succor from, allomothers sibs. In these respects, human primates are outliers. As in other pri-
as well as mothers would have affected infant development, generating mates, infanticide is widely documented across both traditional
novel ape phenotypes which over evolutionary time would have been and modern human societies. But unlike other primates, the likeliest
subjected to novel selection pressures favoring those immature apes perpetrator is the mother herself, usually in the hours or days right
best at monitoring the thoughts and feelings of others and ingratiating after birth. With the exception of cases summarized below, all in-
themselves with prospective caretakers, resulting in the emergence as volving species with obligate allomaternal care, I know of no other
early as several million years ago of more “emotionally modern” apes instances from nature where an experienced nonhuman primate
better able to read, and more interested in, the intentions of others mother inflicted injury on her own infant.
(Hrdy, 2009; in press-a). Here I focus on the extreme variability of re- As in all mammals, new human mothers are acutely sensitive
sponsiveness in pregnant females and new mothers, both those with to the sight, smells and vocalizations of newborns. Yet unlike their
and without prior maternal experiences. self-sacrificingly dedicated Great Ape relations, there are a range of
circumstances where postpartum mothers, including even experienced
Human mothers as outliers among hominoid primates mothers, seem peculiarly insensitive to infantile appeal (kindchenschema)
or infantile signals of distress, voluntarily abandoning or even killing their
Nonhuman ape mothers are all characterized by exclusive care of own infants (Drury et al., 2016-in this issue; Overpeck et al., 1998;
their new infants. They stand out for single-minded dedication to their Resnick, 1970).

Fig. 4. Old World monkey and non-human ape mothers are not known to discriminate based on infantile attributes and will continue to carry the corpse of a dead baby for days as this
young langur monkey mother is doing (S.B. Hrdy/AnthroPhoto).
S.B. Hrdy / Hormones and Behavior 77 (2016) 272–283 277

Maternal rejection of newborn humans is understandably a sensitive Apart from humans, the main (and so far only?) primate instances of
topic. Even anthropologists with relevant data can be reluctant to pub- maternal infanticide known to involve experienced mothers under
lish them. Nevertheless, numerous historical, sociological, institutional natural conditions occurred in species with obligate allomaternal care.
as well as ethnographic sources, testify to the widespread occurrence Every report known to me is summarized below. All occurred either
of postpartum abandonments or infanticide of newborns considered among monogamous species with obligate paternal care or else
premature or otherwise poorly timed (e.g., born too close to an older among monogamous or polyandrous cooperative breeders with
sib); born female when males are preferred, or vice versa, or born defec- alloparental as well as parental care and provisioning. Perhaps because
tive in various real or culturally arbitrary ways rendering them unlikely Callitrichidae have relied on allomaternal care and provisioning many
to prosper. Space here does not permit review of what is a vast, if largely millions of years longer than hominins have, the “dampening” processes
non-quantitative, often anecdotal or otherwise problematic literature involved are even more pronounced.
(see Hrdy, 1999: chapters 12–14; regarding scholarly reluctance to pub-
lish, see pp. 289–297).
Comparative evidence from long-time cooperative breeders

Curiously insensitive, and in the human case, even discriminating, As Bales and Saltzman (2016-in this issue) and Drury et al. (2016-in
primate mothers this issue) stress, it is important to select ethologically appropriate
models. But in the case of social neuroscientists interested in the under-
Whether in industrialized societies where infanticide is against the pinnings of maternal responses under the range of conditions which
law or in traditional societies where abandonment or infanticide are characterized human mothers during the course of evolution, this pre-
regarded as lamentable but sometimes necessary, this under-reported sents special challenges. Neither the often-studied rhesus macaque
and way under-studied phenomena cannot always, or even typically, nor any of our phylogenetically closest Great Ape relations who possess
be dismissed as the result of “deranged” or pathological behavior. human-like neural structures and nearly identical reproductive physiol-
Rather, the scale and patterning of maternal abandonment and infanti- ogies, raise infants the way our hominin ancestors did. But if we turn to
cide among humans (e.g., Badinter, 1981; Bonnet, 1993; Bugos and other primates with cooperative breeding, they are not only phyloge-
McCarthy, 1984; Schiefenhövel, 1989; Sheper-Hughes, 1992) raises netically very distant from humans, they have also been cooperative
the possibility of some dampening of maternal responses in postpartum breeders millions of years longer under conditions of within-group
mothers, a dampening that in some cases permits new mothers to breeding competition far more extreme than those among band-level
discriminate against offspring born at the “wrong” time or possessing hunter–gatherers.
the “wrong” attributes, in ways that have not been observed in other Humans last shared common ancestors with cooperatively breed-
anthropoid primates. Such limited ethological evidence as is available ing New World monkeys in the subfamily Callitrichidae more than
also suggests that mothers may sometimes find it aversive rather 35 million years ago. During much of that time, fast-breeding marmo-
than rewarding to hear cries from a very low birthweight, premature set and tamarin mothers were producing twins or triplets weighing up
or sickly-seeming infant or to look at a baby with facial disfigurement to one fifth of their body weight as often as twice a year, in the process
(Barden et al., 1989; Frodi et al., 1978; Mann, 1992). Such effects may becoming neurologically as well as physiologically adapted to
co-occur or be exacerbated in mothers experiencing severe postpartum allomaternal assistance. I hypothesize that their obligate reliance on
depression, stress, or anxiety (Agrati and Lonstein, 2016-in this issue; allomaternal support was implicated in the evolution in these species
Brummelte and Galea, 2015-in this issue; Lonstein et al., in press; of mechanisms for somehow bypassing ordinary neurophysiological se-
Haim et al., 2016-in this issue). quelae to the hormonal changes accompanying primate gestation.
In respect to relinquishing care or, under certain circumstances, When late-stage pregnant females encounter infants, instead of
ceasing altogether to care for babies, human mothers more nearly responding positively, they attack them. Of seven wild callitrichids ob-
resemble non-primate litter-bearing mammals than other higher pri- served biting to death another female's infant, five were pregnant
mates. Faced with severe scarcity or when disturbed, mammalian (Culot et al., 2011: Table 1). These pregnant females ignored visual
mothers are known to abandon dens or nests, or as in the case of cues or distress vocalizations that in other contexts signal “cuteness”
lions, whole litters. Mothers with multiple young may also discriminate or elicit care, sometimes targeting the vocalizing face.
between infants based on specific attributes such as a kangaroo mother Even stranger, in primates with obligate allomaternal care, new
closely pursued by a predator who jettisons an older joey while mothers may sometimes directly attack their own infants. Under pres-
retaining a tinier joey latched to her teat or a mother with a large litter sure to reduce metabolic costs from carrying and nursing babies, tama-
who shunts aside a runt (e.g., Hoogland, 1994, for prairie dogs; the rin and marmoset mothers in the genera Callithrix and Saguinus, as well
author's personal observation for domestic dogs). Litter-bearing rodents as mothers among monogamous, biparental care species in the genera
as well may differentially care for young in ways that differentially Callicebus and Aotus, pass babies off to others (usually males) after
affects their development (Jenkins et al., 2016-in this issue; Ragan they suckle them. But if babies continue to cling, they may attempt to
et al., 2016-in this issue). forcibly scrape babies off, or bite their hands or feet. In extreme cases,
Over the course of hominin evolution, more closely stacked mothers may bite off hands in their effort to force clinging babies to
dependents emerged in concert with allomaternal provisioning which let go (Cooper, 1964, cited in Hershkovitz, 1977:788; Hoffman et al.,
permitted shorter intervals between births of slow maturing young. 1995:402; Huck and Fernandez-Duque, 2013:378; Ross et al., 2007).
These “as-if” litters would have produced circumstances where hominin Unless fathers pick them up, complaining neonates may be left on the
mothers benefit from being able to calibrate investment in line with ground where they fell (e.g., Epple, 1970:66).
local conditions but also in line with attributes relevant to one infant's Interpreting surprisingly high frequencies of rejected infants among
viability relative to that of a current older, or as yet unborn, sibling. I hy- callitrichids is problematic. Because these tiny, arboreal, monkeys are
pothesize that more discriminative maternal responsiveness at emo- difficult to observe in the forests where they live, most reports of
tional levels not only co-evolved in humans with larger brains, abandonments derive from infants or their corpses, found on the floor
increased foresight and capacities for conscious decision-making and of breeding colonies (Bardi et al., 2001; Epple, 1970:66; Johnson et al.,
various cultural criteria—a generally accepted premise—but also co- 1991; Rothe, 1975:316–317) and even species of primates not known
evolved with a dampening of reflexive maternal responsiveness to reject infants in the wild are known to do so in captivity. But even
postpartum that facilitated the adjustment of a mother's level of com- taking this consideration into account, rates of infant rejection among
mitment in line with how much allomaternal support was available. callitrichids seem unusually high. Furthermore, only among cooperatively
278 S.B. Hrdy / Hormones and Behavior 77 (2016) 272–283

breeding primates have mothers in natural habitats ever been observed to began to bite the face of her vocalizing son, eventually dropping the
deliberately wound their own infants. corpse to the ground (Cäsar et al., 2008).
In tens of thousands of hours that chimpanzees, baboons, and other Apart from humans, only among these New World monkeys with
primates have been observed in nature, mothers are almost unfailingly extensive paternal and/or alloparental involvement do we find primate
protective. Yet in spite of far fewer observation hours, monogamous or mothers under natural conditions sufficiently inured to infantile signals
cooperatively breeding species with obligate reliance on allomaternal of distress as to deliberately injure one, even mutilate a still vocalizing
care are the only nonhuman primates where mothers living in natural face or preferentially consume the brains—a risky predilection in a
settings have been observed to deliberately mutilate, and in three mother who, like most mammals, will consume the placenta right
cases, almost certainly kill, their own infants. These mothers appear after giving birth. Under most conditions such a mother has good rea-
sufficiently inured to infantile signals of distress as to not only push sons to distinguish her neonate from, say, the afterbirth.
their baby away, but in their effort to unburden themselves, bite it Obviously these mothers are not conceiving and gestating babies in
to death. order to discard or recycle them. Abandonment and cannibalism are last
By now, infanticide (mostly by males but also by females) is reported resorts. Faced with insufficient assistance and hence a low likelihood of
for more than 50 species of primates (Palombit, 2012). But in these pro- successfully rearing her infant(s), a cooperatively breeding mother's
simian, monkey and ape cases individuals attack the offspring of others. best option would be to terminate investment, the sooner the better
The only three exceptions of maternal infanticide from nature involve (cf. the Bruce Effect where spontaneous abortions or reabsorption of
species with obligate allomaternal care. It's worth a closer look. fetuses occurs in pregnant mice, langurs or gelada baboons when in
To date there are ten reports of infanticide among wild-living non- the presence of a potentially infanticidal conspecific). Systematic analy-
human primates with either obligate paternal or alloparental care: ses of circumstances surrounding a large sample of captive callitrichid
nine involve cooperatively breeding callitrichids (Bezerra et al., 2007; rejections are consistent with this interpretation. Massimo Bardi et al.
Culot et al., 2011: Table 1; Digby and Saltzman, 2009; Tirado-Herrera (2001) used stepwise multiple regression to evaluate factors affecting
et al., 2001) and one involves monogamous biparental-care Callicebus survival to one year for 1093 Saguinus oedipus born between 1984 and
nigrifrons (Cäsar et al., 2008). In seven of these ten cases, a breeding 1993 at the New England Regional Primate Center. Fifty percent (546)
female killed another female's infant. But in the Callicebus and two were rejected and an additional 12% (134) were killed before they
tamarin cases, living babies were almost certainly bitten to death by could be removed and hand-reared. In 89% of these cases, termination
their own mothers. of investment ended within three days of birth. It is worth noting that
The existence of many firsthand observations of maternal rejections the postpartum timing of these rejections echoes the timing of infanti-
raises the possibility that some of the many dozens of mutilated, often cides among both traditional (Hrdy, 1999) and modern human socie-
headless or partially cannibalized, infants found on the floor of breeding ties. Of 3312 infanticides recorded by the U.S. Centers for Disease
colonies were also bitten by mothers (Bardi et al., 2001). Quite possibly, Control between 1989 and 1998, the incidence peaked in the first
a nonresponsive infant gets reclassified as “prey”, inviting consumption. week, with 82.6% of cases occurring on the first day. Thus, mothers
But the contrast with other primate mothers who remain protective were almost certainly present and implicated (Paulozzi, 2002; see also
of comatose or dead babies for days leads me to suspect some neuro- Bonnet, 1993; Overpeck et al., 1998).
physiological dampening of bonding mechanisms in postpartum Among callitrichids, mothers are the major caretakers immediately
callitrichids. This might help explain why some marmoset mothers fail after birth making it likely that mothers are primarily responsible for
to discriminate between their own infant and unfamiliar infants these rejections (Bardi et al., 2001:167). Bardi's further analysis
matched for age (Saltzman and Abbott, 2005). suggests that these mothers were responding to the availability of care-
takers or to pre- or postpartum signals of paternal willingness to help. In
Neural overrides to gestational and postpartum priming? Bardi's breeding colony sample the best predictor of maternal rejection
was shortage of allomaternal assistance. When older siblings were avail-
Clearly, the acute responsiveness to infantile signals of need and able 54.9% of infants were accepted, compared with 17.2% when absent.
distress that ordinarily emerges during mammalian pregnancy and Mothers in poor health and those bearing triplets were more likely to
which in some species emerges following previous exposure to infants abuse infants. Rates of rejection were also higher for inexperienced
(Akbari et al., submitted for publication; Drury et al., 2016-in this (64.1%) compared with experienced mothers (39.5%) and for younger
issue; Lonstein et al., in press; Numan and Insel, 2003) does not always parents of either sex than for older ones. Nevertheless, infanticide in
operate. Such responses seem particularly variable among primates the wild cases cannot be attributed to inexperienced individuals since
with obligate reliance on allomaternal care and provisioning. In field mothers who abused or rejected offspring were known to have success-
primatology's first report of maternal infanticide (Tirado-Herrera et al., fully reared young previously. Furthermore it is noteworthy that in
2001), a saddle-backed tamarin (Saguinus fusicollis) gave birth in a Bardi's sample, the tamarin mothers without prior caretaking experi-
group containing several males and a second (pregnant) female. On its ence were less, not more, likely than experienced mothers to mutilate
first day of life, the mother dropped her infant several times, only to babies they rejected. Compared with the 8.2% of inexperienced mothers
have it retrieved by one of the males, until when once again reunited who abused infants the rate for parous females was nearly twice that
with its mother, she “took a vertical position on a (tree) trunk, gripped (15.2%) suggesting that mutilations were due to something other than
the infant, and bit into its face. The infant moved its arms and legs. inexperience or incompetence.
Further bites were directed to the face and to the brain case.” The corpse
fell to the ground, revealing that the baby's brain had been entirely
Are callitrichid mothers polling available help?
consumed. In the second case, a mother mustached tamarin (Saguinus
mystax), who had successfully reared three sets of twins at times
Once established, dampening of maternal responsiveness provides
when four or five males were on hand to help, gave birth when there
novel evolutionary opportunities. In the case of callitrichids, it may
were only two or three males as well as a second breeding female in
have provided opportunities for mothers to elicit information about
the group. This mother as well was observed biting and eating
availability of help. Adult male callitrichids, especially progenitors,1
the head of her baby. Her son was vocalizing shortly before, but ob-
servers could not be certain he was still alive when the mother began 1
The usual uncertainty surrounding primate paternity is especially complicated among
biting (Culot et al., 2011:182). In a third case, a black-fronted titi mother callitrichids. Not only can twins and triplets be sired by more than one of the several males
(C. nigrifrons) gave birth to a healthy male carried mostly by her mate a polyandrous mother mates with, but the possibility of cells migrating between fetuses in
until the third day, when, on its being returned to his mother, she utero can result in chimeric individuals (Ross et al., 2007).
S.B. Hrdy / Hormones and Behavior 77 (2016) 272–283 279

respond to infant vocalizations as readily as mothers do (Sánchez et al., good allows them to bounce back after recurrent local population
2014), and sometimes more readily. When mothers push their babies crashes. Yet these tiny-brained, arboreal, squirrel-like monkeys not
(typically twins) off and they vocalize, it is males who usually retrieve only occupy different habitats but utterly different sensory worlds
and carry them, eventually returning them to their mother to nurse. than any ape does. Phylogenetically distant callitrichids also have very
Thus rejecting mothers can poll fathers, gaining information about different reproductive physiologies characterized by postpartum estrus,
their readiness to retrieve and carry infants. Some information about multiple births, a simplex uterus, partible paternity and chimeric prog-
probable male assistance may already have surfaced prior to birth. eny. Nor does anything in the ethnographic record for band-level
Tamarin fathers gain up to 8% of their body weight during a mate's human hunter–gatherers indicate callitrichid-like skews in female re-
pregnancy in anticipation of the extra effort they will expend carrying productive success (e.g., Tardif et al., 2013) or suggest that pregnant
babies (Ziegler et al., 2004). No doubt building up reserves in advance hunter–gatherers are motivated to harm infants born to group-mates.
makes energetic sense, but Toni Ziegler's findings suggest that fathers So far as humans go, the sort of competition between mothers that
may also be secondarily signaling to mothers their willingness to stay would lead to one mother eliminating the offspring of another has
and to help rear young. only been reported for the more stratified, propertied, human societies
The point is, some maternal rejections appear “strategic” rather than such as emerged in the past 20,000 years. Female reproductive compe-
terminal. Detailed observations of parental care among wild tamarins tition can be particularly pronounced in post-Neolithic societies where
(Leontopithecus rosalia) indicate that mothers adjust investment in inheritance of land or other resources are at issue. Such cases include
line with how much others are providing (Bales et al., 2002). This may exploitation of reproductively disenfranchised slaves or subordinates,
explain the curious observation that marmoset males, but not mothers, or wet-nurses coerced into caring for and even suckling the young of
find chimeric infants especially attractive (Ross et al., 2007, for Callithrix more dominant group members at the expense of their own infants'
kuhlii). Perhaps mothers can afford to “relax” if several males (including survival or reproductive opportunities (reviewed in Hrdy, 1999).
progenitors, partial progenitors, or former mates who are possibly However it is unclear how relevant to human neurophysiology such
progenitors) are going to pick up the slack. In any event, if the mother relatively recent customs might be.
repeats her rejections and no other allomother intervenes, the father Clearly, humans are physiologically and cognitively very, very differ-
too may terminate investment. ent from callitrichids, able to consciously assess current dilemmas and
Gauging her mate's probable responsiveness is valuable feedback for to make predictions about the future Yet what human mothers share
a mother to have sooner rather than later. Males with experience or in with these otherwise phylogenetically distant primate relations is an
good condition are especially responsive as reflected both by behavior evolutionary legacy of relying on extensive allomaternal care and provi-
and in higher prolactin levels (Storey and Ziegler, 2015-in this issue). sioning in order to rear young, and perhaps because of it (?) a situation-
Elsewhere, I have speculated that neuroendocrinological changes dur- dependent dampening of maternal responsiveness in the postpartum
ing maternally imposed exposure of males to infants, along with period. It is this respect that humans and callitrichids seem convergent.
prolonged intimate contact with infants, might also facilitate the emer-
gence of male provisioning in callitrichids and may perhaps have done Evolved predispositions versus conscious decisions? Or both?
so in prehuman (hominin) apes as well (Hrdy, in press-b) (Fig. 5).
Chimpanzees and other apes are reflexively possessive of their
A mother is not a mother is not a mother, and certainly not newborns, remaining in protective, skin-to-skin contact with them for
a marmoset months after birth until infants themselves begin to take the initiative
to move about. Women's postpartum responses are far more flexible,
Callitrichids, like early humans, make their living by both hunting for more contingent on external cues and varied than in other apes. Almost
(small) prey and gathering plant foods. In contrast to other Great Apes, everywhere new human mothers tolerate the proximity of familiar (and
but again, like callitrichids, humans are highly adaptable colonizing one assumes, trusted) conspecifics and voluntarily allow them to hold
primates whose potential for rapid reproduction when conditions are their newborns, something no other ape will do. Odder still, under a

Fig. 5. A golden lion tamarin (Leontopithecus rosalia) male passes twin infants back to their mother so they can nurse. Previous mates of the mother, as well as her older offspring, help to
care for infants and also capture prey to feed them around the time of weaning. The more male helpers infants have, the better are their survival prospects (drawing by S. Landry, a gift to
the author).
280 S.B. Hrdy / Hormones and Behavior 77 (2016) 272–283

range of conditions many mothers through history and prehistory have Counteracting the overrides even before we understand them
drastically reduced levels of care for their own offspring (reviewed in
Hrdy, 2009: Chapter 3), a level of apparent disinterest sometimes Patterns revealed by comparisons across species are useful for
dismissed as pathological or as evidence disproving the existence generating hypotheses and predictions. But the comparative method
of any biological basis for maternal “instincts” (Badinter, 1981; is rarely sufficient to identify, much less explain, underlying mecha-
Sheper-Hughes, 1992). It was the last quarter of the 20th century before nisms. It is my hope that the innovative experimental approaches
it was recognized that a mother's lifetime reproductive success can take reviewed in this special issue that have yielded so many insights into
priority over the well-being of particular offspring so that some of this the nature of maternal responses that enhance the survival of imma-
maternal retrenchment can be considered as adaptive (Hausfater and tures can also help us learn how maternal responses come to be so
Hrdy, 1984; Parmigiani and Vom Saal, 1994). much more conditional among cooperative breeders. Until then though,
As mammalian mothers go, women exhibit a dearth of “fixed action do evolutionary and comparative perspectives have anything useful
patterns” post-partum. Although visual and tactile inspections of their to offer?
newborns are routine, other behaviors are less predictable. Rarely do Although only sporadically practiced among post-industrial people,
new mothers lick their babies and (New Agers aside) women never breast-feeding and co-sleeping are primate universals. Around the
consume placentas as do all but a handful (camellids, many marine world, women in traditional societies still breast-feed and sleep with
mammals) of other mammals. There exists far more intra-specific vari- their babies. As amply shown by articles in this special issue, such inti-
ation in the breeding systems of nonhuman primates than previously mate contact with infants has neurophysiological effects correlated
assumed, yet even this variation pales by comparison with the extraor- with maternal responsiveness toward infants (e.g., Meliva-Seitz et al.,
dinarily variable and flexible parenting behaviors observed within and 2016-in this issue; Jonas and Woodside, 2016-in this issue and other ar-
between human societies. Furthermore, humans are virtually the only ticles in this issue) and is inversely correlated with neglect or abuse
apes where experienced mothers are known to voluntarily abandon or (e.g., Strathearn et al., 2009). Anecdotal evidence strongly suggests
harm their own offspring under natural conditions, that intimate contact with nursing infants postpartum can override a
Without doubt, the large human neocortex and extensive capacities mother's reason-based assessment of future prospects, including her
for culture play important roles. Why lick your baby when you can wipe premeditated decision to abandon her infant (Hrdy, 1999:315ff,
it clean, bathe it in water or with oil? As discussed, sapient capacities 454ff). Even absent parturition or breast-feeding, intimate contact
come into play when mothers discriminate between infants based on with infants has transformative effects on the neurophysiology of both
specific physical attributes or cultural ideals. Given that Homo sapiens male and female allomothers (Glocker et al., 2009; Kringelbach et al.,
parents can consciously calculate future prospects for a given infant, 2008; Lonstein et al., in press; Stolzenberg and Champagne, 2016-in
ascribe more or less value to that infant, and respond accordingly this issue, and especially Abraham et al., 2014; also summarized by
(Daly and Wilson, 1980, 1984; Scrimshaw, 1984), it is understandable Storey and Ziegler, 2015-in this issue).
that some social scientists insist that we first rule out the “consequences But holding her baby close and the establishment of breastfeeding
of ordinary practical reasoning” (Driscoll, 2005) or cultural constructs are not foregone post-partum events. For example, human mothers
(Badinter, 1981; Sheper-Hughes, 1992), as well as social pathology may experience delays, or even dampening in postpartum responsive-
(reviewed in Rees, 2009), before invoking neurobiological adaptations. ness. As in other cooperatively breeding primates, their responses
But as committed as most primate mothers – especially those with appear shaped by both social context and available support as well as
prior experience – are, human mothers are not the only primates by particular attributes of their infant, something not yet reported for
whose variable responsiveness in the first hours or days after birth any other primate. In no other primate are mothers known to discrimi-
includes rejection of infants and in extreme cases, even infanticide. nate on the basis of specific infant attributes (robustness, sex, birth
Parallels between humans and other cooperative breeders, including weight, culturally arbitrary traits such as too much or too little hair
tamarins or marmosets with a mere ten grams of gray matter and no or other perceived defects). No doubt, culture and rational assessments
capacities for perspective-taking (Burkart and Heschl, 2007), make concerning a child's future prospects are involved. But another possibil-
clear that higher cortical processes are not essential for mothers to ity is that the dampening of maternal responsiveness to infant
adjust commitment in line with available care. cues postpartum seen in cooperative breeders provided novel opportu-
Biology clearly has considerable influence on maternal and, in nities for selection to operate in this realm, but as to how and through
biparental species, paternal behaviors as well. Yet by themselves, which mechanisms, I am not qualified to even guess. I can however pro-
hormones rarely determine parental responsiveness (Bales and vide neuroscientists with several (hopefully testable) predictions.
Saltzman, 2016-in this issue; González-Mariscal et al., 2016-in this Situation-dependent neural overrides to maternal responsiveness
issue; Lonstein et al., in press; Lynn, 2016-in this issue; Numan and should be situated in more ancient portions of the brain than discrimi-
Insel, 2016-in this issue; Storey and Ziegler, 2015-in this issue; native solicitude based on specific infant attributes, and these discrimi-
Olazábal & Alsina-Llanes, 2016-in this issue). Parity, past experience, nations will not be entirely conscious.
sensitization and learning, along with local conditions, epigenetic The neurophysiological underpinnings for flexible parental commit-
and inter-generational effects sometimes going back several genera- ment are beginning to be explored in humans but remain completely
tions, modulate their impacts (reviewed in Barrett and Fleming, unstudied in other apes. More research is needed but finding ethical
2011; also see Beery et al., 2016-in this issue; Bridges, 2015-in this and humane ways of doing so will be a challenge. Based on recent dis-
issue; Lonstein et al., in press; Lomanowska and Melo, 2016-in this coveries about the neurophysiological bases of parental responsiveness
issue; Mileva-Seitz et al., 2016-in this issue; Stoltzenberg and Cham- (e.g., Abraham et al., 2014; Bell et al., 2014, 2015; Cong et al., 2014;
pagne, 2016-in this issue). When we consider large-brained H. sapi- Dulac et al., 2014; Kim et al., 2016-in this issue; Numan and Young,
ens with all our “later-evolving associative, and mentalizing brain 2016-in this issue; Pereira and Ferreira, 2016-in this issue; Wu et al.,
networks,” as described by Ruth Feldman (2016-in this issue), Kim 2014; Tsuneoka et al., 2015), Rosenblatt's successors in the field of social
et al. (2016-in this issue), and Pawluski et al. (2016-in this issue), neuroscience are well positioned to learn why post-partum responses
the underlying potentials so essential to our extraordinarily flexible in human mothers, even experienced ones, are so much more contin-
family systems, capacities for coordinated endeavors and culture- gent than in most other primates. In the meantime comparative and
generated adaptability to a range of habitats, things become more evolutionary perspectives and hypotheses about the evolutionary func-
complicated still. Yet I am convinced that even the most rational cal- tions of maternal distancing can at least help us understand why some
culations by parents are colored and complicated by humankind's interventions are proving more effective than others and thus worth
deep history of cooperative breeding. pursuing.
S.B. Hrdy / Hormones and Behavior 77 (2016) 272–283 281

In particular, the proposal that a deep hominin legacy of cooperative Bugos, P., McCarthy, L.M., 1984. Ayoreo infanticide: a case study. In: Hausfater, G., Hrdy, S.
(Eds.), Infanticide: Comparative and Evolutionary Perspectives. Aldine de Gruyter,
breeding produced new mothers peculiarly sensitive to their percep- Hawthorne, NY, pp. 503–520.
tions of available assistance can explain why social support sometimes Burkart, J., Finkenwirth, C., 2015. Marmosets as model species in neuroscience and evolu-
proves a better predictor of postpartum commitment than economic tionary anthropology. Neurosci. Res. 93, 8–19.
Burkart, J., Heschl, A., 2007. Perspective taking or behavior reading? Understanding
circumstances (e.g., Bonnet, 1993: 509; Hrdy, 1999: 371ff). It also visual access in common marmosets (Callithrix jacchus). Anim. Behav. 73,
helps to explain why aversive emotions toward infants postpartum 457–469.
tend to co-occur with feelings of “social isolation” (Bonnet, 1993: Burkart, J., Hrdy, S., Van Schaik, C., 2009. Cooperative breeding and cognitive evolution.
Evol. Anthropol. 18, 175–186.
504), why the presence of “doulas” during birth can matter so much Cäsar, C., Franco, E.S., Soares, G.C.N., Young, R.J., 2008. Observed cases of maternal infan-
(Meyer et al., 2001), why the responsiveness of an “at-risk” young ticide in a wild group of black-fronted titi monkeys (Callicebus negrifrons). Primates
mother is enhanced by her grandmother's presence in the same house 49, 143–149.
Cashdan, E. (Ed.), 1990. Risk and Uncertainty in Tribal and Peasant Economies. Westview,
(Spieker and Bensley, 1994), or why even such trivial seeming interven-
Boulder, CO.
tions as brief home visits to pre- and post-partum mothers by a support- Clutton-Brock, T., 2015. Mammalian Societies. J. Wiley and Sons, New York (in press).
ive social worker prove so beneficial (Eckenrode et al., 2000; Olds et al., Cockburn, A., 2006. Prevalence of different modes of parental care in birds. Proc. R. Soc.
1997). Lond. Ser. B 273, 1375–1383 (App. A.).
Crittenden, A.N., Marlowe, F.W., 2013. Cooperative child care among the Hadza: situating
multiple attachment in evolutionary context. In: Quinn, N., Mageo, J.M. (Eds.), Attach-
Acknowledgments ment Reconsidered: Cultural Perspectives on a Western Theory. Palgrave Macmillan,
New York, pp. 67–83.
Culot, L., Liedo-Ferrer, Y., Hoelscher, O., Muñoz Lazo, F.J.J., Huynen, M.-C., Heymannm,
Brief portions of this article were excerpted from an in press chapter E.W., 2011. Reproductive failure, possible maternal infanticide, and cannibalism in
entitled “Of marmosets, men, and the transformative power of babies,” wild moustached tamarins, Saguinus mystax. Primates 52, 179–186.
Daly, M., Wilson, M., 1980. Discriminative parental solicitude: a biological perspective.
in Costly and Cute: How Helpless Newborns Made Us Human, edited by J. Marriage Fam. 42, 177–188.
W. Trevathen and K. Rosenberg. As usual, heartfelt thanks to Sue Carter Daly, M., Wilson, M., 1984. A sociobiological analysis of human infanticide. In: Hausfater,
for mentoring an anthropologist voyaging in the unfamiliar realms of G., Hrdy, S. (Eds.), Infanticide: Comparative and Evolutionary Perspectives. Aldine de
Gruyter, Hawthorne, NY, pp. 487–502.
social neuroscience, to Alison Fleming for asking me to reflect on the ex- Darwin, C., 1874. The Descent of Man and Selection in Relation to Sex. John Murray,
traordinary wealth of new findings expertly reviewed in this special London.
issue, to callitrichidologists Karen Bales, Judith Burkart, Leslie Digby, De Silva, J.M., 2011. A shift towards birthing relatively large infants early in human evolu-
tion. Proc. Natl. Acad. Sci. 108, 1022–1027.
and Lisa Rapaport for criticisms of the earlier book chapter and to
De Silva, J.M., 2015. Brains, birth, bipedalism and the mosaic evolution of the helpless
Toni Ziegler for valuable insights. human infant. In: Trevathen, W., Rosenberg, K. (Eds.), Costly and Cute: How Helpless
Newborns Made Us Human. SAR Press, Santa Fe (in press).
Digby, L., 2001. Infanticide by female mammals: implications for the evolution of social
References systems. In: Van Schaik, C., Janson, C. (Eds.), Infanticide by Males and Its Implications.
Cambridge U Press, Cambridge, pp. 423–446.
Abraham, E., Handler, T., Shapira-Lichter, I., Kanat-Maymon, Y., Zagoory-Sharon, O., Digby, L., Saltzman, W., 2009. Balancing cooperation and competition in callitrichid
Feldman, R., 2014. Father's brain is sensitive to childcare experiences. Proc. Natl. primates: examining the relative risk of infanticide across species. In: Ford, S.M.,
Acad. Sci. 111 (27), 9792–9797. Porter, L.M., Davis, L.C. (Eds.), The Smallest Anthropoids: The Marmoset/Callimico
Agrati, D., Lonstein, J.S., 2015. Affective changes during the postpartum period: Influences Radiation. Springer, New York, pp. 135–153.
of genetic and experiential factors. Horm. Behav. 77, 141–152 (in this issue). Driscoll, C., 2005. Killing babies: Hrdy on the evolution of infanticide. Biol. Philos. 20,
Angelier, F., Wingfield, J.C., Tartu, S., Chastel, O., 2015. Does prolactin mediate parental 271–289.
and life-history decisions in response to environmental conditions in birds? A review. Drury, S.S., Sánchez, M.M., Gonzalez, A., 2015. When mothering goes awry: Challenges
Horm. Behav. 77, 18–29 (in this issue). and opportunities for utilizing evidence across rodent, nonhuman primate and
Akbari, E., Gonzalez, A., Dudin, A., Steiner, M., Fleming, A.S., 2015. Depressed mothers human studies to better define the biological consequences of negative early caregiv-
show increased anxiety and negativity and differential salivary cortisol in response ing. Horm. Behav. 77, 182–192 (in this issue).
to infant cries (submitted for publication). Dulac, C., O'Connell, L.A., Wu, Z., 2014. Neural control of maternal and paternal behaviors.
Badinter, E., 1981. Mother Love: Myth or Reality? Macmillan, New York Science 345 (6198), 765–770.
Bales, K., French, J, N., Dietz, J., 2002. Explaining variation in maternal care in cooperatively Eckenrode, J., Ganzel, B., Henderson, C.R., Smith, E., Olds, D., Powers, J., Cole, R., Kitzman,
breeding mammals. Anim. Behav. 63, 453–461. H., Sidora, K., 2000. Preventing child abuse and neglect with a program of nurse
Bales, K.L., Saltzman, W., 2015. Fathering in rodents: Neurobiological substrates and con- home visitation. J. Am. Med. Assoc. 284 (1), 1385–1397.
sequences for offspring. Horm. Behav. 77, 249–259 (in this issue). Emlen, S., 1995. An evolutionary theory of the family. Proc. Natl. Acad. Sci. 92,
Barden, R.C., Ford, M.E., Jensen, A.G., Rogers-Salyer, M., Salver, K.E., 1989. Effects of cranio- 8090–8099.
facial deformity in infancy on the quality of mother–infant interactions. Child Dev. 60, Epple, G., 1970. Maintenance, breeding and development of marmoset monkeys
819–824. (Callitrichidae) in captivity. Folia Primatol. 12, 56–76.
Bardi, M., Petto, A., Lee-Parritz, D., 2001. Parental failure in captive cotton-top tamarins Feldman, R., Gordon, I., Schneiderman, I., Weisman, O., Zagoory-Sharon, O., 2010. Nat-
(Saguinus oedipus). Am. J. Primatol. 54, 150–169. ural variations in maternal and paternal care are associated with systematic
Barrett, J., Fleming, A.S., 2011. All mothers are not created equal: neural and psychobio- changes in oxytocin following parent–infant contact. Psychoneuroendocrinology
logical perspectives on mothering and the importance of individual differences. 35, 1133–1141.
J. Child Psychol. Psychiatry 52 (4), 368–397. Feldman, R., 2015. The neurobiology of mammalian parenting and the biosocial context of
Beery, A.K., McEwen, L.M., MacIsaac, J.L., Francis, D.D., Kobor, M.S., 2015. Natural variation human caregiving. Horm. Behav. 77, 3–17 (in this issue).
in maternal care and cross-tissue patterns of oxytocin receptor gene methylation in Fernandez-Duque, E., Valeggia, C.R., Mendoza, S., 2009. The biology of paternal care in
rats. Horm. Behav. 77, 42–52 (in this issue). human and nonhuman primates. Ann. Rev. Anthropol. 38, 115–130.
Bell, A.F., Erickson, E.N., Carter, C.S., 2014. Beyond labor: the role of natural and synthetic Fleming, A.S., Lonstein, J.S., Lévy, F., 2015. Introduction to this Special Issue on parental be-
oxytocin in the transition to motherhood. J. Midwifery Women's Health 59 (1), havior in honor of Jay S. Rosenblatt. Horm. Behav. 77, 1–2 (in this issue).
35–42. Frodi, A.M., Lamb, M., Leavitt, L.A., Donovan, C.N., Neff, C., Sherry, D., 1978. Fathers' and
Bell, A.F., Carter, S.C., Colin, S.D., Golding, J., Davis, J.M., Lillard, T.S., Steffen, A.D., Rubin, L.H., mothers' responses to the faces and cries of normal and premature infants. Dev.
Lillard, T.S., Gregory, S.P., Harris, J.C., Connellly, J.J., 2015. Interaction between oxyto- Psychol. 14 (5), 40–49.
cin receptor DNA methylation and genotype is associated with risk of postpartum Glocker, M., Langleben, D.D., Ruparel, K., Longhead, J.W., Valdez, J.N., Griffin, M.D., Sachser,
depression in women without depression in pregnancy. Front. Genet. 6, 243. N., Gur, R.C., 2009. Baby schema modulates the brain reward system in nulliparous
Berkson, G., 1974. Social responses of animals to infants with defects. In: Lewis, M., women. Proc. Natl. Acad. Sci. 106 (22), 9115–9119.
Rosenblum, M.L. (Eds.), The Effect of the Infant on the Caregiver. John Wiley and González-Mariscal, G., Caba, M., Martínez-Gómez, M., Bautista, A., Hudson, R., 2015.
Sons, New York, pp. 233–249. Mothers and offspring: The rabbit as a model system in the study of mammalian ma-
Bezerra, B.M., Da Silva Souto, A., Schiel, N., 2007. Infanticide and cannibalism in a free- ternal behavior and sibling interactions. Horm. Behav. 77, 30–41 (in this issue).
ranging plurally breeding group of common marmosets (Callithrix jacchus). Am. Haim, A., Albin-Brooks, C., Sherer, M., Mills, E., Leuner, B., 2015. The effects of gestational
J. Primatol. 69, 945–952. stress and SSRI antidepressant treatment on structural plasticity in the postpartum
Bonnet, C., 1993. Adoption at birth: prevention against abandonment or neonaticide. brain - A translational model for postpartum depression. Horm. Behav. 77, 124–131
Child Abuse Negl. 17 (4), 501–513. (in this issue).
Bridges, R.S., 2016. Long-term alterations in neural and endocrine processes induced by Hausfater, G., Hrdy, S.B. (Eds.), 1984. Infanticide: Comparative and Evolutionary Perspec-
motherhood in mammals. Horm. Behav. 77, 193–203 (in this issue). tives. Aldine/de Gruyter, Hawthorne, NY.
Brummelte, S., Galea, L.A., 2015. Postpartum depression: Etiology, treatment and conse- Hawkes, K., O'Connell, J.F., Blurton Jones, N.G., 1989. Hardworking Hadza grand-
quences for maternal care. Horm. Behav. 77, 153–166 (in this issue). mothers. In: Standen, V., Foley, R.A. (Eds.), Comparative Socioecology: The
282 S.B. Hrdy / Hormones and Behavior 77 (2016) 272–283

Behavioral Ecology of Humans and Other Mammals. Basil Blackwell, London, Meehan, C., Hawks, S., 2013. Cooperative breeding and attachment among the
pp. 341–366. Aka foragers. In: Quinn, N., Mageo, J.M. (Eds.), Attachment Reconsidered: Cul-
Hawkes, K., O'Connell, J., Blurton Jones, N., Alvarez, H., Charnov, E., 1998. Grandmothering, tural Perspectives on a Western Theory. Palgrave Macmillan, New York,
menopause and the evolution of human life histories. PNAS 95, 1336–1339. pp. 85–113.
Henry, P.I., 2000. Cooperative reproduction in Ituri hunter–gatherers: who cares for Efe Meyer, B.A., Arnold, J.A., Pascali-Bonaro, D., 2001. Social support by doulas during labor
infants? Curr. Anthropol. 41, 836–866. and the early postpartum period. Hospital Physician 37 (9), 57–65http://www.
Hershkovitz, P., 1977. Living New World Monkeys (Platyrrhini), with an Introduction to turner-white.com/pdf/hp_sep01_doulas.pdf.
Primates. U of Chicago Press, Chicago. Mileva-Seitz, V.R., Bakermans-Kranenburg, M.J., van IJzendoorn, M.H., 2015. Genetic
Hewlett, B., Winn, S., 2014. Allomaternal nursing in humans. Curr. Anthropol. 55 (2), mechanisms of parenting. Horm. Behav. 77, 211–223 (in this issue).
1–30. Numan, M., Insel, T.R., 2003. The Neurobiology of Parental Behavior. Springer-Verlag, New
Hiraiwa, M., 1981. Maternal and allomaternal care in a troop of free-ranging Japanese York.
macaques. Primates 22, 309–329. Numan, M., Young, L.J., 2015. Neural mechanisms of mother-infant bonding and pair
Hoffman, K., Mendoza, S., Hennessey, M., Mason, W., 1995. Responses of infant titi bonding: Similarities, differences, and broader implications. Horm. Behav. 77,
monkeys Callicebus molloch to removal of one or both parents: evidence of paternal 98–112 (in this issue).
attachment. Dev. Psychobiol. 28, 399–407. O'Connell, J.F., Hawkes, K., Blurton Jones, N.G., 1999. Grandmothering and the evolution of
Hoogland, J., 1994. Nepotism and infanticide among prairie dogs. In: Parmigiani, S., Vom Homo erectus. J. Hum. Evol. 36, 461–485.
Saal, F. (Eds.), Infanticide and Parental Care. Harwood Academic, Langhorne, PA, O'Connell, J.F., Hawkes, K., Lupo, K.D., Blurton Jones, N.G., 2002. Male strategies and Plio-
pp. 321–337. Pleistocene archeology. J. Hum. Evol. 43 (6), 831–872.
Hrdy, S., 1977. The Langurs of Abu: Female and Male Strategies of Reproduction. Harvard Olds, D.L., Eckenrode, J., Henders, C.R., Kitzman, H., Powers, J., Cole, R., Sidora, K.,
U Press, Cambridge. Morris, P., Petitt, L.M., Luckey, D., 1997. Long-term effects of home visitation on
Hrdy, S., 1999. Mother Nature: A History of Mothers, Infants and Natural Selection. maternal life course and child abuse and neglect. J. Am. Med. Assoc. 278 (8),
Pantheon, New York. 637–643.
Hrdy, S., 2008. Cooperative breeding and the paradox of facultative fathering. In: Bridges, Olazábal, D.E., Alsina-Llanes, M., 2015. Are age and sex differences in brain oxytocin re-
R. (Ed.), The Neurobiology of the Parental Mind. Academic Press, New York, ceptors related to maternal and infanticidal behavior in naïve mice? Horm. Behav.
pp. 405–414. 77, 132–140 (in this issue).
Hrdy, S., 2009. Mothers and Others: The Evolutionary Origins of Mutual Understanding. Overpeck, M.D., Brenner, R.A., Trumble, A.C., Trifiletti, L.B., Berendes, H.W., 1998. Risk
Harvard U Press, Cambridge. factors for infant homicide in the United States. N. Engl. J. Med. 339, 1211–1216.
Hrdy, S., 2016a. Development plus social selection in the evolution of emotionally Palombit, R.A., 2012. Infanticide: male strategies and female counterstrategies. In: Mitani,
modern humans. In: Meehan, C.L., Crittenden, A. (Eds.), Childhood: Origins, J.C., Call, J., Kappeler, P., Palombit, R.A., Silk, J. (Eds.), The Evolution of Primate Socie-
Evolution, and Implications. University of New Mexico Press, Albuquerque (in ties. U of Chicago Press, Chicago, pp. 433–468.
press-a). Parmigiani, S., Vom Saal, F. (Eds.), 1994. Infanticide and Parental Care. Harwood Academ-
Hrdy, S., 2016b. Of marmosets, men, and the transformative power of babies. In: ic, Langhorne, PA.
Trevathen, W., Rosenberg, K. (Eds.), Costly and Cute: How Helpless Newborns Paulozzi, L., 2002. Variation in homicide risk during infancy—United States, 1989–1998.
Made Us Human. SAR Press, Santa Fe (in press-b). Morb. Mortal. Wkly. Rep. 51(9). Centers for Disease Control and Prevention,
Huck, M., Fernandez-Duque, E., 2013. When dads help: male behavioral care during pp. 187–189.
primate infant development. In: Clancy, K.B.H., Hinde, K., Rutherford, J.N. (Eds.), Pawluski, J.L., Lambert, K.G., Kinsley, C.H., 2015. Neuroplasticity in the maternal hippo-
Building Babies: Primate Development in Proximate and Ultimate Perspective. campus: Relation to cognition and effects of repeated stress. Horm. Behav. 77,
Springer, New York, pp. 361–385. 86–97 (in this issue).
Isler, K., van Schaik, C., 2012. Allomaternal care, life history and brain size evolution in Pereira, M., Ferreira, A., 2015. Neuroanatomical and neurochemical basis of parenting: Dy-
mammals. J. Hum. Evol. 63 (1), 52–63. namic coordination of motivational, affective and cognitive processes. Horm. Behav.
Jenkins, J.M., McGowan, P., Knafo-Noam, A., 2015. Parent-offspring transaction: Mecha- 77, 72–85 (in this issue).
nisms and the value of within family designs. Horm. Behav. 77, 53–61 (in this issue). Potts, R., 1996. Humanity's Descent: The Consequences of Ecological Instability. Avon
Jetz, W., Rubenstein, D., 2011. Environmental uncertainty and the global biogeography of Books, New York.
cooperative breeding in birds. Curr. Biol. 21 (5), 72–78. Pryce, C.R., Dobell, M., Martin, R., 1993. Effects of sex steroids on maternal motivation
Johnson, L.D., Petto, A.J., Schgal, P.K., 1991. Factors in the rejection and survival of captive in the common marmoset (Callithrix jacchus): development and application of
cottontop tamarins (Saguinus oedipus). Am. J. Primatol. 25, 91–102. an operant system with maternal reinforcement. J. Comput. Psych. 107, 99–115
Jonas, W., Woodside, B., 2015. Physiological mechanisms, behavioral and psychological (xxx).
factors influencing the transfer of milk from mothers to their young. Horm. Behav. Ragan, C.M., Harding, K.M., Lonstein, J.S., 2015. Associations among within-litter differ-
77, 167–181 (in this issue). ences in early mothering received and later emotional behaviors, mothering, and cor-
Kaplan, H., 1994. Evolutionary and wealth flows theories of fertility: empirical tests and tical tryptophan hydroxylase-2 expression in female laboratory rats. Horm. Behav. 77,
new models. Popul. Dev. Rev. 20 (4), 753–791. 62–71 (in this issue).
Kaplan, H., Hill, K., Lancaster, J., Hurtado, A.M., 2000. A theory of human life history evo- Rapaport, L., 2006. Provisioning in wild golden lion tamarins (Leontopithecus rosalia).
lution: diet, intelligence and longevity. Evol. Anthropol. 9, 156–185. Behav. Ecol. 22 (4), 745–754.
Kim, P., Strathearn, L., Swain, J.E., 2015. The maternal brain and its plasticity in humans. Rees, A., 2009. The Infanticide Controversy: Primatology and the Art of Field Science. U of
Horm. Behav. 77, 113–123 (in this issue). Chicago Press, Chicago.
Koenig, W., Dickinson, J. (Eds.), 2004. Ecology and Evolution of Cooperative Breeding in Resnick, P.J., 1970. Murder of the newborn: a psychiatric review of filicide. Am. J. of
Birds. University of Cambridge Press, Cambridge. Psychiatry 126, 325–334.
Konner, M., 2010. The Evolution of Childhood: Relationships, Emotion and Mind. Harvard Ross, C.N., French, J.A., Orti, G., 2007. Germ-line chimerism and parental care in marmo-
U Press, Cambridge. sets (Callithrix jacchus). Proc. Natl. Acad. Sci. 104, 6278–6282.
Kringelbach, M., Lehtonen, A., Squire, S., Harvey, A.G., Craske, M., Holliday, A., Green, T., Rothe, H., 1975. Influence of newborn marmosets' (Callithrix jacchus) behavior on expres-
Azia, P., Hansen, P., Corneliessen, P., Stein, A., 2008. A specific and rapid neural signa- sion and efficiency of maternal and paternal care. Fifth International Congress of
ture for parental instinct. PLoS 3 (2), e1664. Primatology, Nagoya 1974, pp. 315–320.
Kuzawa, C.W., Chugani, H.T., Grossman, L.I., Lipovich, L., Muzik, O., Hof, P.R., Wildman, Saltzman, W., Abbott, D.H., 2005. Diminished maternal responsiveness during pregnancy
D.E., Sherwood, C.C., Leonard, W.R., Lange, N., 2014. Metabolic costs and evolutionary in multiparous female common marmosets. Horm. Behav. 47, 151–163.
implications of human brain development. Proc. Natl. Acad. Sci. 111 (36), Sánchez, S., Ziegler, T.E., Snowdon, C.T., 2014. Both parents respond equally to infant cues
13010–13015. in the cooperatively breeding common marmoset, Callithrix jacchus. Anim. Behav. 97,
Lazaro-Perea, C., 2001. Intergroup interactions in wild common marmosets, Callithrix 95–103.
jacchus: territorial defense and assessment of neighbours. Anim. Behav. 62, 11–21. Schiefenhövel, W., 1989. Reproduction and sex ratio manipulation through preferential
Lomanowska, A.M., Melo, A.I., 2015. Deconstructing the function of maternal stimulation female infanticide among the Eipos. In: Rasa, A., Vogel, C., Voland, E. (Eds.), The
in offspring development: Insights from the artificial rearing model in rats. Horm. Highlands of Western New Guinea, in: The Sociobiology of Sexual and Reproductive
Behav. 77, 224–236 (in this issue). Strategies. Chapman and Hall, London, pp. 170–193.
Lonstein, J.S., Lévy, F., Fleming, A.S., 2015. Common and divergent psychobiological mech- Schino, G., Troisi, A., 2005. Neonatal abandonment in Japanese macaques. Am. J. Phys.
anisms underlying maternal behaviors in non-human and human mammals. Horm. Anthopol. 126, 447–452.
Behav. 73, 156–185. Scrimshaw, S., 1984. Infanticide in human populations: societal and individual concerns.
Lovejoy, C.O., 1981. The origin of man. Science 211, 341–350. In: Hausfater, G., Hrdy, S. (Eds.), Infanticide: Comparative and Evolutionary Perspec-
Lynn, S.E., 2015. Endocrine and neuroendocrine regulation of fathering behavior in birds. tives. Hawthorne, NY, Aldine de Gruyter, pp. 439–462.
Horm. Behav. 77, 237–248 (in this issue). Sear, R., Mace, R., 2008. Who keeps children alive? A review of the effects of kin on child
Magill, C., Ashley, G.M., Freeman, C.H., 2013. Ecosystem variability and early human survival. Evol. Hum. Behav. 29, 1–18.
habitats in eastern Africa. Proc. Natl. Acad. Sci. 110, 1167–1174. Sheper-Hughes, N., 1992. Death Without Weeping: The Violence of Everyday Life in
Mann, J., 1992. Nurturance or negligence: maternal psychology and behavioral preference Brazil. U of California Press, Berkeley.
among preterm twins. In: Barkow, J.H., Cosmides, L., Tooby, J. (Eds.), The Adapted Spieker, S.J., Bensley, L., 1994. Roles of living arrangements and grandmother social
Mind. Oxford U Press, Oxford, pp. 367–390. support in adolescent mothering and infant attachment. Dev. Psychol. 30 (1),
Marlowe, F., 2010. The Hadza: Hunter–Gatherers of Tanzania. U of California Press, 102–111.
Berkeley. Stolzenberg, D.S., Champagne, F.A., 2015. Hormonal and non-hormonal bases of maternal
Meehan, C., Crittenden, A. (Eds.), 2016. Childhood: Origins, Evolution and Implications. behavior: The role of experience and epigenetic mechanisms. Horm. Behav. 77,
University of New Mexico Press, Albuquerque (in press). 204–210 (in this issue).
S.B. Hrdy / Hormones and Behavior 77 (2016) 272–283 283

Storey, A.E., Ziegler, T.E., 2015. Primate paternal care: Interactions between biology and Turner, S., Gould, L., Duffus, D.A., 2005. Maternal behavior and infant congenital limb
social experience. Horm. Behav. 77, 260–271 (in this issue). malformation in a free-ranging group of Macaca fuscata on Awaji Island. Jpn. Int.
Strathearn, L., Mamun, A.A., Najman, J.M., O'Callaghan, M.J., 2009. Does breastfeeding pro- J. Primatol. 26, 1435–1457.
tect against substantiated child abuse and neglect? A 15-year cohort study. Pediatrics Van IJzendoorn, M., Sagi, A., Lambermon, M., 1992. The multiple caretaker paradox: data
123 (2), 483–493. from Holland and Israel. In: Pianta, R.C. (Ed.), Beyond the Parent: The Role of the
Tardif, S., Ross, C.N., Smucny, D., 2013. Building marmoset babies: trade-offs and cutting Other Adults in Children's LivesNew Directions for Child and Adolescent Develop-
bait. Dev. Primatol.: Progress and Prospects 37, 169–183. ment 57. Jossey-Bass, San Francisco, pp. 5–24.
Tirado-Herrera, E., Knogge, C., Heymann, EW., 2001. Infanticide in a group of wild saddle- Van Schaik, C., Janson, C. (Eds.), 2001. Infanticide by Males and Its Implications.
backed tamarins, Saguinus fusicollis. Am J Primatol 50, 153–157. Cambridge U Press, Cambridge.
Trevathen, W., Rosenberg, K. (Eds.), 2015. Costly and Cute: How Helpless Newborns Made Wu, Z., Autrey, A.E., Bergan, J.F., Watabe-Uchida, M., Dulac, C., 2014. Galanin neurons in
Us Human. SAR Press, Santa Fe (in press). the medial preoptic area govern parental behaviour. Nature 509, 325–330.
Tsuneoka, Y., Tokita, K., Yoshihara, C., Amano, T., Esposito, G., Huang, A.J., Yu, L., Odaka, Y., Ziegler, R., Washabaugh, K.F., Snowdon, C.T., 2004. Responsiveness of expectant cotton-
Shinozuka, K., McHugh, T.J., Kuroda, K.O., 2015. Distinct preoptic-BST nuclei dissociate top tamarins, Saguinus oedipus, to mate's pregnancy. Horm. Behav. 45, 84–92.
paternal and infanticidal behavior in mice. EMBO J. 34 (19), 2385–2481.

You might also like