Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Brain Research Bulletin 93 (2013) 97–109

Contents lists available at SciVerse ScienceDirect

Brain Research Bulletin


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

Review

The neurobiology of abnormal manifestations of aggression—A review of


hypothalamic mechanisms in cats, rodents, and humans
Jozsef Haller ∗
Department of Behavioral Neurobiology, Hungarian Academy of Sciences, Institute of Experimental Medicine, H-1083 Budapest, Szigony utca 43, Hungary

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

Article history: Aggression research was for long dominated by the assumption that aggression-related psychopatholo-
Received 24 August 2012 gies result from the excessive activation of aggression-promoting brain mechanisms. This assumption
Received in revised form 8 October 2012 was recently challenged by findings with models of aggression that mimic etiological factors of
Accepted 9 October 2012
aggression-related psychopathologies. Subjects submitted to such procedures show abnormal attack
Available online 17 October 2012
features (mismatch between provocation and response, disregard of species-specific rules, and insen-
sitivity toward the social signals of opponents). We review here 12 such laboratory models and the
Keywords:
available human findings on the neural background of abnormal aggression. We focus on the hypo-
Aggression
Hypothalamus
thalamus, a region tightly involved in the execution of attacks. Data show that the hypothalamic
Serotonin mechanisms controlling attacks (general activation levels, local serotonin, vasopressin, substance P, gluta-
Vasopressin mate, GABA, and dopamine neurotransmission) undergo etiological factor-dependent changes. Findings
Substance P suggest that the emotional component of attacks differentiates two basic types of hypothalamic mech-
Dopamine anisms. Aggression associated with increased arousal (emotional/reactive aggression) is paralleled by
increased mediobasal hypothalamic activation, increased hypothalamic vasopressinergic, but dimin-
ished hypothalamic serotonergic neurotransmission. In aggression models associated with low arousal
(unemotional/proactive aggression), the lateral but not the mediobasal hypothalamus is over-activated.
In addition, the anti-aggressive effect of serotonergic neurotransmission is lost and paradoxical changes
were noticed in vasopressinergic neurotransmission. We conclude that there is no single ‘neurobiologi-
cal road’ to abnormal aggression: the neural background shows qualitative, etiological factor-dependent
differences. Findings obtained with different models should be viewed as alternative mechanisms rather
than conflicting data. The relevance of these findings for understanding and treating of aggression-related
psychopathologies is discussed.
This article is part of a Special Issue entitled ‘Extrasynaptic ionotropic receptors’.
© 2012 Elsevier Inc. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
2. Hypothalamic attack areas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
3. The neurochemistry of attack-related hypothalamic mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
4. Hypothalamic mechanisms of abnormal aggression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
4.1. Overall activation of the hypothalamus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
4.2. Serotonin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
4.3. Vasopressin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
4.4. Other neurotransmitters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
4.5. The human case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
5. Etiological factor-dependent neuronal plasticity and abnormal aggression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
6. Summary and conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
Conflict of interest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
Acknowledgement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107

∗ Tel.: +36 1 210 9400; fax: +36 1 210 9423.


E-mail addresses: haller.jozsef@koki.mta.hu, haller@koki.hu

0361-9230/$ – see front matter © 2012 Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.brainresbull.2012.10.003
98 J. Haller / Brain Research Bulletin 93 (2013) 97–109

1. Introduction control is evolutionarily conserved (e.g. cats: Romaniuk, 1965;


chicken: Putkonen, 1966; cichlid fish: Demski, 1973; guinea pig:
Animal and human research on aggression followed different Martin, 1976; humans: Bejjani et al., 2002; lizards: Sugerman and
routes for a long time. Beyond an interest in normal aggressive- Demski, 1978; monkeys: Lipp, 1978; mice: Lin et al., 2011; opossum:
ness, human research focused mainly on abnormal manifestations Roberts et al., 1967; rats: Vergnes and Karli, 1969; Kruk et al., 1979).
of aggression that are conceptualized as psychopathologies. Animal In lizards at one extreme, aggression can be induced from several
research focused on natural forms of aggressiveness or electri- brain areas, while at the other extreme the hypothalamus is the only
cal stimulation-induced aggression. The underlying thought was area from where attacks can reliably be elicited by electric stim-
that the focal points of aggression control will provide handles ulation in rats (Kruk, 1991). Hypothalamic structures subserving
for understanding abnormal human aggression and for developing attack were best described in cats and rats and substantial evidence
novel treatment strategies (Siegel et al., 2007). Here we challenge is available in mice, hamsters and humans; therefore, we will focus
this view by showing that the hypothalamic underpinnings of on these species.
aggression are different in different models that were developed There are two hypothalamic regions of interest as it regards the
to mimic aggression-related psychopathologies. Such differences neural control of attacks: a region situated ventrally and medially
are not restricted to the hypothalamus. E.g. the mechanisms that to the fornix which induces affective aggression (“defensive rage”
control of aggression show qualitative differences in two laboratory in the cat, conspecific biting in rats and mice, and outbursts of
models of abnormal aggression. In the hypoarousal-related aggres- aggression in humans), and an area lateral to the fornix from where
sion model, the central amygdala–lateral hypothalamus–ventral emotionally silent predatory attacks can be induced (attacks on rats
periaqueductal gray pathway is activated, rendering the neural in cats and mouse killing in rats) (Fig. 1). A “predatory” area was
control of this type of abnormal aggression similar to that of preda- not demonstrated in mice, hamsters and humans.
tory attacks (Tulogdi et al., 2010). In contrast, increased activation of In cats, two qualitatively different behaviors can be induced by
the medial amygdala–mediobasal hypothalamus–dorsal periaque- the electrical stimulation of the hypothalamus, namely defensive
ductal gray pathway was noticed in a model of hyperarousal-driven rage (affective aggression: threatening vocalizations and postures,
aggression (Toth et al., 2012). These findings show that behav- and strikes upon provocation, i.e. behaviors similar to those shown
iorally and emotionally different forms of abnormal aggression are during territorial aggression and defense) and quiet biting (preda-
controlled by qualitatively different neural mechanisms. tory aggression: killing the pray (e.g. a rat) by biting the back of
The number of models developed to understand abnormal its neck without vocalizations or threat postures) (Siegel et al.,
aggression is increasing, but the mechanisms operating in these 1999). Hypothalamic structures controlling affective and predatory
models were not compared so far. As a first attempt, we compare aggression are topographically and functionally separated: the rel-
here hypothalamic mechanisms in 12 models of abnormal aggres- evant hypothalamic structures (ventral/medial and lateral to the
sion. The hypothalamus was chosen for its tight involvement in fornix, respectively) are not only behaviorally specific, but the two
aggression control (see below). Models were selected for review areas also inhibit each other via GABAergic mechanisms (Cheu and
based on two criteria: hypothalamic mechanisms were studied, Siegel, 1998). The topographic organization of affective and preda-
and subjects convincingly showed abnormal forms of aggression. tory aggression is also revealed by the distinct neural connections
The following criteria were used to identify models of abnormal and the pharmacology of hypothalamic mechanisms (Brutus and
aggression: (i) the models were based on treatments that resem- Siegel, 1989; Shaikh et al., 1984; Siegel et al., 1999, 2007). Thus,
ble etiological factors of aggression-related psychopathologies, and affective and predatory aggressions are controlled by distinct neu-
(ii) aggressive attacks show one or more of the following features: ral circuits in the cat.
(a) mismatch between provocation and response (excessively high The study of hypothalamic structures involved in attack fol-
attack counts, attacks in inappropriate situations), (b) disregard of lowed a different trajectory in the rat. The precise location and
species-specific rules (decreased signaling of attacks, attacks on pharmacological responsiveness of the mediobasal hypothalamus
females or on vulnerable targets), and (c) insensitivity toward the (“hypothalamic attack area”) that controls intraspecific attacks was
social signals of the opponent (sustaining attacks despite submis- described in detail, while brain areas involved in predatory aggres-
siveness on the other side) (Haller and Kruk, 2006). Based on recent sion received less attention. Similar to cats, attacks on conspecifics
findings, this triad of abnormal features was completed by a fourth, can be induced in rats by the stimulation of the mediobasal hypo-
namely “offensive ambiguity” that was seen in two models. Ham- thalamus (“hypothalamic attack area”), which is located ventral
sters repeatedly defeated (“subjugated”) during adolescence were and medial to the fornix, and includes the posterior part of the
highly aggressive against smaller opponents but showed decreased anterior hypothalamus, the relatively cell-sparse region extending
aggressiveness when confronted with opponents of similar size from the fornix to the 3rd ventricle, and the ventrolateral aspects
(Delville et al., 1998). Offensive ambiguity was also seen in rats sub- of the ventromedial hypothalamus (Kruk et al., 1979; Kruk, 1991;
mitted to post-weaning social isolation, which showed increased Lammers et al., 1988). Later studies demonstrated that sponta-
attack counts but were defensive at the same time (Toth et al., neous aggression induces c-Fos activation in the very same region
2011). The main features of the 12 models reviewed here were sum- (Halasz et al., 2002a); moreover, the unilateral electric stimula-
marized in Table 1. Human findings on hypothalamic aggression tion of the mediobasal hypothalamus induces attacks only when
control were also reviewed; the existence of an aggression-related the contra-lateral hypothalamic side also shows c-Fos activation in
psychopathology was used as a sign of abnormal aggressiveness in a manner, and to the extent, similar to that seen in spontaneous
this case. aggression (Halasz et al., 2002b). Although the location of the con-
specific attack site appears similar in rats and cats, the behavioral
response is somewhat different: threats are missing from the rat
2. Hypothalamic attack areas “mediobasal hypothalamic repertoire” (such threats are typical to
hypothalamically-evoked affective aggression in cats), and signs of
The first evidence on the involvement of the hypothalamus defensiveness that are clear in the cat are not obviously present
in aggression control came from stimulation studies performed in the rat. In contrast, predatory attacks are very similar in the
by Hess (1928). Subsequent research showed that biting attacks two species and consist of a rapid bite aimed at the back of the
can be elicited by hypothalamic stimulation in virtually all the neck of the pray (rats for cats and mice for rats). Although it was
species studied so far, suggesting that this mechanism of aggression demonstrated that the electrical and neurochemical stimulation of
J. Haller / Brain Research Bulletin 93 (2013) 97–109 99

Fig. 1. Aggression-related hypothalamic areas in the cat, rat and mouse. (A) Distribution of sites from which defensive rage (red) and quiet biting (blue) are elicited by
electrical stimulation in the cat brain (based on Siegel et al., 1999) (hypothalamic section from the cat; section 160, frame 11.5 from Jasper and Marsan, 1960). (B) Sites
from which conspecific biting attacks (red) and mouse killing (blue) are elicited by electrical stimulation in the rat brain (based on Lammers et al., 1988; Woodworth,
1971) (hypothalamic section from the rat brain at −2.30 mm from Bregma; Paxinos and Watson, 1998). (C) Sites from which conspecific biting attacks can be elicited in
mice by optogenetic stimulation (red) (based on Lin et al., 2011) (hypothalamic section from the mouse brain at Bregma −1.46; Paxinos and Franklin, 2001). 3rd V, 3rd
ventricle; AHp, anterior hypothalamic nucleus, posterior part; ARC, arcuate nucleus; DH, dorsal hypothalamic area; DMH, dorsomedial hypothalamic nucleus; F, fornix; Fil:
nucleus filiformis; IC: internal capsule; LHA, lateral hypothalamic area; OT, optic tract; PVN, paraventricular hypothalamic nucleus; RE, nucleus reuniens; VMH, ventromedial
hypothalamic nucleus; TCA, tuber cinereum area; ZI, zona incerta.
100 J. Haller / Brain Research Bulletin 93 (2013) 97–109

Table 1
The main characteristics of the abnormal aggression models reviewed.

Condition Abnormal features of Species Reference Human etiological factor Reference


attack modeled

Lines selected for Counts increased Mouse Benus et al. (1991) Inherited Miles and Carey
aggression >10-folds; vulnerable and Haller et al. aggression-related traits (1997)
targets; (2006)
females/submissive
opponents attacked
Acute alcohol Counts increased over Mouse Miczek et al. (1992) Alcohol-induced Bennett et al.
mean + 2 SDs aggression (1969)
Anabolic steroid Counts increased 3–5-fold Hamster Melloni and Ferris Hostility and irritability Su et al. (1993)
treatmentc (1996) due to anabolic steroids
Repeated defeat during Counts increased 5 fold Hamster Delville et al. Antisocial behavior Dodge et al. (1990)
adolescence against small intruders; (1998) following child abuse
counts halved against large
intruders
Glucocorticoid Vulnerable targets Rat Haller et al. (2001) Hypoarousal associated Virkkunen (1985)
deficiency attacked; deficient attack antisocial aggression and Woodman and
signaling; decreased Hinton (1978)
autonomic arousal
Cocaine treatmentc Counts increased >5-fold Hamster Ricci et al. (2005) Cocaine-induced Licata et al. (1993)
aggression
Repeated maternal Juveniles: attack counts Rat Veenema et al. Violent and criminal Widom (1989)
separation doubled; large increases in (2006) behavior following early
nape attacks social neglect
Post-weaning social Attacks in inappropriate Mouse Bibancos et al. Aggression problems Chapple et al.
isolation situationsb (2007) following early social (2005)
neglect
Post-weaning social Counts increased 2-folds; Rat Toth et al. (2008) Aggression problems Chapple et al.
isolation vulnerable targets following early social (2005)
attacked; deficient attack neglect
signaling; offensive
ambiguity; hyperarousal
Anabolic steroid Share of aggressive mice Mouse Ambar and Hostility and irritability Su et al. (1993)
treatmentc doubled Chiavegatto (2009) due to anabolic
consumption
Amphetamine Females attacked Vole Gobrogge et al. Amphetamine-induced Cherek et al. (1986)
treatmenta (2009) aggression
Lines selected for Counts increase 2-fold; Rat Neumann et al. Violence elicited by faulty Davidson et al.
anxiety vulnerable targets; (2010) emotion control (2000)
females/narcotized
opponents attacked

Notes. Some of these models were used to study phenomena unrelated aggression; we referred to the first studies addressing abnormal features of aggression.
Stimulation models may fulfill most criteria based on which normal and abnormal manifestations of aggression are differentiated. Nevertheless, this induced behavior may
have little in common with spontaneously occurring abnormal forms of behavior, for which conclusions based on such studies are valuable as it regards the brain mechanisms
of aggression per se, but their conclusions should be dealt with caution as it regards abnormal forms of aggression.
a
For 3 days in adults.
b
Attacks on docile opponents in neutral arenas.
c
Repeatedly during adolescence.

the lateral hypothalamic area induces frog and/or mouse killing in It was recently demonstrated that the optogenetic but not elec-
rats (Vergnes and Karli, 1969; Smith et al., 1970; Bandler, 1970; trical stimulation of the lateral aspects of the ventromedial nucleus
Woodworth, 1971), and recent findings from our laboratory show rapidly induces coordinated and directed biting in mice (Lin et al.,
that the lateral hypothalamus is strongly activated by mouse killing 2011). Behaviorally, this response is different from that seen in the
(Tulogdi et al., 2012), the role of the lateral hypothalamus appears rat, as mice do attack inanimate objects upon stimulation (e.g. a
somewhat different in cats and rats. E.g. some of the early stud- glove), while rats do not (Siegel et al., 1999). Although bites could
ies revealed that the stimulation of the lateral hypothalamus does be elicited from a very restricted hypothalamic area in the mouse, a
induce attacks on conspecifics albeit this response required consid- wider array of hypothalamic mechanisms are activated by territo-
erably larger current intensities than muricide (Woodworth, 1971). rial aggression in this species, suggesting that many hypothalamic
Although no attacks on conspecifics were induced from the lat- structures analogous to the rat hypothalamic attack area are also
eral hypothalamus in other labs (Kruk, 1991; Siegel et al., 1999), it involved in mouse aggression (e.g. the posterior part of the ante-
was shown that the stimulation of the lateral hypothalamus pro- rior hypothalamus, the ventrolateral nucleus of the hypothalamus
motes spontaneous attacks under certain circumstances (Koolhaas, and the tuber cinereum; Duncan et al., 2009; Haller et al., 2006).
1978). The assumption that the lateral hypothalamus is involved It was also shown that individual variation in estrogen receptor ␣
in intraspecific aggression is also supported by c-Fos and brain immunoreactivity in the anterior hypothalamus (part of “hypothal-
imaging studies showing that the lateral hypothalamic area is acti- amic attack area” in rats) is positively correlated with aggressive
vated during resident-intruder conflicts (Clinton et al., 2011; Ferris behavior in mice (Trainor et al., 2006). Moreover, disparate findings
et al., 2008; Tulogdi et al., 2010). Taken together, these findings suggest that the lateral hypothalamus also plays a role in mouse
show that intraspecific aggression and predatory aggression are aggression (Duncan et al., 2009; Hasen and Gammie, 2006). Taken
controlled by hypothalamic areas medial and lateral to the fornix, together, these findings suggest that mouse aggression is controlled
respectively, but the two mechanisms are not as distinct as in the by hypothalamic structures analogous to those controlling aggres-
cat. sion in rats. It is unclear why electric stimulation does not evoke
J. Haller / Brain Research Bulletin 93 (2013) 97–109 101

aggression in mice and why the “optogenetic attack area” is so 3. The neurochemistry of attack-related hypothalamic
limited in this species. mechanisms
In hamsters, the location and extent of hypothalamic struc-
tures involved in aggressiveness were not precisely delimited, In the cat, defensive rage (emotional aggression) is believed
but it is believed that the key regions are the anterior and to be induced by glutamatergic neurons of the mediobasal hypo-
lateral hypothalamic areas which control intraspecific attacks thalamus which project to the periaqueductal gray (Siegel et al.,
(Ferris and Delville, 1994; Knyshevski et al., 2005). Despite 1999, 2007). At the same time, the mediobasal hypothalamus
the poor anatomical description of attack-related hypothalamic inhibits the predation-related lateral hypothalamus via GABAer-
structures, the neurochemical mechanisms subserving attacks gic projections. The area is stimulated by substance Pergic medial
in the anterior hypothalamus were extensively studied in this amygdala neurons and by dopaminergic and noradrenergic inputs.
species. The neurochemical nature of lateral hypothalamic neurons that
The earliest evidence on the involvement of the human hypo- elicit predatory attacks remains unclear. These neurons are under
thalamus in aggression came from the study by Sano et al. (1970) GABAergic projections received from the mediobasal hypothala-
who lesioned the so-called posteromedial hypothalamus and by mus, a pathway that also seems to mediate the indirect suppression
this procedure successfully reduced or abolished aggressiveness of predatory attacks by the Substance Pergic neurons of the medial
in violent patients (Fig. 2). Note that the lesioned region (the amygdala. Both hypothalamic sites are facilitated by dopaminer-
location was somewhat patient-specific) included medial parts gic and cholinergic inputs and inhibited by serotonin. The role of
of the hypothalamus, but also the medial subdivision of the lat- serotonin appears complex as effects exerted via different recep-
eral hypothalamus. The procedure provided excellent results in tors (e.g. 5-HT1A and 5HT2C ) affect aggression in an opposite fashion
subsequent studies as well (Dieckmann et al., 1988; Laitinen, (Hassanain et al., 2003).
2001; Pedrosa-Sanchez and Sola, 2003; Ramamurthi, 1988). A The neurochemistry of hypothalamic mechanisms underlying
case study suggested that the electrical stimulation of the very predation is poorly known in the rat. The mediobasal hypothalamus
same area induces aggressiveness (Bejjani et al., 2002). The behav- (involved in attacks on conspecifics) contains both glutamatergic
ioral response was similar to that seen in animals as it occurred and GABAergic neurons that have a distinct localization within
within seconds, and had no post-stimulation effects on aggres- the area (Hrabovszky et al., 2005). The same study showed that
sion. Importantly, the patient – although clearly remembered it–, a subgroup of glutamatergic neurons also expresses thyrotropin
was not able to explain his behavior, suggesting that aggressive- releasing hormone the role of which remains unclear at present.
ness had no obvious cognitive aspect. Surprisingly, it was recently The local infusion of agonists and antagonists suggest that glu-
shown that the continuous stimulation of the “triangle of Sano” tamatergic neurons are responsible for the induction of attacks
also suppresses aggression in patients with intractable aggression whereas GABAergic neurons locally inhibit attacks (Adams et al.,
(Franzini et al., 2008, 2010; Hernando et al., 2008). Stimulation 1993; Haller et al., 1998; Roeling et al., 1993). Stimulation studies
was continuously applied for prolonged periods (1–2 years), the show that the main input that negatively controls the neurons of
anti-aggressive effect developed slowly over time, and remained the area are serotonergic in nature (Kruk, 1991). The area is under
stable till the stimulation was applied but relapse occurred when the influence of Substance P projections (likely originating from
it was terminated. At the first sight, reduced violence obtained the medial amygdala; Halasz et al., 2008, 2009), of vasopressiner-
by the Sano and the Franzini approaches is intriguing, as lesions gic inputs (Caffrey et al., 2010; Ferris et al., 2008) and is sensitive
and stimulations produce ostensibly opposite effects: one deac- to testosterone (Trainor et al., 2006).
tivates, while the other activates the very same brain region. In hamsters, the anterior hypothalamus (i.e. the “hypothalamic
It is believed, however, that deep brain stimulation is in fact attack area” in this species) is under the control of vasopressin-
a functional lesion (Bennazzous and Hallett, 2000), which ren- ergic inputs (Albers et al., 2006; Ferris and Delville, 1994; Ferris
ders the two methodologies similar in their ultimate outcome. and Potegal, 1988), that seem to integrate other neurochemical and
An interesting example of hypothalamic aggression control is the hormonal influences, namely those of serotonin and testosterone
hypothalamic hamartoma. A subgroup of patients developing this (Delville et al., 1996; Ferris et al., 1997). Attack-related hypothal-
malformation show increased aggression, the removal of which amic mechanisms are also sensitive to the local application of
abolishes aggressiveness (de Almeida et al., 2008; Weissenberger glucocorticoids (Hayden-Hixson and Ferris, 1991).
et al., 2001). The impact of hamartomas on hypothalamic func- Summary. Taken together, these findings show that hypothal-
tion is poorly known, but the findings with this malformation amic mechanisms involved in attack are under the influence
show that hypothalamic functions are tightly linked to aggres- of neural inputs utilizing the neurotransmitters acetylcholine,
sion in humans. Taken together, these findings suggest that there dopamine, GABA, glutamate, noradrenaline, serotonin, substance P
is a region in the human hypothalamus that controls aggres- and vasopressin. These mechanisms are locally controlled by gluco-
siveness, and disparate findings suggest that this area responds corticoids and testosterone. Findings obtained in different species
to stimulation in a manner similar to that seen in laboratory show large overlaps; seemingly species-dependent mechanisms
animals. are likely due to the paucity of studies rather than to real species
Summary. The findings reviewed show that there is a region in differences. E.g. glucocorticoids likely control the function of the rat
the mediobasal hypothalamus, which is tightly involved in the exe- mediobasal hypothalamus (Kruk et al., 2004), although direct evi-
cution of attacks on conspecifics. Although the extent and precise dence (studies employing the local application of corticosterone)
location of this region shows species differences and the behav- are lacking. Although caution is still advisable, the data suggest
iors elicited by its stimulation is also species-dependent, overall, that the neurochemical mechanisms governing aggression-related
its role appears evolutionarily conserved. The role of the lateral hypothalamic areas are universal.
hypothalamic area is more species dependent. In the cat, it belongs
to a topographically different neural network that specifically con- 4. Hypothalamic mechanisms of abnormal aggression
trols predatory aggression. In the rat, it directly controls predatory
attacks but also affects intraspecific aggression. Predatory attack Aggression was studied mainly by stimulation techniques in
centers were not described for mice, hamsters and humans, but cats. Data on natural forms and abnormal manifestations of aggres-
the lateral hypothalamus has a role in aggression control in these siveness are extremely scarce in this species. In contrast, models of
species as well. abnormal aggression became increasingly popular in rodents over
102 J. Haller / Brain Research Bulletin 93 (2013) 97–109

Fig. 2. The hypothalamic region involved in human aggression (based on Sano et al., 1970; coronal section from Mai et al., 1998; 1.3 mm posterior to the midpoint of the
intercomissural line, Plate 23). The red circle indicates the approximate area of the lesion that strongly inhibited intractable human aggression. Note that the name (the
“triangle of Sano”) derives from the shape of electrophysiologically identified brain regions; the lesion per se was circular. 3rd V, 3rd ventricle; BST, bed nucleus of stria
terminalis; Do, dorsal hypothalamic nucleus; LHA, lateral hypothalamic area; LV, lateral ventricle; F, fornix; OT, optic tract.

the last decade. The main characteristics of the abnormal aggres- Tulogdi et al., 2010). Lateral hypothalamic activation was not
sion models reviewed here were shown in Table 1. increased in rats selected for high anxiety, and those submitted to
post-weaning social isolation (Beiderbeck et al., 2012; Toth et al.,
4.1. Overall activation of the hypothalamus 2012). In rats selected for high anxiety neither the mediobasal nor
the lateral hypothalamus was over-activated despite clearly abnor-
One can hypothesize that hypothalamic structures involved in mal attack patterns (Beiderbeck et al., 2012).
attack were overactivated in abnormal aggression if the “hyper- Such discrepancies are difficult to reconcile at present. One obvi-
activation theory” was right, i.e. if abnormal aggression resulted ous conclusion is that the overall increase in aggression-induced
from the exacerbated activation of aggression-controlling brain hypothalamic activation – as shown by c-Fos counts – is not a
regions. However, hypothalamic mechanisms controlling attack general concomitant of abnormal aggression. However, an interest-
underwent interesting, sometimes surprising changes in rodent ing picture emerges when the emotional component of aggression
models of abnormal aggression. is considered. The mediobasal hypothalamus was over-activated
Supporting data. The mediobasal hypothalamus (involved in mainly in models where aggression is associated with hyper-
rivalry attacks) showed increased aggression-induced c-Fos acti- arousal e.g. the post-weaning social isolation and cocaine models
vation in four models, namely in the post-weaning social isolation, both being characterized by increased glucocorticoid and auto-
anabolic steroid, and cocaine models, as well as in rats selected for nomic stress responses (Ansah et al., 1996; Moldow and Fischman,
low anxiety (Toth et al., 2012; Ricci et al., 2007; Knyshevski et al., 1987; Toth et al., 2011). In contrast, the mediobasal hypothal-
2005; Beiderbeck et al., 2012). As it regards the lateral hypothala- amus was not activated in models characterized either by low
mus (implicated mainly in predatory attacks), increased activation glucocorticoid and autonomic stress responses or by low basal
was seen in three models (the hypoarousal and cocaine model, as glucocorticoid levels and heart rates, i.e. by hypoarousal (rat:
well as in low anxiety rats (Beiderbeck et al., 2012; Knyshevski et al., Haller et al., 2004, 2001; mice: Caramaschi et al., 2008; Veenema
2005; Tulogdi et al., 2010). et al., 2004). Although aggression-induced autonomic activation
Non-supporting data. No changes in mediobasal hypothalamic was not directly investigated in rats selected for anxiety, studies
activation was seen in the hypoarousal model, in mice selected on glucocorticoids support the assumption that hypo and hyper-
for high aggressiveness, and in rats selected for high anxiety arousal during aggressive confrontations have an impact on the
(Beiderbeck et al., 2012; Haller et al., 2006; Halasz et al., 2002a; activation of hypothalamic centers involved in attack. In highly
J. Haller / Brain Research Bulletin 93 (2013) 97–109 103

Fig. 3. The association between emotionality and aggression-induced hypothalamic activation patterns as illustrated by a comparison between the glucocorticoid deficiency-
induced hypoarousal model (Tulogdi et al., 2010) with the post-weaning social isolation-induced hyperarousal model (Toth et al., 2012). The two models are directly comparable
as subjects were rats in both cases, behavioral and neuroanatomical analysis was highly consistent, and the findings were obtained by the same group. Upper panel: emotional
and behavioral characteristics in the two models. Lower panel: three-dimensional reconstruction of the relevant hypothalamic region. pink, activation similar to and red,
over-activation compared to that seen in normal rivalry aggression; 3rdV, 3rd ventricle; F, fornix; HAA, hypothalamic attack area (mediobasal hypothalamus); LHA, lateral
hypothalamic area; PVN, paraventricular nucleus of the hypothalamus.

anxious rats that attacked vulnerable targets, the glucocor- are lacking in mice, while the arousal component of aggression is
ticoid response to resident-intruder conflicts was decreased poorly known in other models. Nevertheless the findings briefly
and no aggression-induced hypothalamic activation was noticed reviewed above and the example summarized in Fig. 3 suggests that
(Neumann et al., 2010). In the low anxiety strain that showed the emotional component of abnormal attacks has a large impact
vulnerable attacks and also attacked females and anesthetized on the hypothalamic control of aggression.
opponents, the glucocorticoid response to resident-intruder con-
flicts increased, and the hypothalamus was strongly activated by 4.2. Serotonin
fights.
Summary. The findings briefly reviewed above suggest that The “serotonin hypothesis” is considered a cornerstone of
hypoarousal-associated aggression does not, while hyperarousal- aggression control (Nelson and Chiavegatto, 2001). Theoretically,
associated aggression does correlate with increased mediobasal abnormal aggression may be driven by decreased serotonergic
hypothalamic activation. This relationship is illustrated in Fig. 3 by control, which would indirectly upregulate the function of brain
the comparison of the rat hypoarousal and post-weaning social iso- mechanisms that promote aggression. However, the role of hypo-
lation model. The direct comparability of these models is ensured thalamic serotonin is unexpectedly complex in models of abnormal
by the consistency of behavioral and neuroanatomical analysis, as aggression.
the findings were obtained by the same group. In the hypoarousal Supporting data. In a few models, findings support the notion
model, normal attack counts but abnormal attack features are that serotonin suppresses aggression. In these models (mice
associated with a “predatory-like” hypothalamic activation pattern submitted to post-weaning social isolation, rats submitted to
(normal mediobasal but increased lateral hypothalamic activation; maternal separation and anabolic steroid-treated hamsters) sero-
Tulogdi et al., 2010). In contrast, exacerbated “rivalry-like” hypo- tonin fiber density, and/or the density of various serotonin
thalamic activation pattern was seen in the post-weaning social receptors decreased in the mediobasal hypothalamus that is
isolation (hyperarousal) model (increased mediobasal but normal involved in the control of rivalry attacks (Bibancos et al., 2007;
lateral hypothalamic activation; Toth et al., 2012). Unfortunately, Grimes and Melloni, 2002, 2005; Ricci et al., 2006; Veenema et al.,
evidence on predation-related hypothalamic control mechanisms 2006). Some contradictions do exist; e.g. 5HT1A receptor density
104 J. Haller / Brain Research Bulletin 93 (2013) 97–109

Fig. 4. Serotonin neurotransmission often shows opposite changes in different models of abnormal aggression as illustrated by a comparison between the post-weaning
social isolation and the alcohol-heightened aggression model. The two models are directly comparable as the subjects were mice in both cases, the immunocytochemical
methodologies were highly similar, and the findings were obtained by an overlapping set of authors. Abnormal features of aggression were observed in both models (see
Table 1). The figure was compiled from data published by Bibancos et al. (2007) and Chiavegatto et al. (2010). The Y-axis shows the difference between transcript amounts
obtained in the experimental groups and their controls. *Significantly different from controls in the respective publications.

decreased in hamsters treated with anabolic steroids, but remained in the hypoarousal model which was in sharp contrast with data
unchanged in mice submitted to post-weaning social isolation obtained in controls (Haller et al., 2005). In addition, the 5-HT1A
(Bibancos et al., 2007; Ricci et al., 2006). Overall, however, a partial agonist buspirone ameliorated rivalry aggression but dra-
decrease in hypothalamic serotonin functions is evident in these matically increased attack counts in the hypoarousal model (Haller
models. et al., 2007). In a similar fashion, there was no obvious link between
Non-supporting data. Surprisingly, hypothalamic 5-HT1B and 5- serotonergic neurotransmission and aggressiveness in mice selec-
HT2A receptor mRNA levels doubled in mice in which alcohol tively bred for aggressiveness (Natarajan et al., 2009; Wallinga et al.,
exacerbated aggressive responses (Chiavegatto et al., 2010). Hypo- 2009). Moreover, 5-HT1A and 5-HT1B agonists decreased aggres-
thalamic 5-HT6 receptor expression also doubled in mice treated siveness in highly aggressive feral rats by decreasing serotonin
with anabolic steroids (Ambar and Chiavegatto, 2009), while the release via their presynaptic autoreceptor function (de Boer and
density of serotonin fibers and the number of serotonergic varicosi- Koolhaas, 2005).
ties increased in the anterior hypothalamus of hamsters repeatedly Summary. Taken together, these findings demonstrate that the
defeated during adolescence and that showed “offensive ambigu- anti-aggressive effects of serotonergic neurotransmission are com-
ity” as adults (Delville et al., 1998). Similar findings were obtained promised in certain models of abnormal aggression; moreover,
with N-type Ca channel alpha1B subunit-KO mice where attack serotonin may paradoxically promote aggressiveness in some mod-
counts increased over 5-times yet the serotonin content of the els. Interestingly, these paradoxical interactions between serotonin
hypothalamus doubled without parallel increases in hypothala- and aggressiveness were shown in models where the abnormal
mic noradrenaline and dopamine (Kim et al., 2009). One has to aggression-promoting manipulation was associated with signs of
conclude that certain forms of abnormal aggression are associated hypoarousal (blunted cortisol response to aggression in subjugated
with increased serotonergic neurotransmission in the hypothala- hamsters: Ferris et al., 2005; low basal corticosterone after adolescent
mus. Findings with the post-weaning social isolation and alcohol anabolic steroid treatment: Rejeski et al., 1990; low corticosterone
models are especially intriguing, as the subjects were mice in both stress responses and low basal heart rate in mice selected for high
cases, the findings were obtained by an overlapping set of authors, aggressiveness: Caramaschi et al., 2008; Veenema et al., 2004; low
and the immunocytochemical methodologies were highly similar basal CORT and low autonomic responses to aggression in the rat
(Fig. 4). hypoarousal model: Haller et al., 2001, 2004).
One can argue that the intriguing changes seen at hypothalamic
level were “overwritten” by reduced serotonergic neurotransmis- 4.3. Vasopressin
sion in other brain areas e.g. the prefrontal cortex (Chiavegatto et al.,
2010). However, this argument does not hold true for two reasons. Vasopressin plays important roles in aggression control (see
Firstly, increased serotonergic neurotransmission in abnormally above); as such, one can hypothesize that increased vasopressin
aggressive animals was seen in extra-hypothalamic sites as well. neurotransmission promotes abnormal forms of aggression. This,
E.g. Delville et al. (1998) observed increased number of septal 5-HT however, is not always the case.
varicosities in the hamster “subjugation” model. Chiavegatto et al. Supporting data. Vasopressin neurotransmission appears to pro-
(2010) showed that in addition to the hypothalamus, the expres- mote abnormal manifestations of aggression in several models.
sion of 5-HT1B receptors was increased in the amygdala of mice The interaction has many facets. In the anabolic steroid model,
that showed excessive increases in aggression after alcohol, while the expression of the V1A vasopressin receptor increased in the
Bibancos et al. (2007) reported increased 5-HT6 receptor expres- ventrolateral but not in the anterior hypothalamus (DeLeon et al.,
sion in the hippocampus of mice submitted to the post-weaning 2002). Despite this, V1A antagonists administered into the anterior
isolation model. Secondly, paradoxical interactions were noticed hypothalamus decreased the anabolic steroid-induced aggression
between serotonergic neurotransmission and aggression in two (Harrison et al., 2000). In the amphetamine model, V1A expres-
abnormal aggression models. E.g. there was no significant corre- sion increased in the anterior hypothalamus, and V1A antagonists
lation between raphe serotonergic activation and aggressiveness abolished amphetamine-induced aggression (Gobrogge et al.,
J. Haller / Brain Research Bulletin 93 (2013) 97–109 105

2009). V1A antagonists also abolished exacerbated aggression in model (Ricci et al., 2005) and in mice selected for high aggressive-
the cocaine model, where an increased stimulation-induced vaso- ness (Weerts et al., 1992). In the latter model, findings were in line
pressin release was noticed without changes in vasopressin content with expectations: GABA binding was increased in low aggression
or vasopressin fiber density (Jackson et al., 2005). In the mater- genotype, and was decreased in the highly aggressive genotype. In
nal separation model (the only rat study addressing hypothalamic the former two models, however, increased GABA synthesis was
vasopressin), no change in vasopressin expression was noticed in noticed in the anterior hypothalamus despite the fact that GABA
the anterior hypothalamus, but vasopressin expression (both basal agonists inhibit rivalry aggression when infused into mediobasal
and aggression-induced) increased in the lateral hypothalamus hypothalamic sites (Adams et al., 1993; Roeling et al., 1993).
(Veenema et al., 2006). Despite this variability of findings, increased Summary. The above findings reveal that increased NK1
vasopressin functions are evident in these models. receptor-mediated Substance P and D2 receptor-mediated
Non-supporting data. Intriguingly, decreased anterior hypothal- dopamine neurotransmission promote certain forms of abnor-
amic vasopressin content and decreased vasopressin fiber density mal aggression at the level of the mediobasal hypothalamus.
was noticed in the “subjugation” model (Delville et al., 1998). Some- More conflicting findings were obtained with GABAergic and
what similar findings were obtained in rats selected for anxiety. glutamatergic neurotransmission.
Rats selected for low anxiety generally express a vasopressin gene
that has a reduced transcription rate and consequently have lower
4.5. The human case
vasopressin expression levels than the high anxiety strain (Bunck
et al., 2009; Kessler et al., 2007). As the attack patterns of the
Unfortunately, human data on the role of the hypothalamus in
low anxiety strain are abnormal from more points of view than
abnormal forms of aggression are extremely scarce. It was shown
that of the high anxiety strain, this finding suggests that a higher
in psychopaths that hypothalamic activation correlates positively
abnormality of aggression is paralleled by a lower hypothalamic
with the amount of punishment delivered in response to increased
vasopressin release.
punishment received from a yoked opponent; the response also
Summary. Although the information is incomplete, the role
correlated with increased anger (Veit et al., 2010). An earlier study
of hypothalamic vasopressin in abnormal aggression is also
performed by the same set of authors and employing similar tech-
model-dependent: while it clearly promotes abnormal features
niques showed no such a correlation in healthy controls (Lotze et al.,
of aggression in several models (anabolic steroid, amphetamine,
2007). As anger per se does not induce hypothalamic activation
cocaine, and maternal separation models) the opposite is true for
(Damasio et al., 2000), these findings suggest that the delivery of
the early subjugation model and likely for abnormally aggressive
severe punishment correlates with increased hypothalamic activa-
rats selected for low anxiety behavior.
tion in angry psychopaths but not in angry healthy subjects. A very
recent publication showed that gray matter was increased in the
4.4. Other neurotransmitters
hypothalamus of borderline personality patients, and this increase
correlated positively with the history of traumatization (Kuhlmann
There are neurotransmitter systems that were studied in one
et al., 2012). As the aggressiveness of subjects (similar to border-
or a few models only. This does not diminish the importance of
line personality disordered patients in general) showed a more than
the findings, or the relevance of the respective systems for abnor-
two fold increase, this finding tentatively suggests that traumatic
mal aggression, but the various models of abnormal aggression
experience results in an increased hypothalamic gray matter, which
cannot be compared in their case. E.g. the impact of Substance
may contribute to the expression of aggressiveness. In another
P neurotransmission (the effects of which are mediated by neu-
study, borderline personality patients showed decreased seroto-
rokinin 1 (NK1 ) receptors) was investigated in the hypoarousal
nergic functions in the hypothalamus as indicated by increased
model only (Halasz et al., 2008, 2009). In this model, mediobasal
number of available binding sites for serotonin transporters and/or
hypothalamic neurons expressing NK1 receptors were specif-
decreased competition by lower endogenous serotonin levels (Koch
ically over-activated by aggressive encounters; moreover, the
et al., 2007).
systemic application of NK1 blockers or the selective destruc-
Summary. The scarcity of data precludes firm conclusions on
tion of NK1 receptor expressing mediobasal hypothalamic neurons
the role of the hypothalamus in aggression-related psychopatholo-
by saporin-conjugated Substance-P dramatically reduced violent
gies in humans. In addition, the spatial resolution of current
biting, without affecting milder forms of aggression. Thus, NK1-
brain imaging techniques does not distinguish the various sub-
expressing mediobasal hypothalamic neurons appear specifically
regions of the hypothalamus, which have rather different roles
involved in abnormal features of aggression in this model.
in brain function and behavior. Tentatively, however, data sug-
The role of dopamine was investigated in the anabolic steroid
gest that aggression-related psychopathologies are associated with
model (Ricci et al., 2009; Schwartzer et al., 2009). It was shown that
increased hypothalamic gray matter volume, increased hypothal-
the dopamine production by two hypothalamic nuclei (the nucleus
amic activation during retaliation, and decreased serotonergic
circularis and medial supraoptic nucleus) is increased in this model
control at the level of the hypothalamus. As the aggressiveness of
and the expression of the D2 dopamine receptor is increased in
psychopaths correlated with anger that was induced by the experi-
the anterior hypothalamus. It was suggested that dopamine neuro-
mental punishment received, while borderline patients show signs
transmission promotes abnormally high levels of aggressiveness in
of hyperarousal in terms of both glucocorticoid production and
this model via a mechanism that involves reduced GABAA receptor
autonomic activation (Carvalho et al., 2012; Lyons-Ruth et al., 2011;
expression at the level of the anterior hypothalamus (Schwartzer
Weinberg et al., 2009), the above inference seems to be valid to
et al., 2009). This suggests a disinhibition-like phenomenon at this
emotional/reactive (hyperarousal-driven) aggression.
attack-controlling hypothalamic site.
Finally, the role of glutamatergic and GABAergic neurotransmis-
sion was also investigated in a few models. In the hamster anabolic 5. Etiological factor-dependent neuronal plasticity and
steroid model, glutamate synthesis increased in both the anterior abnormal aggression
and lateral hypothalamus, but equivocal changes were noticed at
the level of their projections (Carrillo et al., 2009; Fischer et al., Intriguingly, the neural changes noticed in the models presented
2007). The role of GABA was investigated in the anabolic steroid here are not temporally linked to the etiological factor that led to
model (Grimes et al., 2003; Schwartzer et al., 2009), the cocaine the development of abnormal aggression. This is especially blatant
106 J. Haller / Brain Research Bulletin 93 (2013) 97–109

Table 2
Neural changes in mediobasal hypothalamic regions in models of abnormal aggression and aggression-related psychopathologies. The models were grouped according to
emotional background.

Model c-Fos 5-HT AVP NK1 GLU GABA DA Arousal

a
Post-weaning social isolation High
Selection for low anxiety High
Early maternal separation High
Cocaine in adolescence High
Amphetamine in adolescence High
b,a
Psychopaths (moderate level) Highc
d a
Borderline personality disorder Highe

Selection for high anxiety Low


Selection for high aggression Low
Glucocorticoid deficiency Low
a
Repeated defeat in adolescence Low

Anabolic steroid during adolescence Unknown


a
Acute alcohol Unknown

( ) increase; ( ) no change; ( ) decrease; (empty cells) no data available; (high arousal) high basal glucocorticoid levels or high glucocorticoid response to aggression
and/or high basal heart rates or high autonomic response to aggression; (low arousal) the opposite pattern; (unknown), not investigated in the specific conditions of the
model.
a
Whole hypothalamus.
b
Activation evaluated by fMRI.
c
Increased anger during punishment/retaliation testing.
d
Increased gray matter volume.
e
General feature (not investigated in the study).

with developmental models where treatments (anabolic steroids, extra-synaptic sites, and by this, can control relatively large popula-
cocaine, amphetamine, maternal separation, defeat) are restricted tions of target cells (dopamine: Caillé et al., 1996; GABA: Olah et al.,
to early periods of life, whereas effects on aggression are expressed 2009; glutamate: Kew et al., 1998; serotonin: Hervé et al., 1987;
in adulthood. A recent study suggests that this is valid also for the also see Vizi et al., 2010 for a review). As hypothesized very early
post-weaning social isolation model, where social deficits were and confirmed later, such non-synaptic receptors not only play
rapidly abolished by long-term re-socialization in adulthood, but an important role in neuronal communication, but preferentially
abnormal features of aggression were resistant to this treatment mediate the effects of pharmacological agents used in psychia-
(Tulogdi et al., 2012). try (Vizi, 1984, 2000). The differentiation of the role of synaptic
The durable effects seen in developmental models suggest that and non-synaptic communication appears especially important for
the behavioral alterations were due to long-term neural plastic- serotonergic neurotransmission, as the neurons of the dorsal and
ity, possibly to epigenetic phenomena. This is strongly supported median raphe (the two main sources of forebrain serotonergic
by the fact that the procedures employed (early stressors, altered fibers) preferentially communicate non-synaptically and synap-
rearing conditions, as well as anabolic steroids and psychostimu- tically, respectively. Their differential study may contribute to a
lants administered during adolescence) induce both synaptic and better understanding of the complex role of serotonergic neuro-
morphological plasticity in the brain regions and neurotransmis- transmission in aggression control.
sion processes relevant for aggression control (hypothalamus on
one side and GABA, glutamate, serotonin, dopamine and vaso- 6. Summary and conclusions
pressin neurotransmission on the other side) (Adriani et al., 2006;
Beauchaine et al., 2011; Grimes and Melloni, 2006; Herman et al., We reviewed here the neuronal mechanisms underlying
2008; Joëls et al., 2004). The same early treatments have epigenetic abnormal manifestations of aggression in 12 models. Each
consequences as well (Bountra et al., 2011; Lesch, 2011). model mimicked an etiological factor of aggression-related psy-
It was shown that the impact of stressors on neural plasticity is chopathologies, and the factors employed induced abnormal
stressor-specific (Kavushansky et al., 2009), and the same applies manifestations of behavior as evidenced by a mismatch between
to early drug treatments: plastic neural changes are drug-specific provocation and response, disregarded species-specific rules,
(Gould, 2010). Therefore, one can hypothesize that the mecha- insensitivity toward the social signals of the opponent, and offen-
nisms activated by developmental models of abnormal aggression sive ambiguity (see Section 1). Multiple abnormalities were present
induce developmental factor-dependent neuronal plasticity; con- in most models.
sequently, the neural underpinnings of abnormal aggression that We focused on the hypothalamus that was tightly involved in
develops in response to this plasticity is also etiological factor- the execution of attacks in all the species studied so far includ-
dependent. ing humans. At the first sight, the models provided conflicting
It is premature to speculate on the precise mechanisms or the data. The overall activation of the mediobasal and lateral hypo-
general nature of the plastic changes that result from the treat- thalamus was increased in some models but not in others, while
ments that promote abnormal manifestations of aggressiveness. serotonergic neurotransmission decreased in some models, but
We outline here only one aspect, namely non-synaptic transmis- paradoxically, the opposite happened in other models. Conflicting
sion. All the neurotransmitters critically involved in the control of findings were reported with hypothalamic vasopressinergic and
aggression (dopamine, GABA, glutamate, serotonin, or vasopressin) GABAergic neurotransmission as well. As shown above, however,
reach relatively distant, high-affinity receptors that are located at the conflict between findings seems to resolve when the emotional
J. Haller / Brain Research Bulletin 93 (2013) 97–109 107

background of abnormal attacks is considered. In Table 2, we sum- Ansah, T.A., Wade, L.H., Shockley, D.C., 1996. Changes in locomotor activity, core
marized findings on the mediobasal hypothalamus by grouping the temperature, and heart rate in response to repeated cocaine administration.
Physiology and Behavior 60, 1261–1267.
models according to their emotional background. Unfortunately, a Bandler Jr., R.J., 1970. Cholinergic synapses in the lateral hypothalamus for the con-
similarly detailed table cannot be compiled for the lateral hypo- trol of predatory aggression in the rat. Brain Research 20, 409–424.
thalamus because of the paucity of data. The data summarized in Beauchaine, T.P., Neuhaus, E., Zalewski, M., Crowell, S.E., Potapova, N., 2011. The
effects of allostatic load on neural systems subserving motivation, mood regu-
Table 2 and disparate data on the lateral hypothalamus suggest lation, and social affiliation. Development and Psychopathology 23, 975–999.
that abnormal aggression is associated with two basic patterns of Beiderbeck, D.I., Reber, S.O., Havasi, A., Bredewold, R., Veenema, A.H., Neumann,
aggression-induced hypothalamic activation: I.D., 2012. High and abnormal forms of aggression in rats with extremes in
trait anxiety – involvement of the dopamine system in the nucleus accumbens.
Psychoneuroendocrinology [Epub ahead of print] PMID: 22608548.
(i) When abnormal aggression is performed on the background Bejjani, B.P., Houeto, J.L., Hariz, M., Yelnik, J., Mesnage, V., Bonnet, A.M., Pidoux, B.,
of increased arousal, the overall activation of the mediobasal Dormont, D., Cornu, P., Agid, Y., 2002. Aggressive behavior induced by intraop-
erative stimulation in the triangle of Sano. Neurology 59, 1425–1427.
hypothalamus increases, the serotonergic control of the Bennazzous, A., Hallett, M., 2000. Mechanism of action of deep brain stimulation.
hypothalamus diminishes, and hypothalamic vasopressinergic Neurology 55, S13–S16.
neurotransmission increases. Bennett, R.M., Buss, A.H., Carpenter, J.A., 1969. Alcohol and human physical aggres-
sion. Quarterly Journal of Studies on Alcohol 30, 870–876.
(ii) In aggression models associated with low arousal, the lat- Benus, R.F., Bohus, B., Koolhaas, J.M., van Oortmerssen, G.A., 1991. Behavioural
eral but not the mediobasal hypothalamus is over-activated, differences between artificially selected aggressive and non-aggressive mice:
the anti-aggressive effects of serotonergic neurotransmis- response to apomorphine. Behavioural Brain Research 43, 203–208.
Bibancos, T., Jardim, D.L., Aneas, I., Chiavegatto, S., 2007. Social isolation and expres-
sion appears lost and vasopressinergic neurotransmission sion of serotonergic neurotransmission-related genes in several brain areas of
decreases. male mice. Genes, Brain, and Behavior 6, 529–539.
Bountra, C., Oppermann, U., Heightman, T.D., 2011. Animal models of epigenetic
regulation in neuropsychiatric disorders. Current Topics in Behavioral Neuro-
Thus, the neural backgrounds of various forms of abnormal sciences 7, 281–322.
aggression show qualitative differences, undermining the tacit Brutus, M., Siegel, A., 1989. Effects of the opiate antagonist naloxone upon hypothala-
assumption that the brain mechanisms of normal and abnormal mically elicited affective defense behavior in the cat. Behavioural Brain Research
33, 23–32.
aggression differ in quantitative terms only. Based on emo- Bunck, M., Czibere, L., Horvath, C., Graf, C., Frank, E., Kessler, M.S., Murgatroyd, C.,
tional backgrounds, we identified two basic mechanisms, one for Müller-Myhsok, B., Gonik, M., Weber, P., Pütz, B., Muigg, P., Panhuysen, M., Singe-
emotional (reactive) and another for “unemotional” (proactive) wald, N., Bettecken, T., Deussing, J.M., Holsboer, F., Spengler, D., Landgraf, R.,
2009. A hypomorphic vasopressin allele prevents anxiety-related behavior. PLoS
aggression. We propose here that findings obtained with etiologi- One 4, e5129.
cally and phenotypically validated models of abnormal aggression Caffrey, M.K., Nephew, B.C., Febo, M., 2010. Central vasopressin V1a receptors
should be considered alternative “neurobiological roads” to abnor- modulate neural processing in mothers facing intruder threat to pups. Neu-
ropharmacology 58, 107–116.
mal aggression rather than conflicting data. This proposal is Caillé, I., Dumartin, B., Bloch, B., 1996. Ultrastructural localization of D1 dopamine
supported by findings obtained with different models by the same receptor immunoreactivity in rat striatonigral neurons and its relation with
groups, where divergent findings cannot be attributed to method- dopaminergic innervation. Brain Research 730, 17–31.
Caramaschi, D., de Boer, S.F., Koolhaas, J.M., 2008. Is hyper-aggressiveness associated
ological differences.
with physiological hypoarousal? A comparative study on mouse lines selected
The findings reviewed here point out an important reason for high and low aggressiveness. Physiology and Behavior 95, 591–598.
why the clinical treatment of aggression-related psychopathologies Carrillo, M., Ricci, L.A., Melloni Jr., R.H., 2009. Adolescent anabolic androgenic
steroids reorganize the glutamatergic neural circuitry in the hypothalamus.
remains problematic, namely the divergence of the brain mecha-
Brain Research 1249, 118–127.
nisms that underlie different forms of abnormal aggression. A more Carvalho, F.S., Beblo, T., Schlosser, N., Terfehr, K., Otte, C., Löwe, B., Wolf, O.T.,
thorough understanding of brain mechanisms may reveal subtypes Spitzer, C., Driessen, M., Wingenfeld, K., 2012. Associations of childhood trauma
of the two basic mechanisms proposed above, or may lead to the with hypothalamic–pituitary–adrenal function in borderline personality dis-
order and major depression. Psychoneuroendocrinology [Epub ahead of print]
discovery of novel types of brain control. Nevertheless, the data PMID: 22444624.
obtained so far strongly suggest that abnormal aggression cannot Chapple, C.L., Tyler, K.A., Bersani, B.E., 2005. Child neglect and adolescent violence:
be considered a unitary phenomenon, and the existence of multiple examining the effects of self-control and peer rejection. Violence Victims 20,
39–53.
mechanisms should be assumed in both research and treatment. Cherek, D.R., Steinberg, J.L., Kelly, T.H., Robinson, D., 1986. Effects of d-amphetamine
on aggressive responding of normal male subjects. Psychiatry Research 21,
257–265.
Conflict of interest Cheu, J.W., Siegel, A., 1998. GABA receptor mediated suppression of defensive rage
behavior elicited from the medial hypothalamus of the cat: role of the lateral
The author declares that there are no conflicts of interest. hypothalamus. Brain Research 783, 293–304.
Chiavegatto, S., Quadros, I.M., Ambar, G., Miczek, K.A., 2010. Individual vulnerability
to escalated aggressive behavior by a low dose of alcohol: decreased serotonin
Acknowledgement receptor mRNA in the prefrontal cortex of male mice. Genes, Brain, and Behavior
9, 110–119.
Clinton, S.M., Kerman, I.A., Orr, H.R., Bedrosian, T.A., Abraham, A.D., Simpson, D.N.,
This work was supported by OTKA Grant no. 82069. Watson, S.J., Akil, H., 2011. Pattern of forebrain activation in high novelty-
seeking rats following aggressive encounter. Brain Research 1422, 20–31.
Damasio, A.R., Grabowski, T.J., Bechara, A., Damasio, H., Ponto, L.L., Parvizi, J., Hichwa,
References R.D., 2000. Subcortical and cortical brain activity during the feeling of self-
generated emotions. Nature Neuroscience 3, 1049–1056.
Adams, D.B., Boudriau, W., Cowan, C.W., Kokonowski, C., Oberteuffer, K., Yohay, K., Davidson, R.J., Putnam, K.M., Larson, C.L., 2000. Dysfunction in the neural cir-
1993. Offense produced by chemical stimulation of the anterior hypothalamus cuitry of emotion regulation – a possible prelude to violence. Science 289,
of the rat. Physiology and Behavior 53, 1127–1132. 591–594.
Adriani, W., Leo, D., Greco, D., Rea, M., di Porzio, U., Laviola, G., Perrone-Capano, de Almeida, A.N., Fonoff, E.T., Ballester, G., Teixeira, M.J., Marino Jr., R., 2008.
C., 2006. Methylphenidate administration to adolescent rats determines plastic Stereotactic disconnection of hypothalamic hamartoma to control seizure and
changes on reward-related behavior and striatal gene expression. Neuropsy- behavior disturbance: case report and literature review. Neurosurgical Review
chopharmacology 31, 1946–1956. 31, 343–349.
Albers, H.E., Dean, A., Karom, M.C., Smith, D., Huhman, K.L., 2006. Role of V1a vaso- de Boer, S.F., Koolhaas, J.M., 2005. 5-HT1A and 5-HT1B receptor agonists and
pressin receptors in the control of aggression in Syrian hamsters. Brain Research aggression: a pharmacological challenge of the serotonin deficiency hypothesis.
1073–1074, 425–430. European Journal of Pharmacology 526, 125–139.
Ambar, G., Chiavegatto, S., 2009. Anabolic–androgenic steroid treatment induces DeLeon, K.R., Grimes, J.M., Melloni Jr., R.H., 2002. Repeated anabolic–androgenic
behavioral disinhibition and downregulation of serotonin receptor messenger steroid treatment during adolescence increases vasopressin V(1A) receptor
RNA in the prefrontal cortex and amygdala of male mice. Genes, Brain, and binding in Syrian hamsters: correlation with offensive aggression. Hormones
Behavior 8, 161–173. and Behavior 42, 182–191.
108 J. Haller / Brain Research Bulletin 93 (2013) 97–109

Delville, Y., Mansour, K.M., Ferris, C.F., 1996. Testosterone facilitates aggression by Haller, J., Toth, M., Halasz, J., De Boer, S.F., 2006. Patterns of violent aggression-
modulating vasopressin receptors in the hypothalamus. Physiology and Behav- induced brain c-fos expression in male mice selected for aggressiveness.
ior 60, 25–29. Physiology and Behavior 88, 173–182.
Delville, Y., Melloni Jr., R.H., Ferris, C.F., 1998. Behavioral and neurobiological con- Haller, J., van de Schraaf, J., Kruk, M.R., 2001. Deviant forms of aggression in glu-
sequences of social subjugation during puberty in golden hamsters. Journal of cocorticoid hyporeactive rats: a model for ‘pathological’ aggression? Journal of
Neuroscience 18, 2667–2672. Neuroendocrinology 13, 102–107.
Demski, L.S., 1973. Feeding and aggressive behavior evoked by hypothalamic Harrison, R.J., Connor, D.F., Nowak, C., Nash, K., Melloni Jr., R.H., 2000. Chronic
stimulation in a cichlid fish. Comparative Biochemistry and Physiology, A: Com- anabolic–androgenic steroid treatment during adolescence increases anterior
parative Physiology 44, 685–692. hypothalamic vasopressin and aggression in intact hamsters. Psychoneuroen-
Dieckmann, G., Schneider-Jonietz, B., Schneider, H., 1988. Psychiatric and neuropsy- docrinology 25, 317–338.
chological findings after stereotactic hypothalamotomy, in cases of extreme Hasen, N.S., Gammie, S.C., 2006. Maternal aggression: new insights from Egr-1. Brain
sexual aggressivity. Acta Neurochirurgica: Supplement 44, 163–166. Research 1108, 147–156.
Dodge, K.A., Bates, J.E., Pettit, G.S., 1990. Mechanisms in the cycle of violence. Science Hassanain, M., Bhatt, S., Siegel, A., 2003. Differential modulation of feline defensive
250, 1678–1683. rage behavior in the medial hypothalamus by 5-HT1A and 5-HT2 receptors. Brain
Duncan, G.E., Inada, K., Farrington, J.S., Koller, B.H., Moy, S.S., 2009. Neural activation Research 981, 201–209.
deficits in a mouse genetic model of NMDA receptor hypofunction in tests of Hayden-Hixson, D.M., Ferris, C.F., 1991. Cortisol exerts site-, context- and
social aggression and swim stress. Brain Research 1265, 186–195. dose-dependent effects on agonistic responding in hamsters. Journal of Neu-
Ferris, C.F., Delville, Y., 1994. Vasopressin and serotonin interactions in the control roendocrinology 3, 613–622.
of agonistic behavior. Psychoneuroendocrinology 19, 593–601. Herman, J.P., Flak, J., Jankord, R., 2008. Chronic stress plasticity in the hypothalamic
Ferris, C.F., Potegal, M., 1988. Vasopressin receptor blockade in the anterior hypo- paraventricular nucleus. Progress in Brain Research 170, 353–364.
thalamus suppresses aggression in hamsters. Physiology and Behavior 44, Hernando, V., Pastor, J., Pedrosa, M., Peña, E., Sola, R.G., 2008. Low-frequency bilateral
235–239. hypothalamic stimulation for treatment of drug-resistant aggressiveness in a
Ferris, C.F., Melloni, R.H., Koppel, G., Perry, K.W., Fuller, R.W., Delville, Y., 1997. Vaso- young man with mental retardation. Stereotactic and Functional Neurosurgery
pressin/serotonin interactions in the anterior hypothalamus control aggressive 86, 219–223.
behavior in golden hamsters. Journal of Neuroscience 17, 4331–4340. Hervé, D., Pickel, V.M., Joh, T.H., Beaudet, A., 1987. Serotonin axon terminals in the
Ferris, C.F., Messenger, T., Sullivan, R., 2005. Behavioral and neuroendocrine con- ventral tegmental area of the rat: fine structure and synaptic input to dopami-
sequences of social subjugation across adolescence and adulthood. Frontiers in nergic neurons. Brain Research 435, 71–83.
Zoology 2, 7. Hess, W.R., 1928. Stammganglien-Reizversuche. Beritche der Gesamte Physiologie
Ferris, C.F., Stolberg, T., Kulkarni, P., Murugavel, M., Blanchard, R., Blanchard, D.C., 42, 554–555.
Febo, M., Brevard, M., Simon, N.G., 2008. Imaging the neural circuitry and chem- Hrabovszky, E., Halasz, J., Meelis, W., Kruk, M.R., Liposits, Z., Haller, J.,
ical control of aggressive motivation. BMC Neuroscience 9, 111. 2005. Neurochemical characterization of hypothalamic neurons involved in
Fischer, S.G., Ricci, L.A., Melloni Jr., R.H., 2007. Repeated anabolic/androgenic steroid attack behavior: glutamatergic dominance and co-expression of thyrotropin-
exposure during adolescence alters phosphate-activated glutaminase and gluta- releasing hormone in a subset of glutamatergic neurons. Neuroscience 133,
mate receptor 1 (GluR1) subunit immunoreactivity in Hamster brain: correlation 657–666.
with offensive aggression. Behavioural Brain Research 180, 77–85. Jackson, D., Burns, R., Trksak, G., Simeone, B., DeLeon, K.R., Connor, D.F., Harrison,
Franzini, A., Marras, C., Ferroli, P., Bugiani, O., Broggi, G., 2008. Stimulation of the pos- R.J., Melloni Jr., R.H., 2005. Anterior hypothalamic vasopressin modulates the
terior hypothalamus for medically intractable impulsive and violent behavior. aggression-stimulating effects of adolescent cocaine exposure in Syrian ham-
Stereotactic and Functional Neurosurgery 83, 63–66. sters. Neuroscience 133, 635–646.
Franzini, A., Messina, G., Cordella, R., Marras, C., Broggi, G., 2010. Deep brain stim- Jasper, H.H., Marsan, C.A., 1960. A Stereotaxic Atlas of the Diencephalon of the Cat.
ulation of the posteromedial hypothalamus: indications, long-term results, and National Research Council of Canada, Toronto.
neurophysiological considerations. Neurosurgical Focus 29, E13. Joëls, M., Karst, H., Alfarez, D., Heine, V.M., Qin, Y., van Riel, E., Verkuyl, M., Lucassen,
Gobrogge, K.L., Liu, Y., Young, L.J., Wang, Z., 2009. Anterior hypothalamic vasopressin P.J., Krugers, H.J., 2004. Effects of chronic stress on structure and cell function in
regulates pair-bonding and drug-induced aggression in a monogamous rodent. rat hippocampus and hypothalamus. Stress 7, 221–231.
Proceedings of the National Academy of Sciences of the United States of America Kavushansky, A., Ben-Shachar, D., Richter-Levin, G., Klein, E., 2009. Physical stress
106, 19144–19149. differs from psychosocial stress in the pattern and time-course of behavioral
Gould, T.J., 2010. Addiction and cognition. Addiction Science & Clinical Practice 5, responses, serum corticosterone and expression of plasticity-related genes in
4–14. the rat. Stress 12, 412–425.
Grimes, J.M., Melloni Jr., R.H., 2002. Serotonin modulates offensive attack in ado- Kessler, M.S., Murgatroyd, C., Bunck, M., Czibere, L., Frank, E., Jacob, W., Horvath, C.,
lescent anabolic steroid-treated hamsters. Pharmacology Biochemistry and Muigg, P., Holsboer, F., Singewald, N., Spengler, D., Landgraf, R., 2007. Diabetes
Behavior 73, 713–721. insipidus and, partially, low anxiety-related behaviour are linked to a SNP-
Grimes, J.M., Melloni Jr., R.H., 2005. Serotonin-1B receptor activity and expres- associated vasopressin deficit in LAB mice. European Journal of Neuroscience
sion modulate the aggression-stimulating effects of adolescent anabolic steroid 26, 2857–2864.
exposure in hamsters. Behavioral Neuroscience 119, 1184–1194. Kew, J.N., Richards, J.G., Mutel, V., Kemp, J.A., 1998. Developmental changes in
Grimes, J.M., Melloni Jr., R.H., 2006. Prolonged alterations in the serotonin neural sys- NMDA receptor glycine affinity and ifenprodil sensitivity reveal three distinct
tem following the cessation of adolescent anabolic–androgenic steroid exposure populations of NMDA receptors in individual rat cortical neurons. Journal of
in hamsters (Mesocricetus auratus). Behavioral Neuroscience 120, 1242–1251. Neuroscience 18, 1935–1943.
Grimes, J.M., Ricci, L.A., Melloni Jr., R.H., 2003. Glutamic acid decarboxylase (GAD65) Kim, C., Jeon, D., Kim, Y.H., Lee, C.J., Kim, H., Shin, H.S., 2009. Deletion of N-type
immunoreactivity in brains of aggressive, adolescent anabolic steroid-treated Ca(2+) channel Ca(v)2.2 results in hyperaggressive behaviors in mice. Journal of
hamsters. Hormones and Behavior 44, 271–280. Biological Chemistry 284, 2738–2745.
Halasz, J., Liposits, Z., Kruk, M.R., Haller, J., 2002a. Neural background of glucocor- Knyshevski, I., Connor, D.F., Harrison, R.J., Ricci, L.A., Melloni Jr., R.H., 2005. Persistent
ticoid dysfunction-induced abnormal aggression in rats: involvement of fear- activation of select forebrain regions in aggressive, adolescent cocaine-treated
and stress-related structures. European Journal of Neuroscience 15, 561–569. hamsters. Behavioural Brain Research 159, 277–286.
Halasz, J., Liposits, Z., Meelis, W., Kruk, M.R., Haller, J., 2002b. Hypothalamic Koch, W., Schaaff, N., Pöpperl, G., Mulert, C., Juckel, G., Reicherzer, M., Ehmer-von
attack area-mediated activation of the forebrain in aggression. Neuroreport 13, Geiso, C., Möller, H.J., Hegerl, U., Tatsch, K., Pogarell, O., 2007. [I-123] ADAM
1267–1270. and SPECT in patients with borderline personality disorder and healthy control
Halasz, J., Toth, M., Mikics, E., Hrabovszky, E., Barsy, B., Barsvari, B., Haller, J., 2008. subjects. Journal of Psychiatry and Neuroscience 32, 234–240.
The effect of neurokinin1 receptor blockade on territorial aggression and in a Koolhaas, J.M., 1978. Hypothalamically induced intraspecific aggressive behaviour
model of violent aggression. Biological Psychiatry 63, 271–278. in the rat. Experimental Brain Research 32, 365–375.
Halasz, J., Zelena, D., Toth, M., Tulogdi, A., Mikics, E., Haller, J., 2009. Substance P neu- Kruk, M.R., 1991. Ethology and pharmacology of hypothalamic aggression in the rat.
rotransmission and violent aggression: the role of tachykinin NK(1) receptors in Neuroscience and Biobehavioral Reviews 15, 527–538.
the hypothalamic attack area. European Journal of Pharmacology 611, 35–43. Kruk, M.R., Halasz, J., Meelis, W., Haller, J., 2004. Fast positive feedback between the
Haller, J., Kruk, M.R., 2006. Normal and abnormal aggression: human disorders and adrenocortical stress response and a brain mechanism involved in aggressive
novel laboratory models. Neuroscience and Biobehavioral Reviews 30, 292–303. behavior. Behavioral Neuroscience 118, 1062–1070.
Haller, J., Abraham, I., Zelena, D., Juhász, G., Makara, G.B., Kruk, M.R., 1998. Aggressive Kruk, M.R., van der Poel, A.M., de Vos-Frerichs, T.P., 1979. The induction of aggressive
experience affects the sensitivity of neurons towards pharmacological treat- behaviour by electrical stimulation in the hypothalamus of male rats. Behaviour
ment in the hypothalamic attack area. Behavioural Pharmacology 9, 469–475. 70, 292–322.
Haller, J., Halász, J., Mikics, E., Kruk, M.R., 2004. Chronic glucocorticoid deficiency- Kuhlmann, A., Bertsch, K., Schmidinger, I., Thomann, P., Herpertz, S., 2012. Mor-
induced abnormal aggression, autonomic hypoarousal, and social deficit in rats. phometric differences in central stress-regulating structures between women
Journal of Neuroendocrinology 16, 550–557. with and without borderline personality disorder. Journal of Psychiatry and
Haller, J., Horváth, Z., Bakos, N., 2007. The effect of buspirone on normal Neuroscience 37 (5), http://dx.doi.org/10.1503/jpn.120039
and hypoarousal-driven abnormal aggression in rats. Progress in Neuro- Laitinen, L.V., 2001. Psychosurgery. Stereotactic and Functional Neurosurgery 76,
Psychopharmacology and Biological Psychiatry 31, 27–31. 239–242.
Haller, J., Toth, M., Halasz, J., 2005. The activation of raphe serotonergic neurons Lammers, J.H., Kruk, M.R., Meelis, W., van der Poel, A.M., 1988. Hypothalamic
in normal and hypoarousal-driven aggression: a double labeling study in rats. substrates for brain stimulation-induced attack, teeth-chattering and social
Behavioural Brain Research 161, 88–94. grooming in the rat. Brain Research 449, 311–327.
J. Haller / Brain Research Bulletin 93 (2013) 97–109 109

Lesch, K.P., 2011. When the serotonin transporter gene meets adversity: the contri- Schwartzer, J.J., Ricci, L.A., Melloni Jr., R.H., 2009. Interactions between the dopami-
bution of animal models to understanding epigenetic mechanisms in affective nergic and GABAergic neural systems in the lateral anterior hypothalamus of
disorders and resilience. Current Topics in Behavioral Neurosciences 7, 251–280. aggressive AAS-treated hamsters. Behavioural Brain Research 203, 15–22.
Licata, A., Taylor, S., Berman, M., Cranston, J., 1993. Effects of cocaine on human Shaikh, M.B., Brutus, M., Siegel, H.E., Siegel, A., 1984. Differential control of aggression
aggression. Pharmacology Biochemistry and Behavior 45, 549–552. by the midbrain. Experimental Neurology 83, 436–442.
Lin, D., Boyle, M.P., Dollar, P., Lee, H., Lein, E.S., Perona, P., Anderson, D.J., 2011. Func- Siegel, A., Bhatt, S., Bhatt, R., Zalcman, S.S., 2007. The neurobiological bases for
tional identification of an aggression locus in the mouse hypothalamus. Nature development of pharmacological treatments of aggressive disorders. Current
470, 221–226. Neuropharmacology 5, 135–147.
Lipp, H.P., 1978. Aggression and flight behaviour of the marmoset monkey Callithrix Siegel, A., Roeling, T.A., Gregg, T.R., Kruk, M.R., 1999. Neuropharmacology of brain-
jacchus: an ethogram for brain stimulation studies. Brain, Behavior and Evolution stimulation-evoked aggression. Neuroscience and Biobehavioral Reviews 23,
15, 241–259. 359–389.
Lotze, M., Veit, R., Anders, S., Birbaumer, N., 2007. Evidence for a different role of Smith, D.E., King, M.B., Hoebel, B.G., 1970. Lateral hypothalamic control of killing:
the ventral and dorsal medial prefrontal cortex for social reactive aggression: evidence for a cholinoceptive mechanism. Science 167, 900–901.
an interactive fMRI study. Neuroimage 34, 470–478. Su, T., Pagliaro, M., Schmidt, P.J., Pickar, D., Wolkowitz, O., Rubinow, D.R., 1993. Neu-
Lyons-Ruth, K., Choi-Kain, L., Pechtel, P., Bertha, E., Gunderson, J., 2011. Perceived rospychiatric effects of anabolic steroids in male normal volunteers. Journal of
parental protection and cortisol responses among young females with border- the American Medical Association 269, 2760–2764.
line personality disorder and controls. Psychiatry Research 189, 426–432. Sugerman, R.A., Demski, L.S., 1978. Agonistic behavior elicited by electrical stimula-
Mai, J.K., Assheuer, J., Paxinos, G., 1998. Atlas of the Human Brain. Academic Press, tion of the brain in Western collared lizards, Crotaphytus collaris. Brain, Behavior
San Diego. and Evolution 15, 446–469.
Martin, J.R., 1976. Motivated behaviors elicited from hypothalamus, midbrain, and Toth, M., Halasz, J., Mikics, E., Barsy, B., Haller, J., 2008. Early social deprivation
pons of the guinea pig (Cavia porcellus). Journal of Comparative and Physiological induces disturbed social communication and violent aggression in adulthood.
Psychology 90, 1011–1034. Behavioral Neuroscience 122, 849–854.
Melloni Jr., R.H., Ferris, C.F., 1996. Adolescent anabolic steroid use and aggressive Toth, M., Mikics, E., Tulogdi, A., Aliczki, M., Haller, J., 2011. Post-weaning social
behavior in golden hamsters. Annals of the New York Academy of Sciences 794, isolation induces abnormal forms of aggression in conjunction with increased
372–375. glucocorticoid and autonomic stress responses. Hormones and Behavior 60,
Miczek, K.A., Weerts, E.M., Tornatzky, W., DeBold, J.F., Vatne, T.M., 1992. Alcohol and 28–36.
“bursts” of aggressive behavior: ethological analysis of individual differences in Toth, M., Tulogdi, A., Biro, L., Soros, P., Mikics, E., Haller, J., 2012. The neural back-
rats. Psychopharmacology 107, 551–563. ground of hyper-emotional aggression induced by post-weaning social isolation.
Miles, D.R., Carey, G., 1997. Genetic and environmental architecture of human Behavioural Brain Research 233, 120–129.
aggression. Journal of Personality and Social Psychology 72, 207–217. Trainor, B.C., Greiwe, K.M., Nelson, R.J., 2006. Individual differences in estrogen
Moldow, R.L., Fischman, A.J., 1987. Cocaine induced secretion of ACTH, beta- receptor alpha in select brain nuclei are associated with individual differences
endorphin, and corticosterone. Peptides 8, 819–822. in aggression. Hormones and Behavior 50, 338–345.
Natarajan, D., de Boer, S.F., Koolhaas, J.M., 2009. Lack of differential serotonin biosyn- Tulogdi, A., Toth, M., Halasz, J., Mikics, E., Fuzesi, T., Haller, J., 2010. Brain mechanisms
thesis capacity in genetically selected low and high aggressive mice. Physiology involved in predatory aggression are activated in a laboratory model of violent
and Behavior 98, 411–415. intra-specific aggression. European Journal of Neuroscience 32, 1744–1753.
Nelson, R.J., Chiavegatto, S., 2001. Molecular basis of aggression. Trends in Neuro- Tulogdi A, Toth M, Barsvari B, Biro L, Mikics E, Haller J., 2012. Effects of resocializa-
sciences 24, 713–719. tion on post-weaning social isolation-induced abnormal aggression and social
Neumann, I.D., Veenema, A.H., Beiderbeck, D.I., 2010. Aggression and anxiety: social deficits in rats. Developmental Psychobiology, in press.
context and neurobiological links. Frontiers in Behavioral Neuroscience 4, 12. Veenema, A.H., Blume, A., Niederle, D., Buwalda, B., Neumann, I.D., 2006. Effects of
Olah, S., Fule, M., Komlosi, G., Varga, C., Baldi, R., Barzo, P., Tamas, G., 2009. Regula- early life stress on adult male aggression and hypothalamic vasopressin and
tion of cortical microcircuits by unitary GABA-mediated volume transmission. serotonin. European Journal of Neuroscience 24, 1711–1720.
Nature 461, 1278–1281. Veenema, A.H., Koolhaas, J.M., de Kloet, E.R., 2004. Basal and stress-induced differ-
Paxinos, G., Watson, C., 1998. The Rat Brain in Stereotaxic Coordinates. Academic ences in HPA axis, 5-HT responsiveness, and hippocampal cell proliferation in
Press, San Diego. two mouse lines. Annals of the New York Academy of Sciences 1018, 255–265.
Paxinos, G., Franklin, K.B.J., 2001. The Mouse Brain in Stereotaxic Coordinates. Aca- Veit, R., Lotze, M., Sewing, S., Missenhardt, H., Gaber, T., Birbaumer, N., 2010. Aber-
demic Press, San Diego. rant social and cerebral responding in a competitive reaction time paradigm in
Pedrosa-Sanchez, M., Sola, R.G., 2003. Modern day psychosurgery: a new criminal psychopaths. Neuroimage 49, 3365–3372.
approach to neurosurgery in psychiatric disease. Revista de Neurologia 36, Vergnes, M., Karli, P., 1969. Interspecific rat-mice agression behavior: effects of
887–897. electric stimulation of the lateral hypothalamus, amygdala and hypopocampus.
Putkonen, P.T., 1966. Attack elicited by forebrain and hypothalamic stimulation in Journal of Physiology 61 (Suppl. 2), 425.
the chicken. Experientia 22, 405–407. Virkkunen, M., 1985. Urinary free cortisol secretion in habitually violent offenders.
Ramamurthi, B., 1988. Stereotactic operation in behaviour disorders. Amygdalotomy Acta Psychiatrica Scandinavica 72, 40–44.
and hypothalamotomy. Acta Neurochirurgica: Supplement 44, 152–157. Vizi, E.S., 1984. Non-synaptic Interactions between Neurons: Modulation of Neuro-
Rejeski, W.J., Gregg, E., Kaplan, J.R., Manuck, S.B., 1990. Anabolic–androgenic chemical Transmission. Pharmacological and Clinical Aspects. John Wiley and
steroids: effects on social behavior and baseline heart rate. Health Psychology Sons, Chichester, New York.
9, 774–791. Vizi, E.S., 2000. Role of high-affinity receptors and membrane transporters in
Ricci, L.A., Grimes, J.M., Melloni Jr., R.H., 2007. Lasting changes in neuronal nonsynaptic communication and drug action in the central nervous system.
activation patterns in select forebrain regions of aggressive, adolescent Pharmacological Reviews 52, 63–89.
anabolic/androgenic steroid-treated hamsters. Behavioural Brain Research 176, Vizi, E.S., Fekete, A., Karoly, R., Mike, A., 2010. Non-synaptic receptors and trans-
344–352. porters involved in brain functions and targets of drug treatment. British Journal
Ricci, L.A., Grimes, J.M., Knyshevski, I., Melloni Jr., R.H., 2005. Repeated cocaine of Pharmacology 160, 785–809.
exposure during adolescence alters glutamic acid decarboxylase-65 (GAD65) Wallinga, A.E., ten Voorde, A.M., de Boer, S.F., Koolhaas, J.M., Buwalda, B., 2009.
immunoreactivity in hamster brain: correlation with offensive aggression. Brain MDMA-induced serotonergic neurotoxicity enhances aggressiveness in low- but
Research 1035, 131–138. not high-aggressive rats. European Journal of Pharmacology 618, 22–27.
Ricci, L.A., Rasakham, K., Grimes, J.M., Melloni Jr., R.H., 2006. Serotonin-1A receptor Weerts, E.M., Miller, L.G., Hood, K.E., Miczek, K.A., 1992. Increased GABAA-dependent
activity and expression modulate adolescent anabolic/androgenic steroid- chloride uptake in mice selectively bred for low aggressive behavior. Psy-
induced aggression in hamsters. Pharmacology Biochemistry and Behavior 85, chopharmacology 108, 196–204.
1–11. Weinberg, A., Klonsky, E.D., Hajcak, G., 2009. Autonomic impairment in border-
Ricci, L.A., Schwartzer, J.J., Melloni Jr., R.H., 2009. Alterations in the anterior hypo- line personality disorder: a laboratory investigation. Brain and Cognition 71,
thalamic dopamine system in aggressive adolescent AAS-treated hamsters. 279–286.
Hormones and Behavior 55, 348–355. Weissenberger, A.A., Dell, M.L., Liow, K., Theodore, W., Frattali, C.M., Hernandez,
Roberts, W.W., Steinberg, M.L., Means, L.W., 1967. Hypothalamic mechanisms for D., Zametkin, A.J., 2001. Aggression and psychiatric comorbidity in children
sexual, aggressive, and other motivational behaviors in the opossium, Didelphis with hypothalamic hamartomas and their unaffected siblings. Journal of the
virginiana. Journal of Comparative and Physiological Psychology 64, 1–15. American Academy of Child and Adolescent Psychiatry 40, 696–703.
Roeling, T.A., Kruk, M.R., Schuurmans, R., Veening, J.G., 1993. Behavioural responses Widom, C.S., 1989. Child abuse, neglect, and adult behavior: research design and
of bicucculline methiodide injections into the ventral hypothalamus of freely findings on criminality, violence, and child abuse. American Journal of Orthopsy-
moving, socially interacting rats. Brain Research 615, 121–127. chiatry 59, 355–367.
Romaniuk, A., 1965. Representation of aggression and flight reactions in the hypo- Woodman, D., Hinton, J., 1978. Catecholamine balance during stress anticipation: an
thalamus of the cat. Acta Biologiae Experimentalis 25, 177–186. abnormality in maximum security hospital patients. Journal of Psychosomatic
Sano, K., Mayanagi, Y., Sekino, H., Ogashiwa, M., Ishijima, B., 1970. Results of Research 22, 477–483.
stimulation and destruction of the posterior hypothalamus in man. Journal of Woodworth, C.H., 1971. Attack elicited in rats by electrical stimulation of the lateral
Neurosurgery 33, 689–707. hypothalamus. Physiology and Behavior 6, 345–353.

You might also like