Download as pdf or txt
Download as pdf or txt
You are on page 1of 41

Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371

Contents lists available at ScienceDirect

Neuroscience and Biobehavioral Reviews


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

Review

The neurobiology of depression and antidepressant action


Paul Willner a,∗ , Jørgen Scheel-Krüger b , Catherine Belzung c
a
Department of Psychology, Swansea University, Singleton Park, Swansea SA2 8PP, UK
b
Center of Functionally Integrative Neuroscience, University of Aarhus, Denmark
c
INSERM 930 and University Francois-Rabelais, Tours, France

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

Article history: We present a comprehensive overview of the neurobiology of unipolar major depression and antide-
Received 24 April 2012 pressant drug action, integrating data from affective neuroscience, neuro- and psychopharmacology,
Received in revised form neuroendocrinology, neuroanatomy, and molecular biology. We suggest that the problem of depression
26 November 2012
comprises three sub-problems: first episodes in people with low vulnerability (‘simple’ depressions),
Accepted 10 December 2012
which are strongly stress-dependent; an increase in vulnerability and autonomy from stress that devel-
ops over episodes of depression (kindling); and factors that confer vulnerability to a first episode (a
Keywords:
depressive diathesis). We describe key processes in the onset of a ‘simple’ depression and show that
Unipolar major depression
Antidepressant drugs
kindling and depressive diatheses reproduce many of the neurobiological features of depression. We
Affective neuroscience also review the neurobiological mechanisms of antidepressant drug action, and show that resistance to
Hippocampus antidepressant treatment is associated with genetic and other factors that are largely similar to those
Amygdala implicated in vulnerability to depression. We discuss the implications of these conclusions for the under-
Ventromedial prefrontal cortex standing and treatment of depression, and make some strategic recommendations for future research.
Anterior cingulate cortex
Nucleus accumbens © 2012 Elsevier Ltd. All rights reserved.
Caudate nucleus
Habenula
Stress
Treatment resistance

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2332
2. Different mechanisms for depression and antidepressant action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2333
3. The psychobiology of depression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2334
3.1. The diathesis/stress model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2334
3.1.1. Diathesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2334
3.1.2. Stress. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2335
3.1.3. Implications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2335
3.2. Effects of stress on the HPA axis and hippocampus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2335
3.2.1. Neurotoxic effects of stress . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2336
3.2.2. Morphological consequences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2337
3.3. Frontal brain circuitry underlying the symptoms of depression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2337
3.4. What drives the changes? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2338
3.4.1. Anhedonia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2339
3.5. Mechanisms mediating affective information-processing biases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2339
3.5.1. Amygdala and nucleus accumbens . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2339
3.5.2. The habenula . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2340
3.6. Failure to cope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2340
3.6.1. Stimulus appraisal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2340
3.6.2. Learned helplessness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2341

∗ Corresponding author.
E-mail address: p.willner@swansea.ac.uk (P. Willner).

0149-7634/$ – see front matter © 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.neubiorev.2012.12.007
2332 P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371

3.7. Rumination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2341


3.7.1. Rumination and recovery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2342
3.8. Basal ganglia involvement in coping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2342
3.9. Summary: stress and depression in conditions of low vulnerability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2343
3.10. Vulnerability to depression: 1. Kindling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2343
3.11. Vulnerability to depression: 2. Predisposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2344
3.11.1. Animal models of vulnerability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2345
3.11.2. The vulnerability brain and the depressed brain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2345
4. Mechanisms of antidepressant action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2345
4.1. Potentiation of monoamine transmission . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2345
4.2. Post-transductional mechanisms of antidepressant action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2346
4.3. Neurogenesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2348
4.3.1. The pathway from synapse to neurogenesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2348
4.3.2. Functional significance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2349
4.4. HPA axis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2349
4.5. Why are antidepressants slow to act? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2350
4.6. Treatment implications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2350
4.7. Treatment resistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2352
4.7.1. Pharmacokinetics, misdiagnosis and comorbidity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2352
4.7.2. Parallels with vulnerability to depression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2353
4.8. Approaches to the treatment of resistant depression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2354
4.8.1. New targets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2355
5. Conclusions and research implications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2356
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2358

1. Introduction an absolute or relative deficiency of monoamines at functionally


important receptor sites in the brain”, with the corollary that
Depression is the commonest psychiatric disorder. It is the most antidepressants work by correcting these deficiencies) has pro-
disabling medical condition, in terms of years lost to disability, and vided the major neurobiological account of depression. Indeed,
it is projected that by 2030 depression will be the foremost con- until recently, it was the only significant hypothesis, and while its
tributor to the worldwide burden of disease (WHO, 2008). In this predominance has been to some extent eclipsed by newer con-
review, we focus on unipolar major depressive disorder, which is cepts over the past decade, it remains the case that the monoamine
defined in DSM-IV (American Psychiatric Association, 1994), as a hypothesis has provided the only significant theoretical frame-
condition characterized by the presence of loss of pleasure or inter- work for antidepressant drug development, proving stubbornly
est in usually pleasurable activities (anhedonia), together with an resistant to the numerous and very expensive attempts by the phar-
array of other features, including anergia, changes in sleep and maceutical industry to break out of the monoamine straitjacket
appetite, sadness, and suicidal ideation. Presentations of unipo- with drugs that act through novel mechanisms. As summarized
lar major depressive disorder (which we shall refer to as simply in Fig. 1, newer antidepressants differ from the older drugs in
‘depression’) can be very variable, but this fact has not featured decreasing the incidence of unwanted side effects and/or narrow-
prominently in the literature that we shall review. ing the neurochemical target, rather than by introducing novel
Depression is characterized by a profoundly negative view of mechanisms of action. However, improvements in both antide-
the world, oneself and the future (Beck, 1967), and this nega- pressant response rates and the slow onset of clinical effect,
tive world-view has been related to negative biases in attention, requiring several weeks of chronic treatment to achieve the full
interpretation and memory (Mathews and MacLeod, 2005). Specif- effect, have been minimal. Tolerability has improved, but dif-
ically, studies of cognitive processing in depression have reported ferences in efficacy are small and difficult to demonstrate, and
increased elaboration of negative information, difficulties disen- there is little evidence that the newer antidepressants are more
gaging from negative material, and deficits in cognitive control efficacious than the older antidepressants. Indeed, one of the old-
when processing negative information, which inter alia explain est antidepressants, the tricyclic clomipramine, remains among
why depressed people experience a high level of negative auto- the most efficacious, alongside the serotonin-noradrenaline reup-
matic thoughts and pathological rumination (Gotlib and Joormann, take inhibitor (SNRI) venlafaxine, the selective serotonin reuptake
2010). Depressed people are particularly vulnerable to negative inhibitors (SSRIs) sertraline and escitalopram, and the atypical
psychological feedback, which has a disproportionately disruptive antidepressant mirtazepine (Montgomery et al., 2007; Cipriani
effect on subsequent performance (Elliott et al., 1996). In addition to et al., 2009). Antidepressants have consistently shown only moder-
an increased response to aversive events, depression is also char- ate response rates, with around 30–40% of patients being classified
acterized by a decreased response to anticipated (McFarland and as non-responders, and the latency of clinical onset remains stub-
Klein, 2008) or actual (Pizzagalli et al., 2008; Chase et al., 2010) bornly long (Trivedi et al., 2006; Holtzheimer and Mayberg, 2011).
rewards, and this provides a cognitive explanation of the core While antidepressant efficacy has been claimed for a number of
symptom of depression, anhedonia. These two complementary non-monoaminergic drugs that are marketed for other indications,
biases, increased negativity and decreased positivity, are central and the failure of some novel agents may to some extent involve
to much of the recent neurobiological literature on depression, increased regulatory requirements, the relative lack of progress
because they play directly into two of the major experimental over the past 50 years is remarkable (Blier, 2010; Baghai et al.,
methodologies, functional neuroimaging and animal models of 2011).
depression. In this paper we present a comprehensive overview of the neu-
Since its introduction almost 50 years ago, the monoamine robiology of unipolar major depression and antidepressant drug
hypothesis (“some, if not all, depressions are associated with action, integrating data from affective neuroscience, neuro- and
P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371 2333

Fig. 1. Antidepressant drug development: new wine in old bottles. The figure summarizes the ways in which the five major newer classes of antidepressant drugs represent
refinements of the mechanisms introduced in the original drug classes, the tricyclic antidepressants and monoamine oxidase inhibitors.

psychopharmacology, neuroendocrinology, neuroanatomy, and in the neural bases of depression and antidepressant action, and
molecular biology, and from preclinical and clinical research. In so the action of antidepressants cannot accurately be described as
doing, we develop a framework for understanding the neurobiology reversing and normalizing the processes that are dysfunctional in
of depression, which also provides a basis for understanding the the depressed brain.
limited success of research in antidepressant drug development. The clearest evidence that the brain of an antidepressant-treated
In Section 3, we first provide a detailed account of a ‘basic’ psy- patient is not in a normal state comes from studies in which
chobiology of depression, which centres on the effects of stress on antidepressant effects are blocked by acute drug treatments. An
neurobiological and psychological functioning in individuals who example is shown in Fig. 2. In this study, after depressed patients
have a low predisposition to become depressed. We next consider had been successfully treated with SSRIs, they were administered a
the mechanisms that underlie various vulnerabilities to depression, low dose of the dopamine (DA) D2 receptor blocker sulpiride. This
and review evidence that these reflect changes in brain function caused a profound return of depressed mood in the treated patients,
that resemble effects of stress, with the result that depression is
more easily precipitated and less stress-dependent. In Section 4,
we review recent research on the mechanisms of antidepressant
action which demonstrates that antidepressants essentially coun- Depressed mood
teract and repair the effects of stress. We also show that the factors 6.0
implicated in resistance to antidepressant treatment largely reca-
pitulate the factors involved in vulnerability to depression, and 5.0
argue that antidepressants are ineffective under these conditions
VAS rating

because stress is of minor importance. In Section 5, we discuss the 4.0


implications of these conclusions for the understanding and treat-
ment of depression, and make some strategic recommendations for 3.0
future research. But first (Section 2), we explain why it is necessary
Con-Plac
to give separate consideration to the analyses of depression and 2.0
Con-Sul
antidepressant action.
Dep-Plac
1.0
Dep-Sul
2. Different mechanisms for depression and antidepressant
action 0.0
0 1 2 3 4 5 6 7
A feature of the monoamine hypothesis of depression that has Time (h)
gone largely unremarked is that it proposes a single mechanism for
both depression and antidepressant drugs: depression results from Fig. 2. A demonstration that the antidepressant-treated brain is not in a ‘normal’
state. Volunteers (Con) or depressed patients who had recovered following SSRI
a decreased functioning in NA and/or 5HT which antidepressants
treatment (Dep) were administered acutely either placebo (Plac) or a low dose of
increase back to normal. The same symmetry is seen in most of the the DA D2 receptor blocker sulpiride (Sul). Sulpiride caused a return of severely
more recent hypotheses that will be discussed below. However, the depressed mood in the patients but not in controls.
assumption of symmetry is incorrect. There are many differences Adapted from Willner et al. (2005).
2334 P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371

but if anything, slightly improved mood in a non-depressed non- 2011) or childhood abuse (Kendler et al., 2002, 2006; Widom et al.,
antidepressant-treated control group (Willner et al., 2005). Exactly 2007). Therefore, interactions between multiple risk genes and
the same effect is seen in studies of serotonin (5HT) or noradren- early environment seem to explain a large part of the variability
aline (NA) depletion. Consumption of a mixture of dietary amino (Krishnan and Nestler, 2008; Caspi et al., 2010), and much of the
acids that omits tryptophan, the precursor of 5HT, leads rapidly to diathesis for depression is laid down in early childhood.
a 70–80% decrease in plasma levels of tryptophan, and a consequent The mechanisms by which these early experiences increase the
decrease in brain 5HT levels; and the tyrosine hydroxylase inhibitor risk of depression (and other psychiatric disorders) include not only
alpha-methyl-para-tyrosine (AMPT) causes a selective depletion biological processes (discussed below) but also psychological and
of the catecholamines DA and NA. Tryptophan depletion causes a psychosocial constructs that convert transient traumatic experi-
severe reinstatement of symptoms in the majority of patients in ences into long-term vulnerabilities. For example, loss of a parent
remission following successful treatment of depression with SSRIs or a poor quality of parental care leads to low self-esteem and emo-
(Delgado et al., 1990, 1999). Similarly, AMPT precipitates a similar tional instability (e.g. Akiskal, 1984; Avagianou and Zafiropoulou,
relapse in depressed patients successfully treated with NA reuptake 2008; Parsons et al., 2010), and may decrease the ability to form
inhibitors (NRIs) (Delgado et al., 1993; Miller et al., 1996). However, close relationships, and so dilute the quality of social support avail-
neither of these manipulations has major effects on mood in people able in later life (e.g. Schoenfelder et al., 2011). There is evidence
who are not depressed or at high-risk for depression (Ruhé et al., that distinct but substantially overlapping neural networks sub-
2007). serve depression and insecure attachment (Galynker et al., 2011).
Similar findings have been reported in animal studies. Both Even high levels of emotions after romantic love and sexual activi-
5HT depletion and NA depletion block the action of SSRIs and ties may increase the risk for later depression in young adolescent
NRIs, respectively, in animal models of depression. However, as girls, as a result of immature and inefficient social coping after fail-
in the clinical situation, while reversing the effects of antidepres- ures at this young age (Davila et al., 2009). Early life experiences
sants, neither manipulation produces a depressive phenotype in also determine characteristic styles of processing information in
untreated animals (Lucki and O’Leary, 2004; Cryan et al., 2004; relation to the self. For example, the negative thinking that charac-
Yalcin et al., 2008). Chronic mild stress (CMS) is a well-validated terizes the depressed person is thought to reflect the activation of
animal model of depression, based on the loss of responsiveness a negative ‘cognitive schema’, learned through adverse childhood
to rewards by animals subjected to a varying schedule of minor experiences such as rejection, criticism, or living with a depressed
stressor (Willner et al., 1987; Willner, 1997a). In a large scale parent (Beck, 1967). Although depressive cognitions are to a large
transcriptomic study in mice, CMS induced a molecular shift in cor- extent state-dependent and their role as vulnerabilities has been
ticolimbic brain areas that was only partly reversed by fluoxetine: questioned, there is now good evidence for the existence of cog-
for example, fluoxetine failed to block over a quarter of the changes nitive diatheses for depression (Alloy et al., 1999; Mathews and
in the amygdala and cingulate cortex, and almost three quarters of MacLeod, 2005; Roiser et al., 2012), which may only be apparent
the changes in the hippocampus, even though the behaviour of the when a depressive schema is activated: for example, girls at risk
mice had returned to normal after treatment (Surget et al., 2009). for depression because their mothers were recurrently depressed
It is evident that antidepressants do not normalize brain attended selectively to negative facial expressions when subjected
activity: mood and behaviour are restored to normal, but the to a depressive mood induction procedure, but not otherwise
antidepressant-treated brain is in a different state from the non- (Joormann et al., 2007).
depressed brain. This is also evidenced by the high rate of relapse if Personality factors, which are moderately heritable, but also
antidepressants are too-rapidly withdrawn following remission of reflect early experience, interact with both cognitive and social
symptoms (Kaymaz et al., 2008). Therefore it is necessary to exam- factors in the etiology of depression (Compass et al., 2004). Much
ine separately the neurobiology of depression and the mechanisms of the influence of both genetics and early traumatic events on
of antidepressant action, rather than conflating these two issues. chronic depressive symptomatology is mediated through the per-
sonality factor of neuroticism (Kendler and Gardner, 2011), which
is one of the strongest risk factors for depression (Enns and Cox,
3. The psychobiology of depression
1997; Christensen and Kessing, 2006). Early-onset depression in
particular (first episode before the age of 30), is characterized by
3.1. The diathesis/stress model
a higher level of neuroticism and a higher prevalence of comorbid
personality disorders, but lower exposure to stressful life events
Individuals within the population vary greatly in their vulnera-
prior to onset, relative to individuals experiencing a later first
bility to psychiatric disorders, including depression. This variation
episode (Bukh et al., 2011). Like depression, a high level of neuroti-
is usually understood within a diathesis/stress model that considers
cism is also associated with a negative information-processing bias
separately issues of vulnerability (the diathesis) and precipitation
(Chan et al., 2007). High levels of negative emotionality (a construct
(the stress) (Monroe and Simons, 1991). Two features of this model
closely related to neuroticism) has been shown to lead, in young
are critical for our present analysis: as the diathesis increases,
people, to the formation of dysfunctional attitudes and other cog-
the level of stress needed to precipitate an episode of depression
nitive vulnerabilities (Lakdawalla and Hankin, 2008; Joiner et al.,
decreases, and the occurrence of an episode of depression itself
2005), while low levels of positive emotionality (pleasure capac-
increases the diathesis for future episodes.
ity) increase vulnerability to depression by reducing social support
(Wetter and Hankin, 2009). Neurotic individuals at high genetic
3.1.1. Diathesis risk for major depression also have an elevated tendency to inter-
We consider first the nature of the depressive diathesis. A pre- act with others in ways that generate stress, and they bear some
disposition (or diathesis) to become depressed may arise in a responsibility for the stressful life events that they encounter
variety of ways, and at different stages of the life cycle. For example, (Hammen, 2006; Kendler et al., 2006).
a number of genetic diatheses have been identified, with a heri- Personality factors and in particular, the personality styles
tability ranging from 31 to 42% (Sullivan et al., 2000; Kendler et al., labelled ‘sociotropy’ and ‘autonomy’, may also account for much
2002, 2006). Some early life experiences are also known to increase of the variability in the symptomatology of depression, includ-
the risk for depression, particularly inadequate emotional contact ing the clinical presentations of autonomous (or ‘endogenous’)
with parents (Robertson and Bowlby, 1952; Roy, 1981; Slavich et al., and reactive depression. Autonomous people obtain pleasure from
P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371 2335

‘doing’ and reaching goals. They have autonomous (‘endogenous’) a pre-existing chronic minor depression (Fogel et al., 2006). How-
depressions characterized by interpersonal distance and hostility, ever, there are many studies demonstrating that the association
hopelessness/suicidality, feelings of failure, and anhedonia: having between severe stress and depression decreases with each succes-
failed in their own terms they lose their motivation, and blaming sive episode, since minor stress episodes may by now be a sufficient
themselves for their failure, they minimize the importance of envi- trigger (Dienes et al., 2006; Stroud et al., 2008; Morris et al., 2010;
ronmental precipitants. Socially dependent people rely for their Slavich et al., 2011). For example, a study of over 2000 women
satisfaction on the approval of others: their depressions are char- reported that over the first six episodes of depression the likeli-
acterized by interpersonal sensitivity, guilt and self-blame, anxiety hood of an episode occurring in any given month increased 15-fold,
and rumination. Their depressions are ‘reactive’: they will show but at the same time, the association with a stressful life event
temporary mood improvements if they receive a strong input of decreased by 75% (Kendler et al., 2000). So over the course of a
attention and reassurance; and because relationships with the lifetime of depressive episodes, the onset of depression becomes
outside world are central to them, the events that precipitate increasingly autonomous. This effect can be described both neu-
depression are much more obvious, and may be exaggerated in robiologically: an episode of depression sensitizes, or ‘kindles’, the
the attempt to regain love or attention (Beck, 1983; Willner, 1985; brain to respond to weaker and weaker precipitants (Post, 1992;
Robins et al., 1997). Contrary to a widespread belief, ‘endogenous’ Stroud et al., 2011), and psychologically: the depressed person
and ‘reactive’ depressions are distinguished by their presentation relies increasingly on negative modes of information processing
rather than by the presence or absence of precipitants: indeed, that come to be activated by increasingly minimal cues (Segal et al.,
endogenous depression is more stress-sensitive, not less (Harkness 1996; Monroe et al., 2007). (These two accounts from different
and Monroe, 2006). stand-points are entirely compatible with one another.)
Thus, the requirement for strong and identifiable stressors to
3.1.2. Stress precipitate an episode of depression decreases as the depressive
A first episode of depression manifests against the background diathesis increases (e.g. several previous episodes of depression and
of a level of vulnerability that is determined by these genetic and a high genetic risk). However, in a person with a weak depressive
experiential factors. The precipitant (or stressor) for a depressive diathesis (e.g. no history of depression and a low genetic risk) it
episode could be an internal event, such as a hormonal challenge is extremely likely that an episode of depression has been precipi-
or a traumatic head injury. More commonly, the onset of depres- tated by stress.
sion is precipitated by external events. The likelihood of entering a
depressive episode is greatly increased following a major adverse 3.1.3. Implications
life event such as bereavement (recent evidence suggests that This analysis suggests the heuristic strategy that we have
interpersonal loss events are particularly powerful precipitants of adopted in the remainder of this review. We propose that a full
depression: Farmer and McGuffin, 2003; Slavich et al., 2010, 2011), understanding of the neurobiology of depression implies a solution
or in relation to an accumulation of chronic minor stresses such to three separate problems: (1) the mechanisms by which stress
as unemployment, poverty, family disharmony or living with sev- precipitates first episodes of depression in individuals with a weak
eral pre-school children (Brown and Harris, 1978; Harkness and depressive diathesis; (2) the kindling process by which depres-
Monroe, 2006). These two types of precipitating events respectively sion becomes increasingly autonomous of stress over repeated
form the basis of two of the major animal models of depression, episodes; and (3) the psychological and neurobiological ‘prekin-
learned helplessness (Seligman, 1975) and CMS (Willner, 1997a). dling’ processes by which some individuals acquire a strong
Bereavement has traditionally precluded a diagnosis of depres- depressive diathesis even prior to their first episode of depression.
sion, but this exclusion has been questioned, on the grounds that We also propose that the first of these problems could be consid-
bereavement related depression is in all respects similar to other ered the more fundamental, because a solution to it would provide
depressive episodes (Corruble et al., 2011), and it is likely to be a basic model within which to develop solutions to the other two
reversed in DSM-V (Zisook et al., 2012). problems: for example, reports of genetic polymorphisms that con-
While the contribution of early life stress to the etiology of the fer vulnerability to depression are impossible to interpret without
depression is well established, the role of stress during adulthood a model in which to understand what role is played by the cells
has been debated, because it appears that many instances of depres- in which the gene product is expressed. In the following sections
sions are not precipitated by stress. However, there is a simple (Sections 3.2–3.9) we first outline the mechanisms that underly
resolution of this controversy. It follows from the diathesis/stress stress-induced first episodes of depression: we are concerned here
concept that a person who has a weak depressive diathesis would with understanding the impact of stress on particularly vulnera-
only succumb to an intense stress, whereas a person with a strong ble brain regions, elucidating the neural circuitry that is impacted
depressive diathesis may succumb to minor or trivial stresses. For by these effects, and relating these neurobiological changes to the
example, in a large twin study, the odds ration for an association cognitive and behavioural features of depression. We subsequently
between severely stressful life events and the onset of a first episode consider the ‘kindling’ (Section 3.10) and ‘prekindling’ (Section
of depression was four times greater in women with the lowest 3.11) mechanisms that increase vulnerability to depression and
level of genetic risk (dizygotic twins whose co-twin had no history decrease the role of stress.
of depression) than in women with the highest level of genetic risk
(monozygotic twins whose co-twin did have a history of depres- 3.2. Effects of stress on the HPA axis and hippocampus
sion) (Kendler et al., 2001). It follows that in a person with a strong
diathesis a precipitating minor stressor (which would not be identi- The major physiological response to stress is an activation
fied as a ‘life event’) might be easily overlooked, whereas a person of neuroendocrine systems, most notably, the hypothalamus-
with a weak diathesis would only succumb to a severe stressor pituitary-adrenal (HPA) axis. In this system, corticotrophin
(Monroe and Harkness, 2005; Harkness and Monroe, 2006). releasing factor (or hormone: CRF/CRH) is released from the para-
Furthermore, the relationship between stress and depres- ventricular nucleus of the hypothalamus to stimulate the pituitary
sion changes over time. Depression is a recurrent condition, in gland to produce adrenocorticotrophic hormone (ACTH), which in
which each episode increases the probability of a further episode turn stimulates the release of glucocorticoids (cortisol in humans
(Solomon et al., 2000; American Psychiatric Association, 1994). or corticosterone in rodents) from the adrenal cortex into the blood
Indeed, one of the most powerful risk factors for major depression is circulation, which inter alia exert negative feedback effects on the
2336 P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371

into the cell with a consequent loss of energy capacity, and a


dlPFC decreased production of brain-derived neurotrophic factor (BDNF),
vmPFC which provides trophic support to cell structure and function.
Through a combination of these effects, prolonged exposure to
HPC
Motor stress (Magarinos et al., 1996) or high levels of glucocorticoids
CN
Sensory outputs (Woolley et al., 1990) causes atrophy of apical dendrites, and ulti-
inputs NAcc
mately, granular cell death (Sapolsky, 2000). Exposure to CMS is
sufficient to cause loss of granule cells (Jayatissa et al., 2008, 2010).
AMG
Consistent with these structural changes, prolonged exposure to
PVN
stress or glucocorticoids impairs hippocampal-dependent mem-
RN
ory processes: spatial memory in animals (Bodnoff et al., 1995;
SN
de Quervain et al., 1998) and verbal declarative memory in peo-
PIT VTA
ple (Newcomer et al., 1999; de Quervain et al., 2000). The effect
of stress to decrease BDNF levels is complemented by a report
Ad Co
of decreased BDNF in the hippocampus of suicide victims (Karege
lHb et al., 2005). However the extreme case of hippocampal cell death
has not been observed in post-mortem samples from depressed or
Fig. 3. A schematic outline of the affective information-processing brain circuitry
steroid-treated patients (Swaab et al., 2005).
described in the text. The green arrows describe a basic input–output pathway via
the amygdala (AMG) and hippocampus (HPC) on the input side and the nucleus The hippocampal inhibitory control of the HPA axis is exerted
accumbens (NAcc), substantia nigra (SN) and caudate nucleus (CN) on the output via several multisynaptic pathways projecting from the subiculum
side. The blue arrows represent the ascending modulation of those structures, as well to the paraventricular nucleus of the hypothalamus. The subicu-
as the ventromedial (vm) and dorsolateral (dl) prefrontal cortex (PFC), by the 5HT lum sends glutamatergic projections to the bed nucleus of the stria
projections of the raphe nuclei (RN) and the DA projections of the ventral tegmen-
tal area (VTA), which are themselves modulated by the lateral habenula (lHb). The
terminalis (Jalabert et al., 2009), the lateral septum, and differ-
red arrows represent descending pathways from the PFC: this part of the diagram ent hypothalamic nuclei, all of which send GABAeric projections
is greatly simplified; in particular, it subsumes the anterior cingulate cortex within to the paraventricular nucleus (Herman et al., 1995, 2002, 2005;
the PFC. The purple arrows show the HPA axis: the paraventricular nucleus of the Cullinan et al., 2008; Belzung and Billette de Villemeur, 2010). All
hypothalamus (PVN) releases CRF, which stimulates the anterior pituitary (PIT) to
of these influences are inactivated by chronic stress: changes in the
release ACTH, which stimulates the adrenal cortex (AdCo) to release glucocorticoids
that feed back at all levels: the dashed lines show inhibitory and negative feed- neuronal activity of all these structures secondary to glucocorti-
back elements, and the full lines show stimulatory elements, including the positive coid stimulation of hippocampal neurons were absent in animals
feedback to the AMG. subjected to CMS, and were restored by chronic antidepressant
treatment (Surget et al., 2011).
pituitary and hypothalamus that limit the degree of activation of High circulating levels of glucocorticoids also promote the
the HPA axis (de Kloet et al., 2005; Holsboer and Ising, 2010). Emo- release of pro-inflammatory cytokines from macrophages and
tional stimuli reach the HPA axis via the amygdala and descending microglial cells, which contribute to the desensitization of GR
pathways from the forebrain (Fig. 3: purple arrows). The amyg- receptors (Raison et al., 2006) and further stimulate the HPA axis
dala exerts excitatory control over the hypothalamus to stimulate (Zunszain et al., 2011). There has been much discussion of the
the HPA axis, which, via increased cortisol levels, acts in a positive parallels between the anhedonic ‘sickness behaviour’ elicited by
feedback manner to further stimulate the amygdala (Duvarci and cytokines and depression (Weisler-Frank et al., 2005; Sharpley and
Pare, 2007). Conversely, the hippocampus exerts inhibitory control Agnew, 2011), though the value of this parallel may be limited
over the HPA axis, such that cortisol stimulation of the hippocam- (Dunn et al., 2005). Elevated levels of cortisol also cause an increase
pus acts in a negative feedback manner to inhibit the HPA axis. in the activity of MAO-A, with a consequent decrease in brain lev-
This negative feedback mechanism is crucial for limiting the activ- els of NA and 5HT (Slotkin et al., 1998). A neuroimaging study
ity of the HPA axis, since without it, the positive feedback loop found increased MAO-A activity throughout the brain of depressed
via the amygdala would cause the system to run out of control patients (Meyer et al., 2006; Meyer, 2012), which would explain the
(Jacobson and Sapolsky, 1991; Herman et al., 1996; Swaab et al., decreased levels of NA and 5HT posited by the monoamine hypoth-
2005; Belzung and Billette de Villemeur, 2010). Descending neg- esis of depression, albeit that direct evidence for the existence of
ative feedback over the HPA axis is also exerted at the level of such decreases remains elusive.
the dorsomedial prefrontal cortex (PFC) and the prelimbic cortex A further effect of stress that has been much discussed in the
(Ulrich-Lai and Herman, 2009; Jankord and Herman, 2008) so acti- recent literature is on neurogenesis, the growth and differenti-
vation of these areas by emotional self-regulation can also improve ation of new cells, which in the adult brain is confined to two
the control over the HPA axis. (This issue is discussed further in areas, the subventricular zone, from where new-born cells migrate
Section 3.6.) into the olfactory bulb, and the dentate gyrus of the hippocam-
pus. Despite their relatively small numbers, there is evidence that
3.2.1. Neurotoxic effects of stress newly generated neurons can affect the functioning of hippocam-
Chronic exposure to glucocorticoids is neurotoxic, and pre- pal circuits: for example, in rodents tested in spatial learning tasks
clinical studies have shown that hippocampal granule cells are known to recruit the hippocampus, levels of the immediate early
particularly sensitive to these effects. Prolonged glucocorticoid gene c-Fos are higher in newly generated adult neurons than in
exposure leads initially to a loss of glucocorticoid receptors (GR) older neurons (Kee et al., 2007). The effect of these cells is ampli-
in hippocampal granule cells, with a consequent disinhibition of fied by modulation of GABAergic interneurons: ablation of new
the HPA axis and a further increase in corticosteroid stimulation neurons caused an increase in the amplitude of spontaneous ␥-
(Raison and Miller, 2003). If further prolonged, the consequences frequency bursts in the dentate gyrus of the hippocampus, as well
of the loss of GR-mediated effects for granule cell function, as as increased synchronization of dentate neuron firing (Lacefield
well as hyper-stimulation through the activation of other, pri- et al., 2012). It has frequently been speculated that the role of
marily glutamatergic mechanisms become severe: they include a neurogenesis in antidepressant action is to support new learn-
hyperactivation of calcium-dependent enzymes leading to the pro- ing of more adaptive cognitions and behaviours (e.g. Castrén and
duction of neurotoxic free radicals, a decrease in glucose transport Rantamäki, 2010). This hypothesis receives indirect support from
P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371 2337

studies showing that suppression of neurogenesis impairs the illness (Sheline et al., 2003; Colla et al., 2007; Tata and Anderson,
learning of a hippocampal-dependent task, context discrimination 2010). Hippocampal volume has been observed to decline during
(Deng et al., 2010; Tronel et al., 2010, 2012). More direct evidence a 3-year prospective study among severely depressed in-patients,
comes from a study in which a genetic mutation was used to pre- particularly those with a incomplete remission and relapses (Frodl
vent the programmed death of newly born neurons, leading to et al., 2008a), and in one meta-analysis was only significant in
an increase in their number, accompanied by an enhancement of patients who had been depressed for more than 2 years or had more
context-discrimination learning (Sahay et al., 2011). Human neuro- than one episode of depression (McKinnon et al., 2009). The weight
genesis continues throughout the lifespan (Eriksson et al., 1998) but of this evidence is that the experience of depression increases the
decreases as a function of chronological age (Amrein et al., 2011), susceptibility of the hippocampus, and after recovery from depres-
and is therefore at a low level in adults. Nevertheless, the number of sion, patients with a history of recurrent depression continue to
newly generated cells in the human hippocampus, while small, may show a decreased hippocampal volume (Neumeister et al., 2005),
be sufficient to be of functional significance (Snyder and Cameron, consistent with the observation that cognitive impairments also
2012). persist following recovery from a depressive episode (Bhalla et al.,
Hippocampal neurogenesis is powerfully suppressed by stress 2006; Preiss et al., 2009).
or by prolonged corticosterone exposure (Wong and Herbert, 2004; In short, neurochemical and structural changes observed in
Mineur et al., 2007; Samuels and Hen, 2011; Hanson et al., 2011; the hippocampus of depressed patients closely resemble those
Petrik et al., 2012). Such effects have been reported in different observed following prolonged exposure to stress or high levels of
species, including tree shrews (Gould et al., 1997) and higher pri- circulating glucocorticoids, and the effects of stress leave a ‘scar’ on
mates (Gould et al., 1998; Perera et al., 2011) as well as rodents the hippocampus that increases the diathesis for depression.
(Mineur et al., 2007; Alonso et al., 2004), and are seen follow-
ing both psychosocial (Gould et al., 1997; Czéh et al., 2002) and 3.3. Frontal brain circuitry underlying the symptoms of
physical stressors (Malberg and Duman, 2003; Pham et al., 2003; depression
Vollmayr et al., 2003). Suppression of neurogenesis has also been
reported in post-mortem samples from elderly depressed patients While the hippocampus is the brain area that is most sensitive
(Lucassen et al., 2010). The dynamics of adult neurogenesis are to the neurotoxic effects of stress, prolonged exposure to high lev-
complex, involving the proliferation of new cells, the formation els of glucocorticoids can also cause damage in many other brain
of new immature neurons, and the insertion of adult born neu- regions, particularly the prefrontal cortex (PFC), where neurode-
rons into a functional network. Stress interferes both with the early generative changes include microglial activation (Hinwood et al.,
phases of this process (cell proliferation) as well as with neuronal 2011), atrophy of pyramidal neurons (Liu and Aghajanian, 2008),
survival and functionality (Czéh et al., 2001; Mirescu and Gould, dendritic retraction (Shansky et al., 2009; McEwen, 2010; Dias-
2006; Oomen et al., 2007; Wong and Herbert, 2004). Importantly, Ferreira et al., 2009), and reduction of synaptic proteins such as
these changes do not occur if chronic stress is predictable, or if PSD95, and synapsin I (Li et al., 2010). These changes are associated
coping with stress is possible: in these situations, neurogenesis is with cognitive deficits in PFC-related tasks (Miracle et al., 2006;
not suppressed, and may actually be stimulated (Lyons et al., 2010; Santini et al., 2004). It has also been reported that levels of N-acetyl-
Parihar et al., 2011). The mechanisms underlying this important aspartic acid, a marker of neurodegenerative processes, decreased
finding are explored further in Section 3.6. in the left dorsolateral PFC as a function of the length of depressive
illness (Nery et al., 2009), indicating that, as in the hippocampus,
3.2.2. Morphological consequences the effects of stress on the PFC are progressive.
These cellular effects of stress have visible morphological Within the PFC, the most stress-sensitive region is the anterior
consequences. Despite some inconsistencies in the literature, meta- cingulate cortex (ACC) (Drevets et al., 2008; Radley et al., 2006;
analyses of structural brain imaging studies have reported that Mayberg et al., 1997; Holmes and Wellman, 2009; McEwen and
there is a reliable decrease in hippocampal volume in patients suf- Gianaros, 2010; Hamani et al., 2011; Pizzagalli, 2011; Price and
fering from major depression (Campbell et al., 2004; Videbech and Drevets, 2010, 2012). A decreased volume of the ACC has also
Ravnkilde, 2004), and in rats exposed to CMS (Delgado et al., 2011), been reliably observed in imaging studies of depressed patients
comparable to the effect seen in Cushing’s disease (Sapolsky, 2000). (Phillips et al., 2003b; Savitz and Drevets, 2009; van Tol et al., 2010).
The effects vary in severity across hippocampal subregions (Cole Alterations in this area might contribute to the increased vulnera-
et al., 2010; Malykhin et al., 2010) and increase across episodes of bility to recurrence: for example, a history of recurrent depression
depression (Cole et al., 2010). In patients with late-onset depres- is associated with a low level of GABA in the ACC (Bhagwagar
sion, the decrease in hippocampal volume was correlated with et al., 2008). With the exception of this region, structural imaging
memory loss (Ballmaier et al., 2008). The major factor in these studies of other brain regions in depression are somewhat incon-
volumetric changes, in both preclinical studies of prolonged stress sistent, with structural damage, including volume and cell loss,
exposure (Tata and Anderson, 2010) and post-mortem studies of reported in many regions but not reliably replicated (Phillips et al.,
depressed patients (Czeh and Lucassen, 2007) is thought to be 2003b; Savitz and Drevets, 2009), albeit that consistency in some
changes in hippocampal morphology and loss of dendrites, rather regions increases with increased severity or chronicity of depres-
than cell loss. Some clinical studies have failed to observe volumet- sion (Lorenzetti et al., 2009). In general, however, these changes do
ric changes (McKinnon et al., 2009; MacQueen and Frodl, 2010). not increase with the duration of depression, or with its recurrence,
However, given the number and range of molecular processes that as it is the case for the hippocampal- and PFC-related changes. For
contribute to these effects, and the observation that the propor- example, the amygdala may be enlarged during a first episode of
tional loss of synapses is far greater than changes in hippocampal depression, but this seems not to persist with recurrence (Frodl
volume (Tata et al., 2006), it is likely that the functioning of the hip- et al., 2003; van Eijndhoven et al., 2009). Moreover, there is some
pocampus is compromised at levels of damage that are not reflected inconsistency between structural and functional imaging studies,
in visible morphological changes. such that volume loss is sometimes associated with increased func-
These findings are potentially of considerable importance for tional activity. Such effects have frequently been reported in, for
the ‘kindling’ hypothesis discussed above, because the degree of example, the amygdala (Savitz and Drevets, 2009).
hippocampal shrinkage is directly proportional to the number and In contrast to the relative inconsistency of structural imaging
duration of prior depressive episodes, and to the total duration of studies, there is a relatively well-accepted consensus regarding
2338 P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371

changes in functional activity in frontal brain regions of depressed Mayberg, 2009; Holtzheimer and Mayberg, 2010, 2011; Hamani
patients. This consensus reflects an increasingly well-established et al., 2011; Monkul et al., 2011).
model of the organization of information processing in the fore- Electrical stimulation of the ACC (deep brain stimulation: DBS),
brain. The model recognizes two distinct but interacting systems: using stimulation parameters that inhibit activity in the area tar-
a ventral “affective” circuit involving the amygdala, the anterior getted, has been found to elicit a prolonged recovery in some
(in rodents, ventral) hippocampus, the ventral striatum, the insu- treatment-resistant patients: over 50% of patients treated with DBS
lar cortex, the ventral (subgenual) part of the ACC and the ventral showed a greater than 50% improvement in depression scores a
and orbital PFC; and a dorsal “cognitive” circuit, involving the pos- year later (Mayberg, 2009; Hamani et al., 2011). DBS of the medial
terior (in rodents, dorsal) hippocampus, the dorsal (pregenual) part PFC has also been reported to elicit antidepressant-like effects in
of the ACC and the dorsolateral PFC. [There is some inconsistency in rats (Hamani et al., 2010). Clinical recovery following DBS of the
terminology with respect to the ventral ACC. In this review we use ACC was associated with a normalization of functional brain activ-
the terms ventral and subgenual ACC interchangeably to refer to ity in treatment responders (Mayberg, 2009; Hamani et al., 2011).
Brodman’s area 25 and the ventral parts of Brodman’s area 24 and Conversely, hyperactivity in this area was reinstated in patients
32 (Phillips et al., 2003a; Mayberg, 2009; Hamani et al., 2011); but after recovery from depression following a decrease in either 5HT
in some papers (e.g. Pizzagalli, 2011; Price and Drevets, 2012), the or DA and NA activity, brought about by a tryptophan depletion
ventral ACC is differentiated into anterior (rostral) and posterior (Neumeister et al., 2004) or AMPT (Hasler et al., 2008) challenge;
(subgenual) regions. Similarly the term ventromedial PFC refers to and in remitted depressives, the reactivity of this region to a sad
the more frontal Brodman’s area 10 but this term is sometimes film clip predicted relapse to depression (Farb et al., 2011). These
used to connote a larger area including the ventral ACC, which findings identify the ACC as a region of particular interest, as will
constitutes a ‘medial prefrontal network’ (Price and Drevets, 2010, be discussed further below.
2012). These inconsistencies reflect the limited resolution of many
neuroimaging studies.] 3.4. What drives the changes?
The ventral system is important for the identification of the
emotional significance of a stimulus, the production of affective While the functional changes in frontal brain regions associated
states and autonomic regulation related to emotionally signifi- with depression and successful treatment have been extensively
cant situations, while the dorsal system is important for executive described and discussed, there has been less consideration of their
function, including selective attention, planning, and effortful origin. As discussed above, there is evidence of stress-induced
regulation of affective states. This understanding is based on a con- neuropathology in many frontal regions, but this evidence is incon-
vergence of evidence from animal and human studies, involving sistent across studies, and often inconsistent with the functional
stimulation by emotionally salient stimuli, focal brain lesions, and changes. While it is possible that stress-induced neurotoxicity
functional neuroimaging techniques (Phillips et al., 2003a; Ernst could account for some of these effects, particularly the hypofunc-
and Paulus, 2005; Mayberg, 2009; Holtzheimer and Mayberg, 2010, tion of dorsal ACC, a simpler hypothesis is that the changes in the
2011; Hamani et al., 2011). functioning of frontal brain circuits could be secondary to struc-
In the human brain, there is also evidence for lateralization of tural and functional changes in hippocampal activity, based on the
emotional information processing, with a greater involvement of anatomical interactions between these two structures (Goldman-
the left hemisphere in appetitive behaviours and of the right hemi- Rakic et al., 1984; Jay et al., 1989; Thierry et al., 2000; Taxidis et al.,
sphere in avoidance behaviours. This distinction can be related to 2010). Indeed, the hippocampal formation, including the subicu-
the more traditional characterization of a left hemisphere involve- lum and the entorhinal cortex, contributes very significantly to the
ment in serial processing, and right hemisphere involvement in overall modulatory control and evaluation of the emotional con-
parallel processing, since approach behaviours require precise, cal- text of information processing by both the dorsal “cognitive” and
culated sequential strategies, while avoidance behaviours, which ventral “affective” systems. This nodal role reflects the anatomical
appear under conditions of danger or threat, require an imme- organization of the entorhinal cortex (which provides the major
diate holistic response (Kinsbourne, 1978; Willner, 1985). While input to the hippocampus), which receives convergent multimodal
lateralization of function has typically been associated with the and highly processed unimodal sensory inputs and has reciprocal
development of language, there is some evidence that lateralized connections with all the higher-order multimodal association cor-
processing has its origins in non-human species (Sullivan, 2004; tex areas, including parietal cortex, temporal cortex and the PFC
Corballis, 2009). (Canto et al., 2008).
From these observations it would be predicted that depression A causal relationship between decreased activity in the hip-
might be associated with decreased functioning in the dorsal “cog- pocampus and the decreased dorsolateral PFC activity that is
nitive” circuit, particularly in the left hemisphere, and increased reliably observed in depressed patients, is suggested by the obser-
functioning in the ventral “affective” circuit, particularly the right vation that decreased hippocampal volume was correlated with
hemisphere, and this is what is observed. Indeed, the functional the extent of executive dysfunction, which is known to be depend-
changes in the depressed brain appear to be exaggerated and pro- ent on the dorsolateral PFC (Frodl et al., 2006). There are several
longed versions of those seen during a normal response to aversive anatomical routes that could mediate this relationship. First, there
stimuli. Depressed patients have been reliably observed to show are direct connections, within the dorsal system, from subiculum
decreased bloodflow, under resting conditions, in the dorsal ACC of the hippocampus to dorsal ACC and dorsolateral PFC (Goldman-
and dorsolateral PFC, with greater effects in the left hemisphere Rakic et al., 1984). There are many reports that the connectivity
than the right, and increased bloodflow in the amygdala and ven- between these two regions is disturbed in, for example, schizophre-
tromedial PFC, with greater effects in the right hemisphere than the nia (e.g. Henseler et al., 2010) or Alzheimer’s disease (Goveas
left. These two sets of changes are associated with different sets of et al., 2011a); by contrast, hippocampal-prefrontal connectivity
symptoms: the decrease in (left) dorsolateral function is associated increased with severity of depression in a sample of depressed older
with psychomotor retardation, apathy and decreased attentional adults (Goveas et al., 2011b; Sheline et al., 2010). Second, there are
capacity, while the increase in (right) ventromedial function is asso- reciprocal inhibitory interactions between dorsolateral and ventro-
ciated with increased punishment sensitivity, anxiety, tension and medial PFC (Phillips et al., 2003a,b; Mayberg, 2009; Hamani et al.,
depressive ruminations (Mayberg et al., 1999; Phillips et al., 2003b; 2011; Pizzagalli, 2011), so a decreased level of activity in the dorsal
Savitz and Drevets, 2009; Vago et al., 2011; Elliott et al., 2011; system could simply follow from increased activity in the ventral
P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371 2339

system. Indeed, there is evidence that this is the case: for exam- Grace, 2005). Monosynaptic projections from the subiculum and
ple, in depressed patients performing a cognitive task, an increased basolateral amygdala have opposite effects on individual nucleus
ventral PFC response to errors was associated with a decreased accumbens neurons (French and Totterdell, 2003), controlling their
recruitment of the dorsal PFC on the following trial (Holmes and firing rate as well as that of VTA dopaminergic neurons (Gill and
Pizzagalli, 2008). A third mechanism, involving the dorsal striatum, Grace, 2011); and there is a similar convergence of monosynap-
is discussed below (Section 3.8). tic subicular and prefrontal cortical inputs to projection neurons
There are also strong reciprocal anatomical connections of the nucleus accumbens (French and Totterdell, 2002; Sesack and
between the hippocampus and amygdala (Pitkänen et al., 2000); Grace, 2010). Finally, the hippocampus may influence activity in DA
their reciprocal functional relationship is nicely illustrated by a (and 5HT) systems via the habenula, a limbic-striatal relay nucleus
study showing that amygdala memory coding was enhanced by that occupies an important position in the depression circuitry, as
lesions of the hippocampus, and vice versa (Richardson et al., 2004). discussed in the following section.
Interestingly, lesions of the basolateral amygdala, or reduction of To summarize, the symptoms of depression appear to reflect
its activity, suppress hippocampal neurogenesis (Kirby et al., 2012). changes in the activity of frontal brain areas that represent a height-
It is also evident from the opposite influences of hippocampus and ening (the ventral system) or an inhibition (the dorsal system) of
amygdala on the HPA axis (Fig. 3) that a decrease in hippocampal their usual role in the processing of and coping with affective infor-
inhibition of HPA activity leads to an increased hormonal stim- mation. Both sets of changes can be related to decreased activity
ulation of amygdala activity. As activity in the amygdala drives in the hippocampus. The data are for the most part correlational,
activity in the ventromedial PFC, an effect that is particularly promi- leaving open the question of where, precisely, the primary changes
nent in emotional disorders (Carlson et al., 2006), it follows that occur that propagate to other regions. However, the greater suscep-
a decreased level of hippocampal functioning causes a reciprocal tibility of hippocampal cells to suffer dysfunctional consequences
increase in the activity of the ventral “affective” system. when exposed to prolonged stress provides a plausible point of
origin for most if not all of these systemic dysregulations.
3.4.1. Anhedonia
Anhedonia, the inability to experience pleasure, is a core symp- 3.5. Mechanisms mediating affective information-processing
tom of depression that, in animal studies, has primarily been biases
associated with a decreased level of activity in the mesolimbic DA
projection from the ventral tegmental area (VTA) of the midbrain As already discussed, the amygdala and nucleus accumbens
to the nucleus accumbens, the major structure of the ventral stri- are key structures in the appraisal of and response to affectively-
atum (Willner and Scheel-Krüger, 1991; Wise, 2008; Treadway and valenced stimuli. As described in the introduction to this review,
Zald, 2011). Animals exposed to CMS are anhedonic to both natural depression is characterized by a heightened response to negative
and drug rewards (Willner et al., 1992) and have subsensitive DA D2 information and a blunted response to positive information (Beck,
receptors in the nucleus accumbens (Dziedzicka-Wasylewska et al., 1967; Disner et al., 2011). Increased activation of the amygdala
1997), as well as lower DA release within the nucleus accumbens in and decreased responsiveness of the nucleus accumbens are key
response to rewarding, but not to aversive, stimuli (Di Chiara et al., mediators of these affective information-processing biases.
1999). A recent study showed that the stress-responsive neuro-
peptide CRF causes DA release within the nucleus accumbens, and 3.5.1. Amygdala and nucleus accumbens
a rewarding effect, but following acute exposure to severe stress, The amygdala is involved in stimulus processing and the nucleus
CRF no longer releases DA, and its effect is now aversive (Lemos accumbens in response selection and contemporary accounts of
et al., 2012). These results parallel the clinical finding of decreased these structures emphasize that these functions are to some extent
responsiveness to reward but no change in responsiveness to pun- independent of the appetitive or aversive context (Sesack and
ishment in the ventral striatum of depressed patients (Robinson Grace, 2010; Gill and Grace, 2011). Thus, the amgdala has been
et al., 2011). The severity of anhedonia is negatively correlated shown to respond not only to aversive stimuli but also to appe-
with activity in the ventral striatum, in both unipolar depression titive stimuli, and to be involved in memory formation for both
(Keedwell et al., 2005; Dunn et al., 2002; Surguladze et al., 2005; types of stimuli, perhaps by coding their affective value (Paton
Epstein et al., 2006) and postpartum depression (Moses-Kolko et al., et al., 2006; Morrison and Salzman, 2010; Murray et al., 2011; van
2011), but positively correlated with activity in the ventromedial Wingen et al., 2010). Likewise, the nucleus accumbens, as output
PFC (Keedwell et al., 2005). Inversely to the PFC, activity in the ven- station of the PFC, is involved in decision-making in relation to
tral striatum increases with antidepressant treatment (Stoy et al., response costs in both appetitive and aversive contexts (Salamone,
2012; Ossewaarde et al., 2011), and DBS of the nucleus accumbens 1994). In both amygdala (Herry et al., 2008; Morrison and Salzman,
is an effective treatment for depression (Schlaepfer et al., 2008; 2010, 2011) and nucleus accumbens (Reynolds and Berridge, 2008;
Bewernick et al., 2010; Grubert et al., 2011). Faure et al., 2008; Richard and Berridge, 2011) there is some degree
Like other aspects of depression, anhedonia might also be of anatomical separation between appetitive and aversive regions.
explained by a stress-induced decrease in hippocampal func- Distinct appetitive and aversion regions within the ventral PFC
tioning. There are several pathways through which stress could and orbital cortex (Kringelbach and Rolls, 2004; Kringelbach, 2005)
influence activity in the ventral forebrain. Information regarding receive information from corresponding regions of the amygdala
the anticipation of reward (Schultz et al., 1997; Treadway and Zald, and project to corresponding regions of the nucleus accumbens
2011) reaches the VTA to activate the mesolimbic DA system via the (Humphries and Prescott, 2010; Haber and Knutson, 2011).
dorsolateral PFC (Ballard et al., 2011), suggesting a potential rela- Notwithstanding these generalized functions, the amygdala has
tionship between anhedonia and decreased activity in this region: traditionally been discussed primarily in relation to its very well-
the projection to the VTA is a very minor output of the dorsolat- established role in fear conditioning and anxiety disorders (Phelps
eral PFC and could be affected by changes in activity too small to and LeDoux, 2005; Etkin and Wager, 2007), and there is some
detect in a neuroimaging study. However, there is also an indirect evidence that the amygdala has a greater involvement in the
excitatory connection from the ventral subiculum of the hippocam- processing of aversive stimuli (Meseguer et al., 2007; Cybulska-
pus to the VTA, via the nucleus accumbens and ventral pallidum, Klosowicz et al., 2009; Dwyer, 2011). Thus, depressed patients
suggesting that a decrease in DA cell firing could follow from a showed not only a greater amygdala activation than controls when
decrease in hippocampal activity (Cooper et al., 2006; Lisman and processing negative or aversive stimuli, such as sad faces (Peluso
2340 P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371

et al., 2009a,b), but also, greater amygdala responses to sad than amygdala) and decreases the salience of positive events (via the DA
to happy faces, whereas the reverse was true in healthy controls projection from the VTA to the nucleus accumbens). The stimula-
(Victor et al., 2010). The increased amygdalar response persists tion of the habenula by stress, leading to an increase in the salience
after removal of the negative stimulus (Schaefer et al., 2002; Siegle of aversive stimuli (preferentially processed in the amygdala) and a
et al., 2007) and this effect is reversed by antidepressant therapy decrease the salience of appetitive stimuli (preferentially processed
(Fu et al., 2007). in the nucleus accumbens), provides a plausible neurobiological
Conversely, the nucleus accumbens has been discussed primar- substrate for the information-processing biases that characterize
ily in relation to reward processing. Indeed, as the structure through depression (Beck, 1967; Disner et al., 2011).
which emotional information gains access to the motor system
(Mogenson et al., 1980), the primary role of the nucleus accumbens
3.6. Failure to cope
could be conceptualized as seeking positive behavioural outcomes:
by generating behaviours aimed at acquiring rewards and avoiding
There is a crucial distinction between exposure to a stressor
punishment, activity in the nucleus accumbens improves hedonic
and the experience of stress. An intense stressor may cause little
tone in both appetitive and aversive environments (Willner, 1985;
stress if coping mechanisms are adequate, while conversely, if cop-
Ikemoto and Panksepp, 1999). Consistent with this formulation,
ing mechanisms fail, a mild stressor may be experienced as highly
depressed patients showed blunted accumbens responses to mon-
stressful. An account of the neurobiology of stress and depression
etary gains but did not differ from controls in their accumbens
must engage with the fact that the critical setting is a failure to cope
response to losses (Pizzagalli et al., 2009). Similarly, depressed
with stress, rather than exposure to stress per se.
patients showed lower activity in the ventral striatum when
reappraising rewards in a reversal-learning task, but not when
reappraising punishments (Robinson et al., 2011). Furthermore, 3.6.1. Stimulus appraisal
depressed patients show difficulty in maintaining positive mood This question has received considerable attention in the psycho-
after a reward, associated with an inability to sustain activity in the logical literature (e.g. Kessler et al., 1985; Bedi, 1999; Olff, 1999;
accumbens over time (Heller et al., 2009), suggesting that decreased Compass et al., 2004; Christensen and Kessing, 2005). The ability
activity within the dorsal PFC may reduce reward sensitivity of the to cope with stress depends in part on the neurobiological mech-
accumbens (Disner et al., 2011), either directly (Ballard et al., 2011) anisms already discussed, but also, and importantly, on how the
or indirectly via the hippocampus, as discussed earlier. stressor is appraised. Indeed, unlike physical stressors, the inten-
sity of which is communicated to the brain directly, psychological
3.5.2. The habenula stressors only come into existence at the point when they are
The lateral habenula is a key structure mediating the response to appraised within the brain. (Or as Shakespeare observed over 400
emotionally negative states (Hikosaka et al., 2008; Hikosaka, 2010), years ago, “there is nothing either good or bad but thinking makes
and the balance of activity between the amygdala and nucleus it so”.) Appraisal of the meaning of an emotion-eliciting situation
accumbens. The habenula is activated by stressful events in humans is a continuous process, and an initial appraisal may be followed
(Ullsperger and von Cramon, 2003) and animals (Matsumoto and by reappraisal in ways that could either increase or decrease the
Hikosaka, 2009). Metabolic activity in this region is increased in ani- perceived stress. The process of reappraisal is an important mech-
mal models of depression (Caldecott-Hazard et al., 1988; Shumake anism of self-regulation that has consequences for mental and
et al., 2003) and following the induction of a depressive episode by physical well-being (Davidson, 2000; Gross, 2002), as exemplified
tryptophan depletion in patients (Morris et al., 1999; Roiser et al., by the central role of reappraisal processes in cognitive behavioural
2009). Conversely, lesions of the lateral habenula reverse depres- therapy.
sive behaviour induced by learned helplessness or CMS (Amat et al., Brain mechanisms of reappraisal have been addressed directly
2001; Yang et al., 2008). Similarly, pharmacological or DBS inhi- by neuroimaging studies in which participants are required to mod-
bition of the lateral habenula has been shown to reverse learned ulate their emotional response to aversive stimuli, using strategies
helplessness (Winter et al., 2011; Li et al., 2011a,b), and benefi- such as imagining negative or positive outcomes to the situa-
cial effects of DBS have recently been reported in one severely tion presented, or by imagining that the situation is happening to
treatment-resistant depressed patient (Sartorius et al., 2010). Expo- themselves or to a stranger. These studies have shown that reap-
sure to stress increases the excitability of habenular output cells, praisal involves increased activation in a diffuse network of PFC
and the extent of this effect is correlated with the degree of learned regions, including elements of both the dorsal and ventral sys-
helpessless (Li et al., 2011a,b). tems. The regions activated depend to some extent on the cognitive
A major output of the lateral habenula is to the dorsal raphe strategy adopted: for example, self-focused regulation recruited
nucleus (DRN), and this projection is responsible for the stress- medial prefrontal regions implicated in internally focused infor-
induced activation of 5HT cells in the DRN (Amat et al., 2001; Takase mation processing, whereas situation-focused regulation recruited
et al., 2004; Yang et al., 2008). A second major output of the lateral lateral prefrontal regions implicated in externally focused infor-
habenula is to the VTA, causing inhibition of DA cells (Hikosaka mation processing (Ochsner et al., 2004; Ochsner and Gross, 2005;
et al., 2008; Hikosaka, 2010; Li et al., 2011a), resulting in anhedo- Roiser et al., 2012).
nia (Friedman et al., 2011). It is noteworthy that both stress and Reappraisal-related activation within the PFC is largely indepen-
antidepressant drugs increase 5HT function. The effect of stress is dent of whether the aim was to achieve an increase or a decrease
mediated by the projection from the DRN to the amygdala, where in negative affect. However, this is not the case in subcortical
it potentiates increases the salience of threat stimuli (Maier and structures. The act of reappraising a stimulus more negatively was
Watkins, 2010). But, as described below, serotonergic antidepres- associated with increased activity in the amygdala and decreased
sants act by potentiating the 5HT projection from the median raphe activity in the nucleus accumbens; conversely, a positive reap-
nucleus (MRN) to the hippocampus. The pro-anxiety effect of 5HT praisal was associated with decreased amygdalar activity (Ochsner
in the amygdala counteracts the antidepressant effect in the hip- et al., 2004; Ochsner and Gross, 2005; Wager et al., 2008). More-
pocampus, and the need to overcome the pro-anxiety effect may over, these changes in subcortical activity within the amygdala and
be one reason for the slow onset of antidepressant action. nucleus accumbens were shown to partly mediate the relationship
Thus, activation of the habenula by stress increases the salience between PFC activity and success in reappraising a negative event
of aversive stimuli (via the 5HT projection from the DRN to the (Wager et al., 2008).
P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371 2341

3.6.2. Learned helplessness the amygdala, leading to an increase in the salience of aversive
Parallel effects are observed in animal studies. Learned helpless- stimuli, and inhibits the nucleus accumbens, leading to a decrease
ness could reasonably be viewed as an animal model of a specific the salience of appetitive stimuli; and the descending control of
appraisal process. In the learned helplessness procedure, animals those processes by the PFC identifies the PFC as the site at which
display a range of behavioural impairments after being subjected the events that result in an episode of depression originate.
to uncontrollable aversive events that are not seen after expo-
sure to controllable events (Seligman, 1975; Maier and Seligman, 3.7. Rumination
1976). This is typically studied using inescapable vs. escapable
footshock, such that animals in the two conditions receive the This conclusion ties in with evidence that the ventromedial PFC
identical amount and pattern of shock, and differ only in their is the crucial region for processing self-related information. For
ability to control it. An elegant series of experiments demon- example this region is activated when people reflect on their own
strated that the perception of control depends on activity within internal state or on the thoughts or intentions of others, and when
the ventral PFC (Maier and Watkins, 2010). Pharmacological inac- they consider their own personality characteristics or receive social
tivation of the ventral PFC (using the GABA agonist muscimol) feedback about themselves from others (Passingham et al., 2010;
during exposure to escapable shock resulted in impairments typ- Lemogne et al., 2012; Murray et al., 2011). Negative self-referential
ical of animals exposed to inescapable shock, such as a failure of schemes are a crucial feature of cognitive models of depression
subsequent escape learning and reduced social investigation. Con- (Beck, 1967; Gusnard et al., 2001). Damage to, or a low level of
versely, pharmacological activation of the ventral PFC (using the activation in, the ventral PFC is associated with an inflated self-
GABA antagonist picrotoxin) during exposure to inescapable shock concept (Beer et al., 2006; Beer and Hughes, 2010), suggesting that
prevented subsequent behavioural impairments (Amat et al., 2005, an elevated level of activation in this area could be responsible for
2008; Christianson et al., 2009; Maier and Watkins, 2010), con- the low self-esteem and sense of social inferiority that characterize
sistent with the antidepressant-like effect of DBS in this region depression (Sloman et al., 2003; Gilbert, 2006).
(Hamani et al., 2010). Moreover, experience of controllability, by Further, high activity in the medial PFC, as well as the amyg-
exposing animals to escapable shock, immunizes them against the dala, is associated with a trait that is predominant in depressed
behavioural impairments caused by subsequent inescapable shock patients, mental rumination, the inability to interrupt the pro-
(Amat et al., 2006), or a different stressor, social defeat (Amat et al., longed processing of negative emotions (Cooney et al., 2010; Siegle
2010; Maier and Watkins, 2010). The learning process responsible et al., 2006). Rumination is associated with self-referential attri-
for this resilience also occurs within the ventral PFC, because it was bution, self-referential appraisal and a failure of inhibition; and
blocked by injection of a protein synthesis inhibitor into this region a decrease in rumination by cognitive reappraisal results in a
during exposure to the controllable condition (Amat et al., 2006). decrease in medial PFC activity (Ray et al., 2005). Depressed patients
Therefore, the ventral PFC appears to be responsible for appraisal of have been reported to show greater activation in the medial PFC and
whether or not it is possible to cope with stress not only in humans ventral ACC during self-referential processing of negative stimuli,
but also in the rat. which was correlated with the severity of depression (Yoshimura
Studies using this model have shown that the PFC–amygdala et al., 2010). Rumination may be reflective (problem solving aimed
interaction operates via the ascending 5HT system. It has long been at alleviating distress) or brooding (passively dwelling on the iniq-
known that the adverse behavioural effects of inescapable shock uity of one’s current state). Reflective and brooding rumination are
are dependent on the integrity of 5HT neurons in the dorsal raphe predictive, respectively, of lower and higher levels of depressive
nucleus (DRN), because pharmacological or surgical inactivation of symptomatology years later, suggesting that reflective rumina-
these neurons prevents learned helplessness (Maier and Watkins, tion is an adaptive cognitive process but brooding is maladaptive
2010). Stimulation of the ventral PFC inhibits DRN 5HT cells via (Treynor et al., 2003).
glutamatergic projections to GABAergic neurons within the DRN The ventral ACC and ventromedial PFC are part of the default
(Hajos et al., 1998; Baratta et al., 2009). When control of stress is mode network (DMN) of cortical and subcortical regions that are
appraised to be absent, this inhibition is lifted, leading to desensi- active when an individual is awake but not focused on the out-
tization of inhibitory 5-HT1A DRN auto receptors (Rozeske et al., side world (Broyd et al., 2009; Lemogne et al., 2012). The DMN is
2011) and increased stress-induced activation of a population of inhibited when engaged in externally focussed task performance
5HT cells in the dorsomedial DRN (Lowry et al., 2008). This in turn and preferentially activated when engaged in internally focussed
results in increased release of 5HT in the amygdala (Amat et al., processes such as daydreaming and memory retrieval. We hypoth-
2006), with a consequent activation of the HPA axis (Fig. 3). esize that when stress is perceived as uncontrollable (an appraisal
The PFC also projects to the habenula, and this projection plays that may be reached rapidly as in a learned helplessness experi-
a key role in setting the balance of negative and positive affectiv- ment or slowly in real world conditions), this results in a decision
ity. The habenula and its outputs are activated by stress without to disengage from externally focussed coping attempts, thereby
regard to the controllability of the stressor (Amat et al., 2001). engaging the DMN, and enabling internally focussed rumination.
However, if stress is appraised as controllable, habenular outputs Rumination, under these conditions, is likely to focus on the cur-
to the amygdala (via 5HT) and nucleus accumbens (via DA), and rent adversity, and significantly, in depressed patients, increased
their behavioural consequences, are over-ridden by descending activity in the DMN is associated with higher levels of maladap-
projections from the ventral PFC (Maier and Watkins, 2010). Indeed, tive depressive ruminations (brooding) and lower levels of adaptive
individual neurons within the PFC project to both the serotonergic reflective ruminations, implying poorer coping (Hamilton et al.,
DRN neurons and the dopaminegic VTA neurons (Vazquez-Borsetti 2011). Furthermore, unlike non-depressed controls, depressed
et al., 2011), and within the DRN these projections terminate on the patients fail to deactivate the DMN when asked to reappraise
same cells innervated by the lateral habenula (Varga et al., 2003). negative pictures as positive (Sheline et al., 2009). The processes
These effects parallel the observation that when negative informa- whereby depressed patients become “stuck in a rut” and unable
tion is reappraised positively by human volunteers, the extent to to self-regulate their emotional state (Holtzheimer and Mayberg,
which increased activity in the ventral PFC results in a decrease in 2010) remain to be fully established. However, three processes
negative emotion depends on the extent to which activity decreases that are known to be involved include: a dysregulated subcortical
in the amygdala and increases in the nucleus accumbens (Wager network and HPA axis; a self-sustaining vicious circle in which
et al., 2008). The stimulation of the habenula by stress activates attention to negative information promotes negative thinking and
2342 P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371

vice versa (Beck, 1967; Disner et al., 2011); and in the longer term, (Watkins et al., 2011). If activity in the ventral ACC is low, then
stress-induced neurotoxicity. these resources are available to be recruited during the course
of psychological treatments (DeRubeis et al., 2008). It is gener-
3.7.1. Rumination and recovery ally assumed that CBT interventions restore emotional control by
Several of the functional changes seen in depression have been providing top–down inhibition of amygdalar responses to threat
found to normalize with successful treatment. A consistent pattern (Egner et al., 2008; Pizzagalli, 2011; Roiser et al., 2012). The ventral
of decreased activity in amygdala and ventral PFC (particularly, the ACC is central to this process by virtue of its having more extensive
ventral/subgenual/rostral ACC) and increased activity in the dor- connections with the amygdala than other cortical regions (Price
solateral and pregenual ACC has been reported in patients who and Drevets, 2010, 2012).
recovered from depression following successful treatment with In contrast to the different response of the ventral ACC to physi-
antidepressant drugs or ECT, but not in patients who failed to cal and psychological treatments, similar changes were seen in the
respond to treatment (Phillips et al., 2003b; Savitz and Drevets, ventromedial and orbital PFC following antidepressant drug treat-
2009; Hamani et al., 2011; Mayberg, 2009; Holtzheimer and ment and CBT, involving a normalization of both resting metabolic
Mayberg, 2011; Pizzagalli, 2011). Similar changes were also seen activity and the response to emotional stimuli (Kennedy et al.,
in depressed patients who responded to placebo treatment, though 2007; Mayberg, 2009). Thus, while the ventral ACC is a critical
fluoxetine and placebo had different effects in other brain regions region for antidepressant effects, it appears that depressed mood is
(Mayberg et al., 2002). generated primarily by the ventromedial PFC (as described already
Not only does activity within the ventral ACC decrease with in Section 3.6).
successful drug treatment of depression, but also, a high level of
activity in prior to treatment is predictive of a successful response 3.8. Basal ganglia involvement in coping
to antidepressants (Mayberg et al., 1997; Pizzagalli, 2011). This
has been hypothesized to mean that the increased activity in the As previously discussed, the ventral ‘affective’ circuit, involv-
ventral ACC may reflect a greater proclivity to engage in adaptive ing the ventral and orbital PFC and the ventral ACC is involved
rumination (Pizzagalli, 2011). Conversely, difficulties in deacti- in the production of affective states, while the dorsal ‘cognitive’
vating the ventral ACC and providing top–down regulation of the circuit is important for executive function, including the effortful
amygdala might foster the emergence of maladaptive emotional regulation of affective states. But in addition to the mechanisms
and self-focused rumination, leading to treatment nonresponse discussed above, the failure of the dorsal PFC to regulate emotion
(Pizzagalli, 2011). Furthermore, as expected given the reciprocal in depression, reflected in an inability to disengage from negative
relationships between the ventral “affective” and dorsal “cognitive” information (Gotlib and Joormann, 2010), may also involve the dor-
systems, the difficulty of deactivating the ventral ACC in depres- sal striatum.
sion is associated with difficulty in activating the dorsal PFC during Activity in the ventral striatum (nucleus accumbens) is trans-
resource-demanding cognitive tasks, resulting in impairments of mitted to the dorsal striatum (caudate nucleus, putamen) via a
executive functioning, cognitive control and emotional regulation well-characterized DA-dependent ‘spiralling’ of efferent informa-
(Ochsner and Gross, 2005; Pizzagalli, 2011). tion from ventral striatum to the substantia nigra and thence to
However, the opposite relationship has been observed for more dorsal striatal regions, so channelling information from limbic
psychological treatments (Roiser et al., 2012): a low level of pre- to cognitive and finally to motor circuits (Haber, 2003; Haber and
treatment activity in the ventral ACC was associated with a better Knutson, 2010). Through this circuit, blockade of DA transmission
response to CBT (Siegle et al., 2006; Fu et al., 2007) or behavioural in the ventral striatum (nucleus accumbens core) causes an increase
activation (Dichter et al., 2010). Indeed, in contrast to the decreased in DA function in the ventrolateral striatum but a further decrease
activity in the rostral ACC that characterizes the response to antide- in DA function in the dorsal striatum (A. Cools, pers. comm.). The
pressant treatment, metabolic activity in this region increased in dorsal striatum, via its projections to the midbrain, innervates the
patients who responded to CBT (Kennedy et al., 2007; Mayberg, dorsolateral prefrontal and dorsal cingulate cortical regions (Haber
2009; Hamani et al., 2011). et al., 2006), and thus occupies a nodal position for the integration
Although there are anatomical subdivisions within the rostral of emotional processes with cognitive control.
ACC (Jones et al., 2005; Palomero-Gallagher et al., 2009) it seems Pathological volume changes in the caudate nucleus and puta-
unlikely that this can explain the different patterns of response to men of depressed patients have been reported, which are particular
physical and psychological treatments because the sites suppor- prominent in late-life depression and treatment-resistant patients
ting these two effects are intermingled (Roiser et al., 2012). The (Bora et al., 2011; Shah et al., 2002; Pizzagalli et al., 2009), and
difference between treatment modalities may rather depend upon involve decreases in gray matter volume (Kim et al., 2008) and
the extent to thinking is dominated by maladaptive rumination. neuronal density (Khundakar et al., 2011). The extent of anterior
Given that increased activity in the ventral ACC is associated with caudate reduction in late-life depressed subjects was associated
rumination (Cooney et al., 2010), the suppression of this activity by with more severe depression (Butters et al., 2009), and may also
antidepressants and other physical treatments may improve the contribute to suicidal behaviour in severely depressed patients
depressed person’s ability to disengage from emotionally negative (Ahearn et al., 2001; Dombrovski et al., 2011; Vang et al., 2010), by
thoughts. Consistent with this account, brief (7-days) SSRI treat- impairing their emotional decision-making and thereby increasing
ment has been found to decrease metabolic responses to negative impulsivity (Wagner et al., 2011). Depressed patients also have a
self-descriptors in neurotic individuals at risk for depression (Di low level of DA activity in the caudate nucleus, which may in part be
Simplicio et al., 2011), an effect that may be mediated by 5HT1A secondary to an impairment of DA transmission in the ventral stri-
receptors (Hahn et al., 2012). However, a high level of rumina- atum, but is also related to an increase in MAO-A activity, the extent
tion interferes with cognitive processes that are used in cognitive of which is associated with the severity of psychomotor retardation
behavioural therapy (CBT), such as shifting attention, disengage- (Meyer, 2012).
ment from negative thoughts, problem solving and other executive Depressed patients have been reported to show enhanced
functions (Ray et al., 2005; Joormann and Gotlib, 2008; Nolen- metabolic activity not only in the ventral ACC, but also in the
Hoeksema et al., 2008), and may impact negatively on the outcome left putamen when attempting to disengage from negative words
of conventional CBT (Haeffel, 2010). Indeed, specific techniques (Eugene et al., 2010; Surguladze et al., 2005). The extent of alter-
of ‘rumination-focussed CBT’ have been developed for this reason ations in the connectivity of the caudate nucleus to other brain
P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371 2343

regions was correlated with disease severity and duration in drug-


naïve first-episode patients (Zhang et al., 2011). Other studies have
suggested that metabolic changes in left caudate nucleus (Gabbay
et al., 2007) and deficits in its connectivity with the resting-state
default mode network (Bluhm et al., 2009) may be early mani-
festations of depression. Recovered depressed patients also show
increased activity in caudate nucleus to aversive stimuli, in contrast
to the decreased response to pleasant stimuli that is typically seen
in the ventral striatum (McCabe et al., 2009). Similarly, increased
activation of the ACC and caudate nucleus by negative stimuli,
which was suggested to reflect the effort to disengage from and
cope with the negative information, has also been reported in
unaffected relatives of depressed patients (Lisiecka et al., 2011;
Levesque et al., 2011). These studies suggest that an altered pat-
tern of dorsal striatal activation by negative stimuli is present not
only in depression but also in risk states for depression, and may
contribute to failure to cope.
In addition to the dopaminergic interaction between the ventral
and dorsal striatum described above, a second mechanism for this
interaction is provided by the innervation of the medial part of the
caudate nucleus by the ventral ACC. Rats exposed to CMS showed
atrophy of not only the ventral ACC but also the dorsomedial
Fig. 4. Stress and depression. The etiological role of stress in depression is medi-
striatum, together with hypertrophy of the dorsolateral striatum
ated by activation of the HPA axis, leading to functional, and later, morphological,
(Dias-Ferreira et al., 2009). The dorsomedial striatum is implicated damage in the hippocampus that results in a widespread disruption of information
in the acquisition and execution of goal-directed actions, whereas processing in the forebrain, with the detailed disruptions determining the specific
the dorsolateral striatum is critical for habit formation (Tricomi symptoms displayed.
et al., 2004; Yin et al., 2005, 2006; Balleine and O’Doherty, 2010).
Consistent with this anatomy, rats exposed to CMS were unable to in particular the DRN with a consequent increase in 5HT release
regulate their behaviour in response to sudden changes in reward within the amygdala and further release of CRF from the hypothal-
contingencies, and continued to respond in habitual ways that were amus and stimulation of the other elements of the HPA axis. As a
no longer appropriate (Dias-Ferreira et al., 2009). These results consequence of prolonged exposure to stress hormones, the hip-
suggest that changes in information processing within the dorsal pocampus undergoes neurotoxic damage, with a loss of synaptic
striatum, more specifically, the balance of activity within the ACC- connectivity and disruption of information processing in the fore-
innervated medial region that is involved in behavioural flexibility brain circuits in which the hippocampus participates. This broadly
and the lateral region that promotes habitual responding, could involves increased activity within a system that includes, inter
go some way to explain the emotional problem-solving difficulties alia, the amygdala and ventral PFC, and decreased activity within a
(Gotlib and Joormann, 2010) that leave depressed patients “stuck in system that includes, inter alia, the dorsal PFC and nucleus accum-
a rut” (Holtzheimer and Mayberg, 2011), as well as the high comor- bens; and the symptoms of depression arise from the enhancement
bidity between depression and compulsive disorders (Dias-Ferreira or inhibition of the information-processing roles of the affected
et al., 2009). regions in the normal brain. (We should add that our discussion
has greatly simplified the complex anatomy of the forebrain and
3.9. Summary: stress and depression in conditions of low has not addressed in any detail the differentiation of subregions
vulnerability within the ventral and dorsal systems.)

Fig. 3 provides a simplified picture of the anatomical systems 3.10. Vulnerability to depression: 1. Kindling
involved in affect regulation that are described above. The green
arrows describe a basic input–output pathway, in which incom- The model described provides a basis against which to consider
ing stimuli are appraised by the amygdala and hippocampus and the nature of vulnerability to depression, which, as outlined above,
information is transmitted to the basal ganglia where an appro- arises from a variety of diatheses, one of the most potent of which
priate response is generated. Affective information-processing is prior experience of depression.
biases are generated by the serotonergic and dopaminergic inputs We noted earlier that there is evidence that the depressed brain
to this system from the DRN and VTA, which are themselves does not fully recover. For example hippocampal volume changes
modulated by the lateral habenula (blue arrows). These biases persist after remission of symptoms (Neumeister et al., 2005; Cole
may be over-ridden by top–down control from the PFC (red et al., 2010), as do reduced frontocingulate functional connectivity
arrows). (Aizenstein et al., 2009) and neurochemical abnormalities in the
The story so far is summarized in Fig. 4, which describes a first ACC and PFC (Bhagwagar et al., 2006, 2008). Consistent with these
episode of depression in a person with a weak depressive diathesis. observations, there are reports of persistent cognitive (Bhalla et al.,
In these conditions, vulnerability to depression is low, and the onset 2006; Preiss et al., 2009) and sensory (Naudin et al., 2012) impair-
of depression is driven by exposure to high levels of stress, which ments, and increased responsiveness to aversive stimuli (McCabe
activate the HPA axis via the amygdala. When stress is prolonged et al., 2009), following recovery from a depressive episode. Persis-
(for example, following an unreplaced loss) this is likely to lead to tent cognitive and neurochemical changes are also seen in rodents
an appraisal that coping attempts have been and will continue to be after recovery from CMS (Elizalde et al., 2008, 2010; Surgetj et al.,
unsuccessful. This decision, which may be reached rapidly or grad- 2009).
ually, involves computations carried out by a distributed network Also as noted earlier, successive episodes of depression are
of PFC regions, but is housed primarily within the ventromedial PFC. triggered by progressively weaker and weaker episodes of stress
It leads to a disinhibition of subcortical stress-responsive systems, (Stroud et al., 2011; Slavich et al., 2011). Recurrent episodes are
2344 P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371

also associated with more intense symptomatology (Robinson a chronic risk factor to contribute to the onset of a depressive
and Sahakian, 2008): for example, recurrent patients displayed episode may be amplified by its association with an episode of
more intense electrophysiological responses to negative words depression.
(Nandrino et al., 2004) and a greater cortisol response to a chal-
lenge with the 5HT-releasing drug fenfluramine (Sher et al., 2004). 3.11. Vulnerability to depression: 2. Predisposition
Because there are very few studies using animal models of recur-
rent depression, little is known about the mechanisms of this effect. Recent research has established that early life stress has endur-
However, the persistent hippocampal damage following an episode ing effects on the adult brain, many of which closely resemble the
of depression could decrease the threshold at which stress acti- changes observed in the depressed brain, and which are therefore
vates the amygdala and engages the processes summarized in Fig. 4. likely to underlie the established cognitive and temperamental vul-
The amygdala may well be involved in the kindling effect across nerabilities to depression (Heim et al., 2008; Pechtel and Pizzagalli,
repeated episodes of depression, given the extensive literature on 2011). For example, people who had been abused or institutional-
amygdala kindling in relation to epilepsy (Post, 2004). ized as children were more anhedonic than non-abused controls,
It was recently reported that the kindling phenomenon was and they had decreased reward-related activity in the nucleus
particularly apparent in relation to a specific type of stressor, inter- accumbens (Dillon et al., 2009; Mehta et al., 2010). Adults who
personal loss: over the first four episodes of depression, there have been traumatized as children also show greater electrophys-
was a 65% decrease in the intensity of stress associated with iological responses to angry faces or voices (Pechtel and Pizzagalli,
the loss that precipitated the episode (Slavich et al., 2011). No 2011), as well as a greater amygdalar response (Taylor et al.,
such change was observed in the intensity of non-loss stressors 2006). A decrease in hippocampal volume is also seen in adults
(though it is unclear whether this reflects a qualitative differ- who experienced childhood trauma (Heim et al., 2008; Pechtel
ence between social and non-social stressors or simply that the and Pizzagalli, 2011), and in one study, decreased hippocampal
intensity of non-social stressors was relatively low in this study). volume was present in depressed patients who had experienced
Like other stressors, social defeat in animals is known to increase childhood trauma, but absent in those who had not (Vythilingam
circulating levels of inflammatory cytokines (e.g. Avitsur et al., et al., 2002). Moreover, both hyperactivity of the amygdala and
2001, 2009). Social stressors such as rejection also increase lev- dorsolateral prefrontal hypoactivity have been reported in non-
els of inflammatory cytokines in people (Dickerson et al., 2009; depressed people with a strong cognitive diathesis for depression,
Slavich et al., 2010). The increasing potency of social stress to elicit a high level of hopelessness; indeed, among depressed patients,
depression (Slavich et al., 2011), and the observation of a greatly hopelessness was more strongly associated than severity of depres-
accelerated age-related increase in cytokine production among sion with these functional changes (Zhong et al., 2011). A decrease
carers, who experience chronic high levels of social stress (Kiecolt- in 5HT transporters has also been reported in the dorsolateral PFC
Glaser et al., 2003), raise the possibility that cytokine-induced of non-depressed people with a genetic diathesis for depression (a
‘sickness behaviour’, which closely resembles the symptomatology depressed co-twin: Frokjaer et al., 2009).
of depression (Weisler-Frank et al., 2005; Sharpley and Agnew, As in depression itself, these neurological vulnerabilities to
2011), might play a relatively greater role in recurrent depressions depression may result from dysregulation of the HPA axis. Many
than in first episodes. This would have implications for treatment. studies in children and adults have reported that early life stress
In addition to structural changes that promote physiologi- leads to a more reactive HPA axis, which probably results from
cal sensitization, associative learning processes may also operate, a decreased expression of glucocorticoid receptors with a conse-
such that weaker stressors are able to elicit symptoms of depres- quent decrease in inhibitory feedback control (Heim et al., 2008;
sion by virtue of a prior association between the physiology of Essex et al., 2011; Wilkinson and Goodyer, 2011). Maternally sepa-
stress and depressive cognitive processes (Robinson and Sahakian, rated rats, an animal model of early life stress, show a similar HPA
2008; Roiser et al., 2012). A similar explanation has previ- dysregulation when adult (Rentesi et al., 2010), as well as decreases
ously been advanced in relation to the depressogenic effect of in levels of BDNF, hippocampal cell proliferation and synaptic plas-
reserpine-induced catecholamine depletion, which, like the effect ticity (Aisa et al., 2009).
of tryptophan depletion is only seen in people with a history As with repeated experience of depression, these effects of early
of depression (Goodwin et al., 1972; Willner, 1985). Tryptophan adversity prime the brain to respond in a depressive manner to
depletion caused different changes in the neural response to emo- lower levels of stress (Slavich et al., 2011). Genetic diatheses appear
tional words in remitted depressive patients and healthy controls to operate in a similar manner. The most studied of these effects is
(Roiser et al., 2009) and similarly, AMPT affected reward processing a functional polymorphism in the promoter region of the serotonin
differently in remitted depressive patients and healthy controls transporter gene, 5HTTLPR, which has been reported to moderate
(Hasler et al., 2009), suggesting that there might be trait-like the effects of stressful life events on depression (Caspi et al., 2003,
abnormalities of monoamine systems in depressive individuals. 2010; Uher and McGuffin, 2008). While one meta-analysis failed to
The associative learning hypothesis makes a general prediction confirm this effect (Risch et al., 2009), a subsequent, larger meta-
that physiological processes that contribute to the symptomato- analysis supported the original finding, but in relation to chronic
logy of depression may be relatively innocuous in people who (family or medical) stress only, not to episodic life events (Karg
have not experienced a depressive episode, but would act as et al., 2011). The same result was reported in a recent prospective
risk factors for a later episode if still present after recovery. For study: the incidence of depression was increased under conditions
example, as already described, increased levels of MAO-A activ- of chronic (but not episodic) stress by the presence of the short
ity have been reported in the brains of unmedicated early-onset allele of the 5HTTLPR gene (Jenness et al., 2011). The short allele
depressed patients, consistent with decreased levels of 5HT, NA produces fewer 5HT transporters and therefore clears 5HT from
and DA (Meyer et al., 2006; Meyer, 2012), and this abnormal- the synapse less efficiently than the long allele, with the result
ity was not corrected by SSRI treatment (Meyer et al., 2009). that the 5HT system responds less flexibly and is more coarsely
Subsequently, high levels of MAO-A activity in the PFC and ACC attuned to changes in demand (Belsky et al., 2009). Consistent
following recovery from depression predicted a recurrent depres- with this observation, individuals at high familial risk for mood
sive episode in the following 6 months (Meyer et al., 2009; Meyer, disorders had lower levels of 5HT transporter binding in the PFC
2012), in the absence of other factors predictive of recurrence. (Frokjaer et al., 2009). The short allele of the 5HTTLPR gene, as
These results are consistent with the notion that the potency of well as a polymorphism of the 5HT1A receptor, was associated
P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371 2345

with increased amygdala activation in depressed patients (Frodl that is essential for the inhibitory effect of DA D2 receptors on cAMP
et al., 2008b; Furman et al., 2011; Munafò et al., 2008). Carriers of signaling, display a depression-like phenotype (Park et al., 2005). A
the short allele have been reported to have increased HPA stress- similar effect was reported in two lines of knockout mice deficient
reactivity (Gotlib et al., 2008), suggesting that these effects may in the sigma1 receptor (Sabino et al., 2009). Depression-like phen-
result from a potentiation of 5HT action in the amygdala (Lowry otypes are also seen following genetic manipulation of the HPA
et al., 2008; Rozeske et al., 2011; Furman et al., 2011). Conversely, axis. For example, conditional mouse mutants overexpressing CRF
carriers of the long allele respond better to treatment of depres- either in the entire central nervous system (CRH-COE-Nes) or only
sion with SSRIs (Serretti et al., 2007), probably, as discussed below, in the forebrain (CRH-COE-Cam) both displayed increased rapid eye
by potentiating 5HT in the hippocampus. A polymorphism of the movement (REM) sleep, which is a characteristic symptom of major
BDNF gene has been reported to protect against the depressogenic depression (Kimura et al., 2010). Mice deficient in CRF2 receptors
effect of the 5HTTLPR polymorphism (Pezawas et al., 2008) and (CRF2−/− mice) also display depressive-like behaviours, accom-
to potentiate the clinical response to antidepressants (Zou et al., panied by increased hippocampal levels of a variety of activated
2010). (phosphorylated) protein kinases (Todorovic et al., 2009). Finally,
Many post-mortem studies of depressed patients have reported depressive-like behaviour is seen in ovariectomized ER␤ knockout
an increase in the density of cortical 5HT2A receptors, which mice (Rocha et al., 2005), which seems to be mediated by hip-
is associated with a decrease in 5HT release (Arango et al., pocampal estrogen receptors as intra-hippocampal administration
1997). In an in vivo study, 5HT2A receptor density was correlated of estrogenic compounds elicits an antidepressant-like effect (Walf
with the severity of dysfunctional attitudes (Meyer et al., 2003; and Frye, 2007). For the majority of genes that have been reported
Meyer, 2012). High levels of 5HT2A binding in the PFC were also to influence vulnerability in animal models that reproduce aspects
found in unmedicated patients following recovery from depres- of depression, it is not yet known whether they are also of clinical
sion (Bhagwagar et al., 2006). This marker was correlated with relevance.
the degree of neuroticism, and in particular, the sub-trait of vul-
nerability, in non-depressed volunteers (Frokjaer et al., 2008), but
3.11.2. The vulnerability brain and the depressed brain
only in those at high familial risk for depression (Frokjaer et al.,
While the evidence base concerning the mechanisms underlying
2010). Thus, while an increase in cortical 5HT2A receptors may be
vulnerability to depression is less complete than that for depression
a marker of the depressed state, it also appears to be a trait marker
itself, the available evidence indicates strongly that both repeated
of vulnerability to depression.
experience of depression and other diatheses – both experien-
tial and genetic – prime the brain to enter a state of depression
3.11.1. Animal models of vulnerability
more readily by reproducing some of the neurobiological features
Vulnerability to depression can also be studied in animal mod-
of depression. These parallels between the vulnerable brain and
els (Willner and Mitchell, 2002b). Like depressed people, rats
the depressed brain are observed both in the clinic and in animal
bred to display helplessness are anhedonic (Vollmayr et al., 2004)
models of depression. As a result of this priming, and the conse-
and show decreased activity in the dorsal PFC but increased
quent depression-like brain changes, depressions that occur in the
activity in the ventral ACC (Shumake et al., 2000) and lateral
presence of a strong diathesis may be precipitated by lower levels of
habenula (Shumake et al., 2003), and other metabolic changes
stress, and accompanied by a smaller physiological stress response.
that resemble those seen in depression (Shumake and Gonzalez-
The treatment implications of this different physiology are illus-
Lima, 2003), as well as lower levels of 5HT uptake and BDNF
trated by a study comparing the effectiveness of treatment of
in the hippocampus (Aznar et al., 2010). A similar mouse model
depression with a serotonergic antidepressant (nefazodone), CBT,
has recently been described (Yacoubi et al., 2011). Vulnerability
or their combination. In patients with a weak depressive diathesis
can also be approached via the individual differences in suscep-
(no early life stress), the drug treatment was somewhat better than
tibility to stress that exist naturally in a large rodent population.
CBT and the effects were additive for the combination treatment.
For example, vulnerability to social defeat has been shown to be
However, in patients who had experienced early life stress, CBT
mediated by increases in the electrical activity of DA-producing
was significantly better than nefazodone, and adding nefazodone
neurons in the VTA (Krishnan et al., 2007), while resistance to
to CBT did not improve the outcome. This result was interpreted
learned helplessness is mediated by induction of the gene encod-
to suggest that psychotherapy might be an essential element of
ing the transcription factor FosB in the periaqueductal grey region
treatment for depressed patients with childhood trauma (Heim
(Berton et al., 2007). A recently gene profiling study identified sev-
et al., 2008). An alternative interpretation is that the effectiveness
eral hundred genes that were differently regulated in vulnerable
of some antidepressants may depend on the extent to which the
and resistant mice characterized as vulnerable or resistant to CMS
patient is currently stressed. We return to this issue below.
(Strekalova et al., 2011), and as discussed further below, parallels
in gene and protein expression between vulnerability to CMS and
resistance to antidepressant treatment have also been reported 4. Mechanisms of antidepressant action
(Bisgaard et al., 2007; Wilkinson et al., 2009; Christensen et al.,
2011). 4.1. Potentiation of monoamine transmission
A third approach involves the generation of transgenic mice
with modification of genes associated with resilience or vulner- Antidepressant drugs are assumed to act primarily via
ability. There are now many such examples. For example, knock-in monoaminergic mechanisms, but there has been considerable
mice expressing the variant of the BDNF gene (Met-66 BDNF debate in the literature as to whether they potentiate transmission
homozygotus mice) that protects against depression in people at monoaminergic synapses, as originally proposed, or decrease it
(Pezawas et al., 2008) show resilience in the social defeat depres- (Segal et al., 1974). However, neurotransmitter depletion studies
sion model (but not in models of anxiety) (Krishnan et al., 2007; provide strong evidence that the primary action of antidepressant
Chen et al., 2006). Genetic disruption of the acid-sensing ion drugs is as conceived originally: enhancement of neurotransmis-
channel-1a (ASIC1a) produced an antidepressant-like effect in the sion at 5HT and NA synapses. As discussed earlier (Section 2), in
forced swim test, which was reversed by restoring ASIC1a in the both human and animal studies, the effects of SSRIs are blocked
amygdala of ASIC1a (−/−) mice (Coryell et al., 2009). Conversely, by antagonizing transmission at 5HT synapses, while the effects
mice lacking the prostate apoptosis response 4 (Par-4), a protein of NRIs are blocked by antagonizing transmission at NA synapses.
2346 P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371

As to the location of the relevant synapses, there is evidence that mesolimbic DA system (Cador, 1992; Humphries and Prescott,
antidepressant actions can be antagonized by various interventions 2010; Sesack and Grace, 2010; Haber and Knutson, 2011). This cir-
within the hippocampus (reviewed below) or prefrontal cortex cuitry provides the basis for the role of the nucleus accumbens
(Furr et al., 2011; Bondi et al., 2010). as a ‘limbic-motor interface’, in which DA controls the transmis-
The clinical selectivity of 5HT depletion for 5HT-selective sion of information from limbic forebrain structures to the motor
antidepressants (Delgado et al., 1990, 1999) and NA depletion system (Mogenson et al., 1980) feeding forward into the dor-
for NA-selective antidepressants (Delgado et al., 1993; Miller sal striato-pallido-thalamic-prefrontal psychomotor loop that has
et al., 1996) also demonstrates another important point: that 5HT- been characterized in both humans and animals (Haber and Knut-
and NA-selective antidepressants act via different mechanisms, son, 2011; Humphries and Prescott, 2010). The expression and
involving 5HT synapses and NA synapses, respectively. Studies of functional sensitivity of DA D2 receptors in the nucleus accum-
antidepressant action in animal models of depression replicate bens is decreased by CMS and restored by antidepressant treatment
these effects. 5HT depletion reliably blocks the effects of SSRIs in (Dziedzicka-Wasylewska et al., 1997), suggesting that the interac-
animal models of depression without affecting the action of NRIs. tion between DA antagonists and antidepressants (Fig. 2) occurs
Studies of NA depletion show essentially the opposite pattern: via a primary action of antidepressants in hippocampus, amygdala
blockade of the action of NRIs but not SSRIs (Lucki and O’Leary, or PFC, leading to changes in the output to the nucleus accumbens
2004), albeit that these effects are less consistent: for example, that in turn modulates DA D2 receptor sensitivity (Willner, 1997b).
mutant mice deficient in NA do not respond either to NRI or to SSRI
(Cryan et al., 2004), which may reflect the fact that 5HT cells in the 4.2. Post-transductional mechanisms of antidepressant action
DRN are silent in the complete absence of a tonic NA input from
the locus coeruleus, mediated through alpha1-adrenergic recep- Fig. 5 summarizes the post-synaptic mechanisms engaged by
tors (Gartside et al., 1997; Tremblay and Blier, 2006). A limitation an increase in synaptic concentrations of monoamines. With the
of these studies is that they involved acute antidepressant treat- single exception of the 5HT3 receptor, the synaptic actions of
ment in models (the forced swim and tail suspension tests) that 5HT and NA are mediated by their binding to G-protein coupled
are useful as antidepressant screening tests, probably because they receptors that stimulate or inhibit synthesis of the intracellular sec-
mainly reflect the enhanced activity of monoaminergic systems in ond messenger cyclic AMP (cAMP). This leads in turn to changes
the normal rodent brain, but have low validity as models of depres- in the activity of protein kinase A (PKA), which interacts in a
sion (Willner and Mitchell, 2002a). However, similar effects have complex fashion with other protein kinases (MAPK and CaMK),
been described in the more valid CMS model: chronic treatment resulting in changes in the activity of the nuclear transcription
with the SSRI citalopram and the NRI desipramine reversed dis- factor cAMP response element-binding (CREB). Following chronic,
tinct cognitive impairments in animals subjected to CMS, which but not acute, administration, antidepressants of all classes have
were blocked by acute antagonism of transmission at 5HT and NA been found to increase the expression of CREB in the hippocampus
synapses, respectively (Furr et al., 2011; Bondi et al., 2010). (Nibuya et al., 1996). The control of CREB expression is relatively
While actions of different classes of antidepressant drug origi- well understood, but the precise mechanisms by which antide-
nate at different sites, the various cortical or subcortical routes to pressants stimulate CREB remain uncertain (Tardito et al., 2006).
antidepressant action converge at a point downstream from the pri- However, three relevant observations are that: the critical recep-
mary actions at 5HT and NA synapses. For example, after chronic tor for 5-HT-mediated effects is the 5HT1A receptor (Santarelli
treatment, antidepressants of all classes increase the responsive- et al., 2003); 5HT1A-receptor activation inhibits PKA but stimu-
ness of D2 DA receptors in the nucleus accumbens, the terminal lates MAPK (Polter and Li, 2010); and inhibition of the MAPK
integrative area of the mesolimbic DA system (a specific effect pathway blocked the effects of both serotonergic and noradren-
that is not seen in other brain regions). The functional signifi- ergic antidepressants in the forced swim test (Duman et al., 2007).
cance of this effect has been investigated using the CMS model. According, it appears that the MAPK pathway is responsible for the
In a series of studies, the recovery from CMS-induced anhedo- effect of antidepressants to increase CREB expression. Conversely,
nia following antidepressant treatment was reversed by an acute CREB expression was decreased, in the dentate gyrus of the hip-
challenge with low doses of D2 receptor-blocking drugs. This low pocampus specifically, in anhedonic rats subjected to CMS (Grønli
level of D2 antagonism reversed the action of SSRI, NRI or mixed et al., 2006). CREB expression has also been found to be decreased
5HT-NA uptake inhibitors, while having no effect in non-stressed in post-mortem studies of depressed suicide victims; conversely,
or non-antidepressant-treated animals (Muscat et al., 1990, 1992; CREB expression was increased by pre-mortem antidepressant
Sampson et al., 1991). Significantly, like the effects of 5HT and NA treatment (Pittenger and Duman, 2008). In animal models, over-
synthesis blockade described above, the effect of D2 receptor block- expression of CREB produces antidepressant-like effects (Blendy,
ade is also seen in patients: as already seen in Fig. 2, administration 2006). However, these effects are region-specific, as desipramine
of a low dose of the D2 blocker sulpiride caused a profound return and fluoxetine both decreased CREB function as well as CREB-
of depressed mood in depressed patients successfully treated with regulated target gene expression in the nucleus accumbens, and
SSRIs (Willner et al., 2005). blocked the effects of stress (Chartoff et al., 2009), while inhibition
As antidepressants do not affect DA uptake (except in the PFC, of CREB in this brain region produces an antidepressant-like effect
where the effect is seen reliably only on acute administration, and (Newton et al., 2002). Thus CREB appears to have opposite func-
reflects the fact that in PFC, DA is cleared from synapses primarily tions in the hippocampus (antidepressant) and nucleus accumbens
by the NA transporter: Tanda et al., 1996), the DA-antidepressant (pro-depressant).
interaction probably occurs indirectly. As illustrated in Fig. 3, the CREB is a member of a much larger family of structurally and
ascending 5HT (and NA) projections terminate within the three functionally related transcription factors, which also include iso-
structures that have been identified as the predominant players forms of cAMP response element modulator (CREM), ICER and
in our current understanding of the neuronal network underly- activating transcription factors (ATFs). These proteins can homo-
ing depression, the hippocampus, amygdala and PFC, all of which dimerize or heterodimerize to regulate gene transcription. ATF2,
project to the nucleus accumbens, where their projections ter- ATF3, and ATF4 are all induced by stress in the nucleus accumbens
minate in partially overlapping fields, and to some extent on and their viral-mediated over-expression in the accumbens induces
the same cells (particularly within the nucleus accumbens shell: emotion-related behaviours: over-expression of ICER or of ATF2
French and Totterdell, 2002, 2003), that are innervated by the elicits depressive-like behaviours while over-expression of ATF3
P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371 2347

Fig. 5. Cellular and molecular mechanisms of antidepressant action (red arrows). Antidepressants act by increasing levels of 5HT (blue circles) and/or NA (green circles),
which is generally achieved via inhibition of the 5HT and/or NA transporters. Serotonin then binds to 5-HT2B positive (stimulatory) auto-receptors located in the raphe
nuclei and to 5-HT1A, 5-HT4 and 5-HT7 post-synaptic receptors within the hippocampus, as well as to 5-HT1A receptors located on hippocampal neural progenitor cells.
NA binds to ␣2 auto-receptors located in the locus coeruleus, and to hippocampal ␤-adrenergic receptors. Within the post-synaptic hippocampal neurons, activation of
both serotonergic and noradrenergic receptors elicits activation of a cytoplasmic cascade of intracellular messengers. ␤-adrenergic receptors as well as 5-HT4 and 5-HT7
receptors are coupled to Gs, so their stimulation will sequentially activate cAMP and PKA. On the other hand, 5-HT1A receptors are coupled with Gi or Gq, which activates
Ca2+-dependant cascades as well as MAPK. All of these pathways lead to phosphorylation of CREB in the nucleus of the cell, which induces transcription of the BDNF gene
into pro-BDNF. In the cytoplasm, pro-BDNF will mature into mBDNF which is then trafficked to dendrites and axons. Once released, mBDNF binds to TrkB receptors located
on neural progenitor cells, which will contribute to the maturation of these cells and their differentiation into new hippocampal neurons. The following actions have been
shown to be essential for the neurogenetic and depression-relevant behavioural effects of SSRIs: stimulation of 5HT2B auto-receptors; inhibition of synaptic 5HT uptake;
stimulation of post-synaptic 5HT1A receptors, and stimulation of MAPK, CREB, and BDNF. In parallel, antidepressants inhibit membrance pumps such as MDR-PGP and
so increase access of Gc to the brain, resulting in a raised intracellular level of glucocorticoids (Gc), which bind to glucocorticoid receptors (GR). (These effects have been
demonstrated in embryonic stem cells. The figure assumes that a similar process occurs in neural progenitors. It is unknown whether this also takes place in mature neurons.)
Upon binding, GR activate PKA, leading to increased neurogenesis (probably via the CREB-BDNF pathway). GR also translocate to the nucleus, where the activated receptor
can activate or repress transcription of specific genes, which inter alia causes a resensitization of GR.

and ATF4 has the opposite effect (Green et al., 2008). Induction of Zetterström et al., 1998). These effects parallel the clinical effects,
CREB in the basolateral amygdala produced an antidepressant-like as increased BDNF is observed only after chronic antidepressant
response in the learned helplessness model, similar to the effect in treatment, with the exception of ECT, where the effect is seen more
the hippocampus: (Wallace et al., 2004). rapidly.
CREB regulates the activity of a number of genes, with con- The BDNF gene contains several exons that can be alterna-
sequent changes in the production of several proteins, the most tively spliced to form different mRNA transcripts, all coding for an
studied of which is brain-derived neurotrophic factor, BDNF. Like identical BDNF protein. Social defeat induces a severe decrease in
CREB, hippocampal BDNF and BDNF mRNA levels are decreased hippocampal BDNF mRNA levels that is mediated via the decreased
by chronic stress, and increased by antidepressant treatment, expression of only two of the transcripts (III and IV). Interest-
although an increase in the BDNF protein is not always seen (Nibuya ingly, the effect of chronic imipramine to oppose the decrease
et al., 1995; Russo-Neustadt et al., 1999; Coppell et al., 2003; in hippocampal BDNF mRNA is also mediated via the same two
Molteni et al., 2006; Duman and Monteggia, 2006; Castrén et al., transcripts. The effect of social defeat relates to an increase in
2007; Martinowich et al., 2007). Also like CREB, BDNF levels are histone H3-K27 dimethylation at the Bdnf P3 and P4 promoters
decreased in the brains of suicide victims studied post-mortem, with no change at the other Bdnf promoters. However, chronic
but increased by pre-mortem antidepressant treatment (Karege imipramine does not reverse this effect as it increases BDNF expres-
et al., 2005; Pittenger and Duman, 2008). An increase in hippocam- sion via a different mechanism, H3-K9 dimethylation that is not
pal BDNF mRNA levels is also observed after electroconvulsive altered by chronic stress (Tsankova et al., 2006). This means that
shock treatment (Nibuya et al., 1995; Duman and Vaidya, 1998; chronic imipramine and chronic stress act via different mechanisms
2348 P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371

at the level of histone dimethylation. This again suggests that a of BDNF also increases survival of newborn hippocampal cells
depression-like signature is still present in the hippocampus after (Schmidt and Duman, 2010). Chronic antidepressant treatments
successful antidepressant therapy. also increase hippocampal neurogenesis. This effect occurs only in
The response to antidepressant treatment in the forced swim the dentate gyrus of the hippocampus, but not in other regions
test was blocked in mutant mice with full or forebrain-specific that support neurogenesis (Surget et al., 2008). It is seen not only
impairment of either BDNF or its receptor, Trk-B (Saarelainen et al., with commercially available antidepressants, but also with puta-
2003; Monteggia et al., 2004; Deltheil et al., 2008; Ibarguen-Vargas tive antidepressants from very different pharmacological classes,
et al., 2009). A region-specific knock-down of BDNF in the den- including not only antidepressants that act at monoaminergic tar-
tate gyrus or ventral subiculum of the hippocampus (but not in gets (Malberg et al., 2000; Manev et al., 2001) but also compounds
the CA1 or CA3 fields) also induces depression-like behaviour and acting via non-monoaminergic pathways, such as glutamatergic
confers resistance to the effects of antidepressant drugs such as ligands (Yoshimizu and Chaki, 2004), the synthetic cannabinoid
desipramine and citalopram (Adachi et al., 2008; Taliaz et al., HU210 (Jiang et al., 2005), tianeptine (Czéh et al., 2001; McEwen
2010). Conversely, a single BDNF infusion into the hippocampus et al., 2002), compounds acting on the stress axis such as CRF1
has been reported to produce a long-lasting recovery of depressive- or vasopressin V1b receptor antagonists (Alonso et al., 2004),
like behaviours in the learned helplessness model of depression, or a melanin-concentrating hormone (MCH) antagonist (David
comparable to that produced by chronic systemic antidepressant et al., 2007). This stimulant effect of antidepressant-like treat-
treatment (Shirayama et al., 2002). In contrast to these effects in ments on hippocampal neurogenesis was also observed with
the hippocampus, BDNF increases depression-like behaviour when non-pharmacological therapies, such as electroconvulsive therapy
injected into the VTA, while inhibition of BDNF signaling in the (Malberg et al., 2000) or vagus nerve stimulation (Revesz et al.,
nucleus accumbens has the opposite effect (Eisch et al., 2003; 2008). Increased hippocampal neurogenesis is also seen following
Berton et al., 2006a,b; Krishnan and Nestler, 2008). So like CREB, other treatments that have antidepressant-like effects in rodents,
the effects of BDNF are region-specific, with opposite effects in such as physical excercice (van Praag et al., 1999; Kronenberg et al.,
the hippocampus and nucleus accumbens. However, chronic sys- 2003) or rearing in enriched environments (Kronenberg et al., 2003;
temic BDNF treatment has antidepressant-like effects (Schmidt and Veena et al., 2009; Schloesser et al., 2010). The stimulant effect of
Duman, 2010), suggesting that the effect in the hippocampus pre- antidepressants on neurogenesis that is observed in adult rodents
dominates. is not seen in very young or very old animals (Hodes et al., 2009;
CREB also regulates the expression of another neurotrophin, Couillard-Despres et al., 2009), and this might be relevant to antide-
vascular endothelial growth factor (VEGF) (Lee et al., 2009). Simi- pressant treatment resistance in children and older adults. Inbred
lar to BDNF, VEGF expression in the hippocampus is decreased by mouse strains have been reported to vary in the extent to which
stress and increased by chronic antidepressant treatment; block- neurogenesis was stimulated by fluoxetine, and this effect was seen
ade of VEGF receptors antagonized the behavioural effects of only in strains that also showed a positive behavioural response to
antidepressants; and VEGF infusion into the lateral ventricle had antidepressant treatment (Miller et al., 2008). Significantly, these
antidepressant-like effects (Warner-Schmidt and Duman, 2007; effects require chronic treatment consistent with the clinical situ-
Pittenger and Duman, 2008). The fact that antagonism of hippocam- ation. Antidepressant treatment has also been reported to increase
pal expression of either BDNF or VEGF was sufficient to block neurogenesis in the human hippocampus, studied post-mortem
antidepressant action suggests that an increased stimulation of (Boldrini et al., 2009).
both of these neurotrophins may be required. There is evidence Interestingly, neuronal progenitors and immature neurons
suggesting that hippocampal levels of several other neurotrophins express the BDNF receptor TrkB, and conditional deletion of the
are also increased by chronic antidepressant treatment, (Schmidt TrkB gene in these cells both decreases their proliferation and
and Duman, 2007): some of these effects might similarly turn out survival, and suppresses the effects of antidepressant dugs on pro-
to be obligatory. liferation and survival (Sairanen et al., 2007; Bergami et al., 2008;
Li et al., 2008). This demonstrates that BDNF mediates the stimu-
4.3. Neurogenesis lation of neurogenesis by antidepressants. Similarly, central VEGF
infusion increases cell proliferation and the survival of immature
Neurotrophins have multiple roles in the developing nervous neurons (Jin et al., 2002; Warner-Schmidt and Duman, 2007). Fur-
system and also in the adult brain; they influence synaptic plas- ther, VEGF knockout mice have lower levels of hippocampal cell
ticity, the morphology of dendritic spines, and the proliferation, proliferation and immature neurons (Sun et al., 2006), while viral-
differentiation and survival of neurons (Kandel, 2001; Schmidt and mediated VEGF over-expression has the opposite effect (Cao et al.,
Duman, 2007). In the adult mammalian brain, new neurons are 2004). It is therefore likely that the involvement of VEGF in the
found mainly in the dentate gyrus of the hippocampus, and to effect of antidepressants is also via its effect on neurogenesis.
a lesser extent the olfactory bulbs, although some evidence sug-
gests that neurogenesis may also occur in other areas, including 4.3.1. The pathway from synapse to neurogenesis
the hypothalamus, amygdala, neocortex, piriform cortex, substan- Stimulation of monoamine transmission is required for antide-
tia nigra and striatum (Gould, 2007). In the subgranular zone of pressants to increase neurogenesis. Behavioural and neurogenetic
the dentate gyrus, new adult neurons originate from a process that effects of SSRIs are absent in mice after genetic or pharmacological
includes different stages: first, Type-1 cells (radial-glia-like stem inactivation of the 5HT2B receptor, which is a positive (stimu-
cells) divide, which enables them to maintain themselves (by pro- latory) autoreceptor situated on 5HT cell bodies, the absence of
ducing other Type-1 cells) as well as to generate Type-2 daughter which prevents SSRIs from increasing synaptic 5HT concentra-
cells (neural progenitor cells). These divide symmetrically produc- tions (Diaz et al., 2012). Under normal conditions, serotonergic
ing Type-3 cells (neuroblasts) which migrate into the granule cell antidepressants increase neurogenesis by increasing the impact
layer. Type-4 cells are post-mitotic: they extend axons toward the of 5HT at 5HT1A receptors: neurogenesis is positively correlated
CA3 region, leading to the development of mature granule cells. with levels of 5HT in the hippocampus (Brezun and Daszuta, 1999,
Chronic administration of BDNF directly into the hippocam- 2000; Banasr et al., 2004; Radley and Jacobs, 2002), and the effect
pus increases neurogenesis (Scharfman et al., 2005), which, as of the SSRI fluoxetine was absent in mutant mice lacking the
discussed earlier, is suppressed by chronic stress (Samuels and 5HT1A receptor (Santarelli et al., 2003). These mice continued to
Hen, 2011; Hanson et al., 2011). Chronic peripheral injection respond to tricyclic antidepressants, presumably acting through
P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371 2349

noradrenergic mechanisms (Santarelli et al., 2003). However, with the effects seen following X-irradiation. This suggests that a
the noradrenergic system is also involved in neurogenesis, as degree of maturation, to the point when the new cells are forming
cell proliferation in the subgranular zone is decreased by NA effective synaptic connections, is necessary for a new cell to con-
depletion while noradrenergic stimulation has the opposite effect tribute to the therapeutic action of antidepressant drugs.
(Kulkarni et al., 2002; Rizk et al., 2006). Further, the proneurogenic While neurogenesis appears to play a critical role in antidepres-
action of imipramine is speeded by blockade of ␣2 -adrenoceptor sant action, blockade of neurogenesis, using X-irradiation or genetic
receptors using yohimbine (Yanpallewar et al., 2010). Conversely, tools, does not in itself induce a depressive phenotype (Santarelli
the ␣2 -adrenoceptor agonists clonidine and guanabenz decrease et al., 2003; Airan et al., 2007; Bessa et al., 2009; Holick et al., 2008;
adult hippocampal neurogenesis through a selective effect on the Revest et al., 2009; Samuels and Hen, 2011; Surget et al., 2008,
proliferation, but not the survival or differentiation of progenitors 2011). This is not particularly concerning, because, as discussed
(Yanpallewar et al., 2010), while ␣2 -adrenoceptor antagonists such above, a depressive diathesis is only activated in the presence of
as dexefaroxan have the opposite effect (Rizk et al., 2006). As these stress. Of greater concern, X-irradiation also did not influence sus-
effects persist in dopamine ␤-hydroxylase knockout (Dbh−/− ) mice ceptibility to stress (Santarelli et al., 2003; Surget et al., 2008, 2011),
(which lack NA) it has been suggested that they are related to the and CMS decreased neurogenesis equally in both CMS-sensitive and
presence of ␣2 -heteroceptors on progenitor cells. CMS-resilient animals (Jayatissa et al., 2010). But again, this would
As illustrated in Fig. 5, taken together, these studies confirm be expected if neurogenesis is involved primarily in repairing dam-
that proneurogenic effect of chronic antidepressant administration age to the hippocampus. One recent study, using a genetic model
operates via (1) inhibition of 5HT or NA uptake, causing (2) stim- that reduced neurogenesis in both the hippocampus and in the
ulation of 5HT1A and noradrenergic receptors with effects on (3) olfactory bulbs, reported that this did not cause a depressive-like
intracellular second messengers and (4) protein kinases, leading to behaviour per se, but did increase stress-responsiveness (a pro-
increased expression of (5) CREB and (6) BDNF and VEGF, which (7) longed increase in corticosterone) and depressive-like behaviours
stimulate neurogenesis. when confronted with an acute stress (30 min restraint) (Snyder
et al., 2011). However, the stress-responsiveness seen in this study
4.3.2. Functional significance could be related to peripheral side effects associated with both the
The causal role of neurogenesis in antidepressant action has genetic model (Bush et al., 1998), and the drug used to initiate the
been investigated in a series of studies using the CMS model, with suppression of neurogenesis (Wei et al., 2011). The opposite effect
two behavioural endpoints, feeding in a novel environment and was not found using a genetic model in which adult neurogenesis
grooming: both of these behaviours are suppressed by CMS and was increased (Sahay et al., 2011).
reinstated by chronic, but not acute antidepressant treatment. Neu- In addition to increasing neurogenesis, chronic antidepressant
rogenesis was blocked by X-irradiation of the hippocampus, using treatment has also been reported to increase dendritic arborisation
lead screens to shield other brain areas. X-irradiation of the hip- in the hippocampus (Wang et al., 2008), and to increase the expres-
pocampus also blocked the reversal of CMS-induced behavioural sion of plasticity-related proteins, including neural cell adhesion
changes by either the SSRI fluoxetine or the tricyclic antidepressant molecule (NCAM) in the hippocampus and medial PFC. The study of
imipramine (Santarelli et al., 2003; Surget et al., 2008, 2011). These Bessa et al. (2009) found that, while MAM abolished neurogenesis,
effects were also seen for physiological endpoints such as altered it had no effect on other trophic effects of antidepressants. Specifi-
negative feedback over the HPA axis: CMS attenuated the suppres- cally, CMS decreased the volume of both hippocampus and medial
sion of glucocorticoid release in a dexamethasone suppression test, PFC, which was associated with atrophy of the dendrites of pyrami-
and fluoxetine restored normal HPA function in intact animals but dal cells, involving shortening and loss of dendritic spines. CMS also
not in irradiated animals, demonstrating that the action of fluox- decreased expression of the genes coding for NCAM in hippocam-
etine required newly born hippocampal cells (Surget et al., 2011). pus and PFC, and another plasticity protein, synapsin 1 in prefrontal
While X-irradiation blocked behavioural effects that are only seen cortex. All of these changes were normalized by chronic antidepres-
after chronic antidepressant treatment, it did not affect behavioural sant treatment. This study provides support for the importance of
effects of fluoxetine that are typically seen after acute administra- structural changes in the hippocampus and medial PFC in depres-
tion (e.g. decreased immobility in the forced swim test) (David et al., sion and antidepressant action. Indeed, since newly born neurons
2009). The opposite manipulation, amplification of neurogenesis, need to mature for several weeks before their presence is essential
did not produce an antidepressant-like phenotype in the forced for antidepressant action (Mateus-Pinheiro et al., 2011), it is likely
swim and novelty-suppressed feeding tests (Sahay et al., 2011), that these other trophic actions of antidepressants mediate their
but this experiment was conducted in ‘non-depressed’ animals: the short-term behavioural effects.
effects of this manipulation in, for example, the CMS model have
not yet been explored. 4.4. HPA axis
Discrepant results have been reported when blockade of
neurogenesis was achieved pharmacologically, using the drug, As discussed earlier, a critical event in the breakdown of
methylaxoxymethanol (MAM), rather than by X-irradiation. This homeostatic regulation under conditions of prolonged stress and
study examined three behavioural endpoints in the CMS model, consequent glucocorticoid over-exposure is a decrease in the
anhedonia (decreased sucrose preference), increased immobility in responsiveness of glucocorticoid receptors (GR) on hippocampal
the forced swim test, and novelty-suppressed feeding. MAM abol- granule cells, leading to disinhibition of the HPA axis and a fur-
ished the effect of antidepressant drugs on neurogenesis, but did ther increase in corticosteroid stimulation (Raison and Miller,
not interact with antidepressants on any of the behavioural meas- 2003). Antidepressants normalize HPA function, and part of this
ures (Bessa et al., 2009). However, this is probably related to the effect involves a restoration of GR responsiveness. This effect
experimental protocol used in this study, as MAM was applied is independent of the monoamine uptake-inhibiting actions of
during the 2 weeks of antidepressant treatment, in contrast to antidepressants, because it is seen in neuronal cell cultures
X-irradiation which was applied several weeks earlier (Santarelli (Okugawa et al., 1999; Anacker et al., 2011a). In embryonic stem
et al., 2003; Surget et al., 2008, 2011). When MAM was adminis- cells, antidepressants influence a number of intracellular signalling
tered during the 2 last weeks of the antidepressant treatment, but systems in vitro and also inhibit cellular membrane pumps such as
behavioural testing was delayed until 4 weeks later, the antidepres- the multidrug-resistant glycoprotein (MDR-PGP) (Yau et al., 2007).
sant effect was blocked (Mateus-Pinheiro et al., 2011), consistent Inhibition of MDR-PGP by antidepressants increases the access of
2350 P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371

glucocorticoids to the brain, increasing intracellular glucocorticoid However, while there is increasing evidence that normalization
concentrations (Fig. 5), and leading, after some days of treatment, to of cognitive biases to emotional stimuli may represent an important
an increase in the sensitivity of GRs (Carvalho and Pariante, 2008). early stage in the normalization of the frontal brain circuits that are
The effect of chronic SSRI treatment to increase neurogenesis in disordered in depression, evidence that this serves as a platform
vitro was blocked by a GR antagonist, and was only seen in the pres- for relearning new adaptive associations, and that this relearning
ence of a GR agonist (Anacker et al., 2011b). These findings suggest is the critical event in antidepressant action, is presently lacking.
that restoration of GR responsiveness by chronic antidepressant Two predictions can be derived from this hypothesis, neither of
treatment, and their subsequent stimulation by glucocorticoids, which is well supported. First, if experience of altered affective
may be critical for stimulation of neurogenesis by antidepressants. biases is essential for antidepressants to improve mood, then rapid
However, there is inconsistency between these in vitro findings and onset of antidepressant action should not be possible. However,
the effects of antidepressants in vivo, since the in vitro effects are as discussed below, at least four treatments are known that do
mediated by an increase in intracellular levels of PKA (Anacker et al., have a rapid onset: DBS (Hamani et al., 2011), ECT (Sienaert, 2011),
2011b), but, as discussed earlier, in vivo SSRIs increase transmis- intravenous ketamine (Zarate et al., 2006a,b), and REM-sleep depri-
sion at 5HT1A receptors, which decreases PKA levels (Polter and vation, which is effective after a single night, albeit only until the
Li, 2010). Consequently, the relevance to their clinical action of next sleep (Dallaspezia and Benedetti, 2011). Assuming that at least
these monoamine-independent effects of antidepressants, which some of these treatments resemble antidepressant treatment in
have been described in embryonic stem cells in vitro, is at present eliciting a positive affective bias, there is a potential associative
unclear. Alternatively, these actions may indeed be relevant, but in account of these effects: just as tryptophan depletion can acti-
brain regions other than the hippocampus, such as the PFC. vate a depressive schema by virtue of a prior learned association
between negative information-processing biases and low mood, so
antidepressant treatments might activate a non-depressed schema
4.5. Why are antidepressants slow to act? by virtue of a lifelong association between positive information-
processing biases and high mood. But this account would apply
The traditional account of the slow onset of antidepressant equally to conventional antidepressants, and does not depend on
action is that this relies on neurobiological effects that develop experience gained while under the influence of antidepressant
slowly under chronic drug treatment, such as increases or decreases drugs. The second prediction is that later episodes of depression
in the sensitivity of populations of neuroreceptors, and/or intrin- should be easier to treat than first episodes. This follows because the
sically slow processes that depend on changing patterns of gene association between drug-induced positive biases and improved
expression, such as neurogenesis and synaptic remodelling. More functioning has already been learned during the first episode,
recently, an alternative ‘cognitive neuropsychological’ model has and therefore should be easier to activate. However, as discussed
been proposed, which posits that antidepressants cause relatively below, far from being easier to treat, later episodes of depression
rapid changes in the emotional information-processing biases, become increasingly treatment-resistant. With the exception of
but improvements in mood are slow because the patient must these ‘natural experiments’, the central tenet of the cognitive neu-
first relearn emotional associations (Harmer, 2008; Harmer et al., ropsychological model of antidepressant action, the critical role of
2009a; Clark et al., 2009; Pringle et al., 2011; Roiser et al., 2012). experience under drug treatment, has not yet been tested.
There is good evidence that some emotion-processing effects
of antidepressants may occur rapidly. Single doses of SSRI or NRI 4.6. Treatment implications
antidepressants have been reported to increase the processing of
positive information in both healthy volunteers (Harmer et al., The model developed in this review is summarized in
2003a,b; Browning et al., 2007) and depressed patients (Harmer Figs. 4 and 6. As already described, Fig. 4 summarizes the processes
et al., 2009b). Single doses of SSRIs do not reverse negative that follow an appraisal that stress is unmanageable, leading, via
information-processing biases: on the contrary, responses to threat prolonged exposure to high levels of corticosteroids, to a compro-
are increased (Harmer et al., 2003a; Browning et al., 2007). This mised hippocampal-cingulate system with an increase in activity
probably results from a potentiation of 5HT release in the amyg- of the ventral PFC and amygdala and a decrease in activity of the
dala, which, as discussed above, is a critical element of the response dorsal PFC and nucleus accumbens, which mediate an increased
to stress (Maier and Watkins, 2010), and may correspond to the sensitivity to aversive events and a decreased response to rewards
increase in anxiety reported by many patients in the first few days (anhedonia), respectively. The effects of currently used monoamin-
of SSRI treatment, which counteracts the antidepressant effect of ergic antidepressants (Fig. 6) are mediated primarily by stimulation
5HT release in the hippocampus. However, a week of antidepres- of serotonergic and noradrenergic synapses in the hippocampus,
sant administration to volunteers did decrease threat responses which increases the production of neurotrophins and neurogene-
(Harmer et al., 2004), and also decreased amygdalar activation by sis leading to reorganization and/or repair of the stress-induced
threat stimuli (Harmer et al., 2006). Though requiring repeated drug morphological damage, involving recently born neurons, and a
treatment, these changes occur earlier than changes in mood are rebalancing of activity in forebrain circuits, with a normalizing
typically reported. decrease in the impact of aversive events and increase in the impact
Tryptophan-depletion studies provide a second line of sup- of rewards. These two sets of processes are largely independent:
port for the cognitive neuropsychological account of mood antidepressants reinstate normal functioning but do not directly
disturbance. As discussed above, tryptophan depletion does not counteract the effects of stress.
generally impair mood in people who have never been depressed. Like any model of this kind, this is, of course, an over-
However, tryptophan depletion does produce a variety of nega- simplification: for example, stress and antidepressants may also
tive information-processing biases in never-depressed volunteers affect other structures directly, in particular, the PFC, amygdala,
that resemble those reported in depressed patients (Harmer, nucleus accumbens and habenula. Antidepressant effects have
2008). As described above, it has been argued that during an been reported following direct injection of antidepressants into the
episode of depression, an association is learned between these amygdala (Garrigou et al., 1981; Duncan et al., 1986; Shirayama
information-processing biases and low mood, such that subse- et al., 2011) and PFC (Sherman and Petty, 1980), or following
quently, tryptophan depletion may evoke low mood via this interventions targeted at the nucleus accumbens (Corrigan et al.,
association (Harmer, 2008; Robinson and Sahakian, 2008). 2000; Krishnan et al., 2007), and DBS is clinically effective when
P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371 2351

Fig. 6. How antidepressants restore the affective information-processing brain circuitry (cf. Fig. 3) that has been disrupted in depressed patients: the example of a SSRI. First
(step 1), the antidepressant increases serotonin concentrations (yellow points) in the median raphe nucleus (RN) as well as in the hippocampus (HPC). This stimulates the
release of neurotrophins such as BDNF (step 2) and hippocampal neurogenesis (step 3). As the new hippocampal neurons develop they increase not only the function of the
hippocampus, but also the function of hippocampal outputs (step 4). This modifies the function of areas to which the outputs of the hippocampus project (step 5) and thus
restores most parts of the system to an almost normal state. See legend to Fig. 3 for other abbreviations.

applied to the prefrontal cortex (Hamani et al., 2010, 2011), nucleus The possibility of achieving antidepressant effects at multiple
accumbens (Schlaepfer et al., 2008; Bewernick et al., 2010; Grubert points within the depression circuitry provides a framework for
et al., 2011) or habenula (Sartorius et al., 2010). As the habenula understanding both the mechanism of action of many putative
is responsive to light and shows intrinsic circadian rhythmicity, antidepressant drugs and their limitations. We consider here the
this is also a promising candidate for involvement in the antide- prospects for agents that promote neural repair, neuroprotective
pressant effects of light therapy (in seasonal affective disorder) drugs, and anti-stress agents.
and REM (rapid eye movement) sleep deprivation (Dallaspezia and As discussed above, agents that promote neural repair includ-
Benedetti, 2011). While current antidepressants are thought to act ing CREB, BDNF and other trophic factors such as VEGF, have
primarily within the hippocampus, it appears that actions at any been reported to show antidepressant-like effects when adminis-
point within the circuitry associated with depression may in prin- tered within the hippocampus (Blendy, 2006; Warner-Schmidt and
ciple promote recovery. Duman, 2007; Pittenger and Duman, 2008). Histone deacetylase
However, an implication of the same distributed circuitry is that inhibitors are another class of drugs that promote the expression
drug actions may be inconsistent at different sites within it, and of plasticity-related gene expression, and these drugs have been
this creates both problems and opportunities. As discussed above, reported to elicit antidepressant-like effects when infused within
it appears that stress and depression are associated with high levels either the hippocampus or nucleus accumbens (Covington et al.,
of 5HT activity in amygdala and PFC, but low levels in hippocam- 2009, 2011). (Peripheral administration of these drugs has not
pus (Lowry et al., 2008; Elizalde et al., 2010; Rozeske et al., 2011). yet been studied, but might be expected to cause extensive side-
Conversely, the effect of SSRIs to increase 5HT activity is benefi- effects.) Estrogens, which have frequently, but inconsistently, been
cial in hippocampus, but detrimental in amygdala and PFC, and reported to show clinical antidepressant effects and to potentiate
this may limit the effectiveness of SSRIs. The addition of a NRI the effects of conventional antidepressants, also have neurotrophic
component or a DA reuptake inhibitor further increases 5HT lev- actions and increase the expression of BDNF within the hippocam-
els in hippocampus but restricts 5HT levels in the PFC, probably pus (Estrada-Camarena et al., 2010). However, as noted above,
because NA uptake inhibition also potentiates DA in PFC but not while BDNF has antidepressant-like effects in the hippocampus,
elsewhere (Weikop et al., 2007a,b), which might explain the greater it causes exactly the opposite, pro-depressive effect when admin-
efficacy of SNRIs such as venlafaxine. This example illustrates a istered within the nucleus accumbens (Eisch et al., 2003; Berton
theoretical limitation for antidepressant drug development: that et al., 2006a,b), which potentially limits the usefulness of strategies
peripherally administered agents are not anatomically specific in that target BDNF directly, albeit that peripheral administration of
their actions, and actions remote from the site of antidepressant BDNF to rats had antidepressant-like effects, including reversal of
action may counteract the benefits. Perhaps problems of this kind CMS-induced anhedonia (Schmidt and Duman, 2010).
could be circumvented by using nano-technology to deliver drugs Neuroprotective agents have also been proposed as antide-
to specific brain areas, but this intriguing prospect is not yet within pressants. Because of its high level of oxygen consumption, the
reach. brain is at particular risk from the toxicity caused by reactive
2352 P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371

oxygen species: high levels of oxidative stress have been reported of those who do recover relapse within 6–12 months (Warden
in post-mortem studies of the depressed brain, and antidepressants et al., 2007; Holtzheimer and Mayberg, 2011): this, alongside the
decrease oxidative stress. However, while many anti-oxidants are slow onset of antidepressant action, provides the major impetus for
under preclinical investigation, including several of plant origin, the continuing need for improved treatments. For example, in the
there is minimal evidence that anti-oxidants are clinically effective largest open-label study of the effectiveness of antidepressant ther-
as antidepressants (Ng et al., 2008). The situation is similar for the apy, the Sequenced Treatment Alternatives to Relieve Depression
neurosteroids dihydroepiandosterone (DHEA) and DHEA sulphate, (STAR*D) trial, fewer than one-third of patients achieved remission
which include neuroprotection among their many neurochemical with an adequate trial of a standard antidepressant after up to 14
actions: a relationship between plasma concentration and mood weeks of treatment. Some further improvement may be obtained
has been reported, but the literature regarding their clinical efficacy by switching non-responders to another antidepressant from a dif-
in depression is highly inconsistent (Maninger et al., 2009). ferent class, but even after treatment with two antidepressants and
Many putative antidepressants act primarily as anti-stress 24 weeks of treatment, only half of the patients in the STAR*D study
agents. For example, HPA antagonists counteract the neuroen- remitted (Trivedi et al., 2006; Rush et al., 2011). The intractabil-
docrine effects of stress, by inhibiting glucocorticoid synthesis, ity of this problem raises the question of whether the efficacy of
antagonising receptors for corticosteroids, vasopressin or CRF, or antidepressant treatments is intrinsically limited. For example, if,
blocking the actions of inflammatory cytokines, release of which is as suggested, the key event in the onset of depression involves
stimulated by corticosteroids (Holsboer, 2000). All of these drugs a computation within the PFC (concerning the controllability of
have shown some antidepressant-like activity in animal models adverse events), the mechanisms encoding the elements of this
of depression (e.g. Koo and Duman, 2009). A cytokine antagonist decision may not be differentiated neurochemically in ways that
has shown antidepressant activity in patients treated for psoria- would render them differentially susceptible to drug treatment.
sis, but this effect is difficult to interpret because it accompanied Happily, the evidence that drug treatment-resistant depressions do
an improvement in their physical condition (Tyring et al., 2006). respond to ECT, DBS and, as discussed below, intravenous ketamine,
HPA antagonists have also shown clinical antidepressant activity, refutes this nihilistic prospect by demonstrating that treatment
but so unreliably and with unacceptable side effects, that clinical resistance can in fact be overcome.
development was stopped (Schüle et al., 2009; Zorrilla and Koob,
2010). This outcome is particularly disappointing because patients 4.7.1. Pharmacokinetics, misdiagnosis and comorbidity
with elevated HPA activity as assessed by the cortisol response to a A major cause of treatment resistance, which is clinically impor-
combination of dexamethasone and CRH (Heuser et al., 1994) have tant but neurobiologically trivial, arises because antidepressant
a poor prognosis (Paslakis et al., 2010). drugs may have difficulty in reaching the brain in sufficient con-
A part of the reason so many novel compounds fail to progress centrations. Many single nucleotide polymorphisms (SNPs) of liver
from animal models of depression into the clinic is that, in order cytochrome P450 proteins have been identified that alter the
to be marketed, a new antidepressant would need to demonstrate metabolism of antidepressants (Brøsen, 2004; Yin et al., 2006). For
superiority to currently available drugs, in terms of onset of action, example, non-responders to the SSRI citalopram were found to
efficacy, or side effect profile. Current drugs have a relatively benign carry the allele of the enzymes CYP2D6 and CYP2C19 associated
side effect profile, so the challenge for a new agent is to decrease with extensive metabolism of citalopram and treatment response
onset latency or to improve on efficacy, either by increasing the was improved by adding an inhibitor of these two enzymes to
therapeutic effect in patients who respond to treatment, or by elic- increase plasma levels of the drug (Bondolfi et al., 1996). Different
iting a response in treatment-resistant patients. Phase 3 studies antidepressants vary greatly in their interaction with cytochrome
to demonstrate such profiles will be lengthy and costly, and ani- P450 enzymes (Preskorn, 1997). The association between antide-
mal models with proven validity in relation to treatment-resistant pressant plasma levels related to genetic differences in liver
patients are lacking. HPA antagonists and neuroprotective agents metabolism has not always been observed (Grasmäder et al.,
have not demonstrated superiority in these respects; rather, the 2004; Peters et al., 2008), but considering that many antidepres-
reverse. This unreliability could, of course, reflect the fact that the sants show non-linear dose–response functions (Perry et al., 1994;
role of stress in depression decreases with successive episodes DeVane and Gill, 1997) this is hardly surprising.
and with the strength of the depressive diathesis (Kendler et al., Antidepressants may also be ineffective because they fail to
2001), and so in many people is minimal: indeed, only around penetrate the blood brain barrier (BBB) sufficiently. The function-
30% of depressed patients show a significant increase in CRF lev- ing of the BBB relies crucially on transporter molecules such as
els (Arborelius et al., 1999). Perhaps anti-stress measures would be P-glycoprotein (P-gp) (Cordon-Cardo et al., 1989; Thiebaut et al.,
more effective early in the course of depression, and particularly, 1987). P-gp is a member of the ATP-binding cassette (ABC) trans-
in first episodes of depression in people with a weak diathesis, in porter protein superfamily (Ambudkar et al., 1999), and is encoded
whom the effect of stress is most pronounced. However, according by the ABCB1 gene, also known as the multidrug-resistance 1
to the current understanding of antidepressant action, the mech- (MDR1) gene (Callen et al., 1987). A SNP in the ABCB1 gene pre-
anism by which anti-stress agents achieved their antidepressant dicts antidepressant resistance in patients receiving drugs that
effect would be to enable natural repair processes to proceed, have been identified as substrates of P-gp, but not drugs that are not
protected from the countervailing effect of stress. But as antide- substrates of P-gp such as mirtazapine. Therefore, putative antide-
pressants are efficient in stimulating neurogenesis and related pressants should be analysed for their P-gp substrate status, which
processes of cell repair even in the presence of stress, it might can be determined by using abcb1 knockout mice.
prove impossible to demonstrate superiority via anti-stress meas- Another cause of treatment resistance that does not relate
ures alone. In principle, protection from stress might augment the directly to the neurobiology of depression is misdiagnosis (Souery
effectiveness of antidepressants, but in practice there is little evi- et al., 1999). For example, depression in older adults may be an
dence for a synergistic effect of this kind. early sign of dementia arising from vascular causes or Alzheimer-
type pathology. Late-onset depression (patients having their first
4.7. Treatment resistance depressive episode after the age of 60 years) is associated with
MRI hyperintensities (Fujikawa et al., 1993; Alexopoulos et al.,
As many as 30% of patients treated with antidepressants fail 1999; Krishnan et al., 1997; see Camus et al., 2004 for a review)
to respond, despite multiple treatment attempts, and around 50% and greater neuropsychological impairment (Salloway et al., 1996;
P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371 2353

Lesser et al., 1996; Alexopoulos et al., 1997), suggesting that than those who were treatment-resistant (MacQueen et al.,
ischemic cerebrovascular disease might increase vulnerability to 2008).
depression in late life by causing structural damage affecting Treatment resistance is also associated with other factors that
prefrontal–subcortical circuits (Alexopoulos et al., 1999). A poor increase the diathesis for depression, including neurotic premor-
response to antidepressant therapy is seen particularly in older bid personality, familial predisposition for affective disorder, or
patients who have high vascular risks (Iosifescu et al., 2005), includ- the experience of multiple loss events (Souery et al., 1999). Treat-
ing burden of cardiac disease (Alexopoulos et al., 2004), and high ment resistance is also associated with several genetic diatheses
levels of cholesterol (Sonawalla et al., 2002; Papakostas et al., 2003, for depression, such as the short allele of the 5HTTLPR gene (Zobel
2004). This has also been demonstrated in an animal model: mice and Maier, 2010; Licinio and Wong, 2011), and polymorphisms of
that were fed a high-fat diet became resistant to the antidepressant genes coding for the CB1 receptor (Domschke et al., 2008; Juhasz
effect of fluoxetine in the CMS model (Isingrini et al., 2010). This et al., 2009) and the 5-HT1A receptor (Kato and Serretti, 2010). But
subtype of depression may be related to ischemic lesions causing not all genetic variants associated with antidepressant treatment
cognitive deficits. As antidepressants do not target the pathophy- resistance are related to vulnerability to depression, such as genes
siological lesions shown by these patients, it is hardly surprising coding for a glutamate receptor (dystrobrevin-binding-protein 1
that their ability to improve symptoms may be poor. The same gene: Arias et al., 2009) and the 5HT2A receptor (Horstmann and
applies to depressive features shown by patients with Alzheimer’s Binder, 2009; Licinio and Wong, 2011). There are also some genetic
disease, where depression is frequent and may precede the cog- variants that are associated with treatment resistance for which
nitive dysfunction. In this case, the depressive symptomatology the association with vulnerability has not been evaluated. These
is caused by histopathological lesions in brain areas involved in include the C allele of the −182T/C SNP in the promoter region of
depression, including prefrontal cortex, hippocampus and cingu- the noradrenaline transporter (Shiroma et al., 2010) and a variant
late cortex and would not be improved by therapies targeting of the gene coding for the neurotrophin VEGF (Viikki et al., 2010).
monoaminergic neurotransmission, as the symptoms are caused A high level of activity in the HPA system also predicts treatment
by Alzheimer-type plaques and tangles. A number of studies have resistance. Treatment outcome is associated with polymorphisms
reported that patients who perform poorly on tests of executive of genes coding for proteins involved in HPA functioning, including
function are likely to be poor treatment responders (Dunkin et al., the CRH receptor 1 gene and the hsp90 co-chaperone FKBP5, a part
2000; Baldwin et al., 2004; Gorlyn et al., 2008; Pimontel et al., 2012). of the mature glucocorticoid receptor (GR) heterocomplex that reg-
These findings may to some extent reflect the inclusion of patients ulates the GR response (Kato and Serretti, 2010). Patients who fail
with neuropathological lesions of prefrontal cortex. to show a decrease in HPA activity after 2–3 weeks of treatment
Treatment resistance may also arise because depression occurs are unlikely to recover from depression with prolonged treatment
in the context of other medical disorders, with no primary dys- (Brouwer et al., 2006; Ising et al., 2007), and if they do recover are
function of the brain circuitry related to depression. For example, very likely to relapse (Zobel et al., 2001; Appelhopf et al., 2006).
thyroid dysfunction (Gold et al., 1981) or folate deficiency (Gilbody This suggests, as discussed previously, that a major component of
et al., 2007) can cause depressive symptomatology and treatment antidepressant action is the anti-stress effect, and that additional
resistance, which may respond to thyroid hormone (Abraham et al., actions may be needed for improvements in treatment.
2006) or folate (Fava and Mischoulon, 2009) supplementation. Other studies have explored the association of brain activ-
The factor most strongly associated with antidepressant treat- ity with treatment resistance. As discussed earlier, the most
ment resistance is the presence of comorbid anxiety conditions, convincing evidence points to a relationship between antide-
including generalized anxiety disorder, panic disorder and social pressant resistance and low pretreatment activity in the rostral
phobia (Souery et al., 2007). These conditions are associated with ACC (Mayberg et al., 1997; Pizzagalli, 2011), which was observed
a decreased density of 5HT1A receptors in cortex and midbrain in 19 out of 23 studies included in a recent meta-analysis
(Meyer, 2012). The likely explanation for this type of antidepres- (Pizzagalli, 2011). This marker is a predictor of treatment outcome
sant resistance is that these three anxiety disorders are themselves not only in patients treated with conventional monoaminergic-
difficult to treat pharmacologically (Bystrisky, 2006; Pollack et al., acting antidepressants such as SSRIs, but also in patients treated
2008). It goes beyond the scope of this review to examine this prob- with intravenous ketamine (Salvadore et al., 2011) or by non-
lem, but given the close psychological relationship between anxiety pharmacological modalities such as ECT (McCormick et al., 2007),
and depression, it seems inevitable that the presence of an ongoing transcranial magnetic stimulation (TMS: Kito et al., 2008), or sleep
anxiety disorder would interfere with recovery from depression. deprivation (Wu et al., 1999). As discussed above, this marker is
Consistent with high anxiety, a poor response to antidepressant specific to physical treatments, as the opposite relationship is seen
treatment is also seen in patients with high levels of activity in the with psychological treatments, where a low level of activity in the
amygdala prior to treatment (Saxena et al., 2003). It has also been rostral ACC predicts a better outcome (Pizzagalli, 2011; Roiser et al.,
argued that cases of intense normal sadness are frequently misdi- 2012).
agnosed as depression, where perhaps a response to antidepressant The similarities between the factors associated with vulner-
treatment should not be expected (Horvitz and Wakefield, 2007). ability to depression and those associated with antidepressant
treatment resistance, prompt the question of whether activity in
the rostral ACC, identified as one of the most reliable markers of
4.7.2. Parallels with vulnerability to depression antidepressant treatment resistance, is also associated with vulner-
Setting aside treatment resistance that is associated with ability to depression. A number of observations are consistent with
pharmacokinetic factors, misdiagnosis and comorbidity, most of this hypothesis. For example, depressed patients showed increased
the other factors identified as contributing to treatment resis- rostral ACC activation in response to errors (Holmes and Pizzagalli,
tance are similar or identical to those that confer vulnerability 2008), and the same effect has been observed in non-depressed vol-
to depression. In particular, recurrent depressions are more dif- unteers who carry the short allele of the 5HTTPR gene that confers
ficult to treat than first-episode depressions (O’Reardon et al., vulnerability to depression (Holmes et al., 2010). Also consistent
2007; Souery et al., 2007; Kaymaz et al., 2008; Rush et al., with an association between high rostral ACC activity and vulnera-
2011). This may reflect the neurological damage associated bility to depression are observations that high rostral ACC activity is
with recurrent depression, as suggested by the observation that associated with depressed mood in healthy (non-depressed) chil-
patients who responded to treatment had a larger hippocampus dren (Boes et al., 2008), and with pessimism (Sharot et al., 2007)
2354 P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371

Table 1 and genetic diatheses for depression. Antidepressants are most


Some sources of resistance to antidepressant drugsa .
effective in people with a low depressive diathesis (first episode
Factors unique to treatment resistance and few vulnerability factors), whose depressions are associated
Pharmacokinetics with high levels of recent stress. As depressions become increas-
SNPs of liver cytochrome 450 enzymes
ingly autonomous of stress (later episodes and other vulnerability
Failure to penetrate the blood–brain barrier
Misdiagnosis
factors), conventional monoaminergic antidepressants become less
Vascular or Alzheimer-type pathology effective and different strategies are needed to address treatment
Thyroid insufficiency resistance.
Folate defiency
Bipolar disorder
4.8. Approaches to the treatment of resistant depression
Normal sadness
Comorbidity
General anxiety disorder Resistant depression is typically treated by supplementation
Panic disorder of ongoing antidepressant therapy. The two classic augmentation
Social phobia
strategies are the addition of thyroid hormone (T3) or anti-manic
Post-traumatic stress disorder
Factors common to treatment resistance and vulnerability to depression
drugs (lithium or anticonvulsants). Both are widely used, albeit that
Developmental the evidence for their effectiveness is largely in combination with
Familial predisposition to depression tricyclics and there is less evidence that they enhance the effect of
Childhood trauma newer antidepressants (Connolly and Thase, 2011). There is good
Multiple loss events
evidence that the response to SSRIs can be augmented by a more
Neuroticism
History of previous depressive episodes recent strategy, the addition of a second-generation antipsychotic
Physiological agent, such as quetiapine or aripiprazole (Komossa et al., 2010;
HPA dysregulation Kennedy et al., 2011; Connolly and Thase, 2011), administered at
Decreased hippocampal volume
doses much lower than those that would be used in the treatment
Increased amygdalar activity
High activity in the rostral ACC
of psychosis. At these low doses, second-generation antipsychotics
Genetic act selectively at DA synapses in the VTA, resulting in an increase
Short allele of the 5HTTLPR gene DA release in the nucleus accumbens and PFC, and as partial 5HT2
5HT1A receptor polymorphism receptor antagonists; both actions may contribute to the antide-
CB1 receptor polymorphism
pressant effect (Möller, 2005; Blier and Blondeau, 2011). However,
a
The table summarizes only factors that have been identified in the clinic. Further the benefits of augmentation are small (Gaynes et al., 2012).
examples of commonality between vulnerability and treatment resistance that have
As discussed earlier, a good response to antidepressant drug
been identified in animal models are discussed in the text.
treatment is associated with changes in metabolic activity in the
ventral ACC and nucleus accumbens, which are also produced
and negative emotionality (Santesso et al., 2012) in non-depressed by DBS of these structures. There is some evidence that simi-
adults. lar metabolic changes in the PFC and ACC can be achieved by
Parallels between vulnerability to depression and treatment repeated transcranial magnetic stimulation (rTMS) but it is not yet
resistance have also been observed in animal studies. Per- clear whether rTMS can improve on the effectiveness of antide-
haps unsurprisingly, antidepressant responsiveness was lower in pressant drugs (Dell’osso et al., 2011). Strategies to increase DA
mutant mice expressing low levels of BDNF in the hippocampus activity in the nucleus accumbens are similarly embryonic. A recent
(Adachi et al., 2008; Ibarguen-Vargas et al., 2009). Congenitally study reported an effect on congenital helplessness of the MAO-B
helpless rats are also said to be resistant to antidepressant drug inhibitor deprenyl, which preferentially protects DA rather than NA
treatment (Vollmayr et al., 2004). However, the extent of antide- or 5HT; uniquely, this effect required experience of repeated test
pressant resistance in these animals is unclear, as the data have trials, suggesting some form of learning process that has not been
not been published in full: reponsiveness to fluoxetine has been described for any other antidepressant (Schulz et al., 2010). There is
reported (Shumake et al., 2010), but there are no published compar- some evidence that DA D2/D3 receptor agonists have antidepres-
isons comparing the effectiveness of antidepressants on acquired sant activity (Rakofsky et al., 2009), and triple uptake inhibitors,
and congential helplessness. Antidepressant-like properties in con- which add DA uptake inhibition to the NA and 5-HT uptake inhibi-
genitally helpless rats have been reported for a highly selective tion shown by tricylics and SNRIs, have been confirmed as effective
5HT2A antagonist (Patel et al., 2004), consistent with the clinical antidepressants (Chen and Skolnick, 2007), but again, it has not yet
picture of increased 5HT2A receptors in difficult-to-treat vulner- been established if they have superior efficacy to antidepressants
able patients. The CMS model has been used to make direct with a more restricted spectrum of activity.
comparisons of vulnerability and treatment resistance: similar It remains the case that electroconvulsive therapy (ECT) is the
patterns of gene expression (Wilkinson et al., 2009; Christensen most effective treatment for resistant depression, with remission
et al., 2011) and protein expression (Bisgaard et al., 2007) were rates in excess of 75% after a course of treatment (Sienaert, 2011),
observed between animals that were resilient to CMS, and those though this may include a powerful placebo component, as is also
that developed anhedonia but subsequently responded to antide- the case after the classical antidepressant treatments (Rasmussen,
pressant treatment; these two groups differed in gene and protein 2009; Read and Bentall, 2010). It also remains the case that the
expression from animals that developed anhedonia but were mechanism of action ECT is poorly understood (Bolwig, 2011).
treatment-resistant. ECT is more effective than antidepressant drugs in stimulating
Broadly speaking, therefore, it appears that two sets of factors hippocampal neurogenesis (Malberg et al., 2000), but generalized
are largely responsible for antidepressant treatment resistance, seizures cause many other neurochemical changes that are also
which are summarized in Table 1. On the one hand, there are factors associated with antidepressant-like effects (Merkl et al., 2009),
that are not directly related to depression, including pharmacoki- including an increase in striatal DA D1, D2 and D3 receptor bind-
netics, misdiagnosis and comorbidity. On the other hand, there ing (Strome et al., 2007), which is implicated in the mechanism
are factors that are essentially the same as those that are associ- of action by the fact that a polymorphism of the gene coding for
ated with vulnerability to depression, including recurrent episodes, COMT, an enzyme that degrades DA, is associated with the treat-
early life stress and its psychological sequelae, neurotic personality, ment response to ECT (Huuhka et al., 2008). ECT also powerfully
P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371 2355

reverses dysfunctions of the HPA axis, probably via direct actions single injection of ketamine also reversed not only the anhedo-
within the hypothalamus (Bolwig, 2011). nic and other behavioural deficits in the CMS model, but also the
The factors responsible for the superiority of ECT to antide- atrophy of dendritic spines (and consequent loss of synapses) in
pressant drug treatment remain uncertain, though it is probably the PFC (this study did not assess hippocampal synapse loss), and
relevant that unlike antidepressants, ECT is not subject to the associated electrophysiological deficits (Li et al., 2010, 2011a,b;
pharmacokinetic problems such as liver metabolism or blood–brain Duman et al., 2012).
barrier penetration. However, the fact that it is possible to improve The antidepressant effects of NMDA antagonists may appear
on the efficacy of antidepressants (and the reluctance of many clini- paradoxical, given that glutamate levels are decreased in depres-
cians to use ECT) encourages continuing research to develop more sion, and an NMDA antagonist further depresses glutamate
effective drug treatments. One strategy is to build molecules that function. However, ketamine indirectly increases presynaptic glu-
combine neuroreparative effects in the hippocampus with actions tamate release in the PFC, via its interaction with NMDA receptors
elsewhere within the depression circuitry. We briefly review four on GABA interneurons (Moghaddam et al., 1997), which increases
examples. stimulant effects at the other major class of glutamate recep-
tors, AMPA receptors. The clinical response to ketamine was
4.8.1. New targets negatively correlated with a marker of glial cells in the ventral
A compelling case has been made for an involvement of endoge- PFC (Salvadore et al., 2011), perhaps because a decrease in glial
nous cannabinoid systems in depression: depression is associated cells would mean lower clearance of glutamate from the synapse
with decreased levels of circulating endocannabinoids (Gorzalka and therefore a larger increase in glutamate concentration and
and Hill, 2011; McLaughlin and Gobbi, 2012), and single nucleotide a greater stimulation of AMPA receptors. This effect is essential
polymorphisms in the CB1 receptor have been reported to increase for the antidepressant action of ketamine because it is blocked
neuroticism and vulnerability to develop a depressive evidence by AMPA receptor antagonists (Maeng et al., 2008). Accordingly,
following exposure to life stress (Juhasz et al., 2009), as well knockout mice lacking an AMPA receptor sub-unit have been
as increasing resistance to antidepressant treatment (Domschke reported to show a depressive phenotype (Chourbaji et al., 2008).
et al., 2008). Decreasing endocannabinoid signalling produces in Ketamine itself is unlikely to be developed for general antide-
animals a depressive-like phenotype (including anhedonia) and pressant use because it may produce psychotomimetic effects, it
increased HPA activity, while endocannabinoids and CB1 agonists has abuse potential and it is neurotoxic at high doses (Behrens
have antidepressant-like effects (Gorzalka and Hill, 2011; Hill et al., et al., 2007), but these observations do presage the potential devel-
2009; McLaughlin and Gobbi, 2012). There are as yet no clinical opment of new antidepressants based on actions at glutamate
data on cannabinoid agonists, but there is strong clinical evidence synapses. Examples of drugs under development include positive
of depressogenic effects of the cannabinoid antagonist rimona- modulators of AMPA receptors and antagonists of several sub-
bant, including depressive effects in people with no prior history types of metabotropic glutamate receptor (Sanacora et al., 2008;
of depression (Hill and Gorzalka, 2009). Rimonabant also sup- Hashimoto, 2011). Antagonists at mGluR2/3 receptors in particular
pressed positive affective memories and responses to reward in may be promising targets, as the density of these receptors in the
the nucleus accumbens and ventral PFC (Horder et al., 2009, 2010). PFC is increased in depression (Feyissa et al., 2010).
Like antidepressants, facilitation of endocannabinoid signalling A third example is the novel (and expensive) antidepressant
enhances monoaminergic transmission, increases cellular plastic- agomelatine, which, since its introduction less than 10 years ago,
ity and neurotrophin expression in the hippocampus, and decreases has been shown to be among the most effective antidepressant
HPA activity, and there is evidence that these effects depend in drugs available, and has been approved for the treatment of depres-
part on actions within the hippocampus (Gorzalka and Hill, 2011; sion in over 40 countries (Carney and Shelton, 2011). Agomelatine
McLaughlin and Gobbi, 2012). However, there is also evidence that is an agonist at 5HT2C and melatonin receptors, and both of these
at least a part of the antidepressant-like cannabinoid effect may actions appear to be essential for its antidepressant properties:
be mediated at CB1 receptors within the prefrontal cortex, which agomelatine decreases stress-induced glutamate release in the PFC
inhibit the release of glutamate and GABA (Bambico et al., 2007; and increases BDNF in hippocampus and PFC, and hippocampal
McLaughlin and Bambico, 2012). neurogenesis, but agonists at either 5HT2C or melatonin recep-
The glutamate system provides a second example of drugs tors alone do not share these effects (Racagni et al., 2011). But,
that may have wider, or at least, different, anatomical targets differently from other antidepressant drugs, agomelatine also nor-
than current antidepressants. There are many inconsistent reports malizes circadian rhythms by its action at melatonin receptors in
of changes in glutamate levels in the brains of people who are the suprachiasmatic nucleus (Quera Salva et al., 2011; Dallaspezia
depressed (both post-mortem and in vivo), but a consensus is and Benedetti, 2011). As the habenula, like the suprachiasmatic
developing that glutamate levels are decreased in the PFC and ACC; nucleus, also shows intrinsic circadian rhythmicity, and is respon-
changes have also been reported in hippocampus, but less fre- sive to melatonin (Hikosaka, 2010), a potential role for the habenula
quently and less consistently (Sanacora et al., 2008; Hashimoto, in the antidepressant action of agomelatine must be a distinct pos-
2011). Furthermore, genetically modified mice with impaired glu- sibility.
tamate function exhibit a depressive-like phenotype (Tordera et al., Cognitive enhancers present a fourth potential route to
2007, 2011). At the same time, there is extensive evidence for improved antidepressant efficacy, building on the proposal that
antidepressant-like effects of glutamate NMDA receptor antag- antidepressant-induced neuroplasticity might serve as the basis
onists in animal models of depression (Sanacora et al., 2008; for new learning (e.g. Castrén and Rantamäki, 2010). There is some
Hashimoto, 2011), albeit that the weak NMDA antagonist meman- evidence that combined treatment with antidepressants and CBT
tine was ineffective in the only double-blind clinical trial (Zarate has significant advantages over either treatment modality alone
et al., 2006a). Following an initial report on seven patients (Berman (Pampallona et al., 2002; de Maat et al., 2007; Roiser et al., 2012),
et al., 2000), a large number of studies, including two randomized but the differences are unimpressive, and relate in part to treatment
controlled trials (Zarate et al., 2006b; Diazgranados et al., 2010) compliance. (Indeed, one study reported that non-depressed vol-
have reported rapid antidepressant effects following a single intra- unteers treated with an SSRI and a cognitive intervention showed
venous infusion of the NMDA antagonist ketamine, which appear smaller decreases in negative thinking than participants treated
within 2 h, last for around a week, and are seen in patients who with either intervention alone: Browning et al., 2011). An alter-
are refractory to treatment with conventional antidepressants. A native approach contemplates combining CBT with a cognitive
2356 P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371

enhancer to promote the learning process that occurs during CBT. ‘Normal’ brain ‘Depressed’ brain
Compounds under investigation include ampakines (Knapp et al., - + +
2002), agents that interact with the nicotinic cholinergic receptor vm-PFC dl-PFC onset
- dl-PFC
vm-PFC NAcc
(Mineur and Picciotto, 2010), and erythropoietin, a neuroprotective AMG NAcc
AMG
compound that, in addition to antidepressant effects in animal recovery
models and initial clinical trials also improves memory in animal
models of neurodegenerative conditions and in patients with cog-
nitive decline (Miskowiak et al., 2012). However, the possibility that
a cognitive enhancer could promote both adaptive and maladap-
tive change, injects an element of hazard into this strategy (Branchi,
2011). Hippocampal
damage
We finally consider a strategic issue, taking as an example the Involvement
inhibitory neurotransmitter GABA: whether the search for novel of stress in
antidepressants should begin from what is known about conven- precipitation Vulnerability
of depression (pre-existing
tional antidepressants or from a different starting point. There
or ‘scar’)
is increasing evidence that cortical GABA levels are decreased
in depression (Hasler et al., 2007; Croarkin et al., 2011; Hasler Treatment Brain
resistance repair
and Northoff, 2011), particularly so in treatment-resistant patients
(Price et al., 2009). Chronic antidepressant treatment decreases Antidepressant response
activity at GABA-B receptors (Cryan and Slattery, 2010; Ghose et al.,
2011). This suggests that GABA-B antagonists might have antide- Fig. 7. Some key conclusions. The upper part of the figure summarizes some of the
pressant properties, a hypothesis that has been amply confirmed structures involved in the shift between the ‘normal’ brain and the ‘depressed’ brain,
which is mediated by the balance between responsiveness to negative (amygdala:
in animal models but not yet tested clinically (Cryan and Slat-
AMG) and positive (nucleus accumbens: NAcc) information, and between rumi-
tery, 2010; Ghose et al., 2011). However, there is no good reason nation (ventromedial prefrontal cortex: vm-PFC) and cognitive control of emotion
to suppose that an approach derived from the known actions of (dorsolateral prefrontal cortex: dl-PFC). In depression, activity is increased in the
currently used antidepressants would improve on their clinical AMG and vm-PFC and decreased in the NAcc and dl-PFC, resulting in an inbalance
within this circuitry. This brief summary omits some key structures (e.g. habenula)
efficacy: on the contrary, the problems that remain uncorrected
and processes (e.g. appraisal) discussed in the text. The lower part of the figure
in treatment-resistant patients are more likely to be rectified by shows the relationship between the role of stress in the onset of depression and the
agents that have novel actions. Dysfunctions of GABA could be cor- effectiveness of monoaminergic antidepressants in treating depression. Antidepres-
rected by actions at either the GABA-B receptor, which is modulated sants are effective against depressions that are precipitated by high levels of stress,
by antidepressants, or the GABA-A receptor, which is not. It was because they promote the repair of a damaged hippocampus. However, in the pres-
ence of vulnerability factors, which could be constitutional and/or consequential to
recently reported that a GABA-A agonist, gaboxadol, not only has
previous episodes of depression, depression is precipitated by lower level of stress
antidepressant-like effects in the CMS model comparable to those of stress, causing less of an insult to the hippocampus, and thereby decreasing the
of escitalopram, which is among the most efficacious antidepres- response to antidepressant treatment.
sants in clinical use, but also potentiated the effect of escitalopram,
such that a higher reponse rate was achieved when the two drugs
were co-administered than with either drug alone. Gaboxadol did needed. Understanding the neurobiology of depression has helped
not increase hippocampal neurogenesis, and its mechanism of researchers uncover a number of novel targets for antidepressant
action has not yet been established (Christensen et al., 2012). How- therapies, which are under investigation in animal models, case
ever, GABA, acting via GABA-A receptors, promotes the survival reports, and small open-label studies. While many of these new
and maturation of newly born hippocampal neurons (Wang and approaches have suggested antidepressant potential, pivotal trials
Kriegstein, 2009), and it is plausible that this could be the mecha- will be needed to clarify which of these treatments may be realized
nism whereby gaboxadol acts synergistically with a conventional clinically. Over the next few years, the treatment of depression may
antidepressant. There is some evidence that GABA-enhancing anti- be improved by the introduction of several of these novel interven-
convulsants such as carbamazepine and sodium valproate have tions that no longer rely solely on monoamine reuptake inhibition.
antidepressant efficacy as monotherapy or adjunctive treatment However, the history of the past 50 years does not give cause for
(Vigo and Baldessarini, 2009), which perhaps should be further unrestrained optimism.
studied in light of the reported effects of gaboxadol. Perhaps some new conceptual approaches are needed. In this
From this brief review it is evident that the approaches review, starting from the established role of monoamines in antide-
to treatment-resistant depression are largely consistent with pressant action, we have attempted to expand the horizon to
attempts to intervene in the depression circuitry summarized integrate this early understanding into the wider neurocircuitry
above, but also include other approaches that were arrived at that is now known to be involved in the symptomatology of depres-
empirically and for which the mechanisms of action are not well sion, and the cellular and systemic mechanisms that underlie the
understood. Animal models of treatment resistance are now being ability of antidepressants to provide relief. We have focussed also
developed (Samuels et al., 2011), which may contribute to a better on the overlapping constellations of constitutional and experiential
understanding of existing strategies and improved prospects for factors that promote both vulnerability to depression and resis-
the treatment of those patients who currently are difficult to help. tance to treatment. Some of our key conclusions are summarized
in Fig. 7. The evidence suggests that depression results from an
inbalance between forebrain systems involved in processing and
5. Conclusions and research implications responding to affective information. This can arise either from the
action of high levels of stress, causing damage to the hippocam-
Despite the efficacy of currently available antidepressant med- pus, or from the combination of low levels of stress with a variety
ications and somatic therapies, residual depressive symptoms and of factors that confer vulnerability to depression, which may be
relapse are common. This creates a challenge for clinicians as they present from an early age, or may represent ‘scars’ left by previous
seek to eliminate symptoms completely and help fully recover depressive episodes. Monoaminergic antidepressants (i.e. almost
patients. To reach these goals, improved treatment strategies are of those currently available) act primarily by repairing a damaged
P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371 2357

hippocampus, and are therefore ineffective in depressions where antidepressant drug action, but in so doing, we have tended, in the
stress has had only a minor precipitating role. These conclusions are interests of clarity, to smooth over discrepancies and inconsisten-
based firmly on the position that depression is a stress-related con- cies in the literature. For example, we have given prominence to the
dition, while acknowledging that the role of stress may sometimes role of neurogenesis in antidepressant action, albeit that, according
be minor and decreases during the progression through a life his- to a recent commentary the neurogenesis hypothesis “has triggered
tory of depressive episodes. This distinction has important strategic reactions ranging from messianic conviction to total scepticism”
implications for both preclinical and clinical research. (Anacker and Pariante, 2012). As discussed above, many of the
From the perspective of a diathesis/stress model, animal mod- apparent inconsistencies in this area can be resolved: the neuro-
els of depression are mostly models of a first episode of depression. genetic effect of antidepressants is strain dependent, but so too is
They include models in which depressive-like behaviour is precip- their behavioural effect (Miller et al., 2008); the blockade of antide-
itated by external events, typically stressors of various kinds, and pressant action by suppression of neurogenesis appeared to be
models of a depressive diathesis involving individuals rendered technique-dependent (Bessa et al., 2009), but was later shown to be
vulnerable by virtue of their history (for example, a brain lesion) or actually time-dependent, involving relatively mature young neu-
genetic endowment (including both inbred strains and genetically rons (Mateus-Pinheiro et al., 2011); and while neurogenesis is not
modified animals) (Willner and Mitchell, 2002a,b). With very few necessary for all behavioural effects of antidepressants (David et al.,
exceptions (e.g. Isingrini et al., 2010), animal models of depression 2009), the forced swim test, where neurogenesis is not required,
have not addressed the progressive vulnerability that develops over is an acute test that is not considered a valid model of depres-
successive episodes of depression. Consequently, the preclinical lit- sion. Nevertheless, there are some exceptions that cannot readily
erature has almost nothing to say about the mechanisms by which be assimilated to the neurogenesis story: in particular, there are
depression becomes progressively more autonomous of stress. As antidepressant-like drugs that promote neurogenesis but whose
discussed above, this is a critical issue for both vulnerability to behavioural effects are not blocked by suppression of neurogene-
depression and resistance to antidepressant treatment. sis, including a melanin-concentrating hormone (MCH) antagonist
Conversely, clinical research rarely distinguishes between first (David et al., 2007), and CRF1 and vasopressin V1b antagonists
and recurrent depressive episodes. If, as envisaged here, the man- (Surget et al., 2008). But there is no reason to expect that all antide-
ner in which patients enter a depressive episode changes across pressants must hit the identical spot. There is evidence that the
episodes, then this could potentially account for much of the incon- site of action of the HPA (CRF1 and V1b) antagonists may be the
sistency in the neurobiological literature, in relation to, for example amygdala (Salome et al., 2006), and there is every possibility that
HPA activity or the changes reported in structural and functional the habenula may be the site of action of the MCH antagonist.
imaging studies. It is also a distinct possibility that if the neuro- The DBS literature demonstrates that antidepressant effects can
biological mechanisms change across episodes of depression, then be elicited from a variety of sites within the depression circuitry,
different antidepressant strategies might be appropriate at differ- including the habenula and the nucleus accumbens (Hamani et al.,
ent stages, and this might account for the clinical failure of so many 2010). The nucleus accumbens is likely to be the site of antidepres-
compounds predicted to show antidepressant activity from pre- sant action of DA agonists such as pramipexole (Willner, 1997b),
clinical studies. notwithstanding that they also stimulate neurogenesis (Albrecht
So this analysis suggests two research strategies. Firstly, it will and Buerger, 2009). So the intense attention focussed on hippocam-
be necessary to develop animal models of repeated depressive pal neurogenesis may be a case of looking in the light. There are
episodes. As the object is to approximate the clinical situation, very good reasons to shine an intense light on the hippocampus: it
these models should aim to use realistic conditions based, for does seem to be a critical region for the neuropathological effects
example, on the CMS model rather than, for example, the tail of stress, and repair within this region is important, if not uni-
suspension or the forced swim tests, which have only the most versally critical, for recovery. But other parts of the depression
tenuous relationship to depression (Cryan and Holmes, 2005). And circuitry that have been less brightly illuminated also merit atten-
as the rationale is to investigate changes in the role of stress, it tion.
would be better in general to use environmental precipitants (e.g. A broader focus is essential because, according to our cur-
CMS) rather than a labour-saving short cut such as corticosterone rent understanding, the hippocampus is almost incidental to
administration (David et al., 2009; Samuels and Hen, 2011), which the symptomatology of depression. Neurogenesis simply main-
could be a useful way of identifying systems of interest, but pre- tains an efficient level of functioning within the hippocampus
empts investigation of important elements of the stress system. (Airan et al., 2007), which is concerned primarily with contex-
The research envisaged in this proposal would provide a more tual learning (Rudy, 2009). The importance in depression of the
solid backdrop against which to test the mechanisms of depres- hippocampal formation, including the subiculum and entorhi-
sive diatheses, such as genetic polymorphisms, than is currently nal cortex, derives from its connections with other regions that
available, and could engender novel leads for antidepressant drug have more direct involvement in the psychological (amygdala,
development. PFC and other areas of association cortex, nucleus accumbens)
Secondly, clinical research should rigorously attempt to exam- and physiological (hypothalamus) phenomena that characterize
ine separately first and recurrent episodes of depression (Meyer, depression. The most critical regions appear to be the ante-
2012). Indeed, the population of greatest research interest at rior cingulate and ventral prefrontal cortex, but, as discussed,
present are first-episode depressed patients with a weak depres- all parts of the system can be related to specific symptoms of
sive diathesis, in whom the effects of stress are greatest. It should depression: in particular, rumination and negative self-referential
be possible in this homogeneous group to iron out the inconsis- attributions reflect hyperactivity in amygdala and ventral pre-
tencies in the literature and achieve a much clearer picture of frontal regions, and anhedonia reflects hypoactivity in the nucleus
the neurobiological mechanisms. In tandem with the preclinical accumbens and dorsal PFC. It is obvious that such a complex
research, this would provide a sounder basis for understanding network will generate a wide variety of symptom patterns, depend-
the growth of the depressive diathesis over episodes of depres- ing on the degree of dysfunction elicited in each component,
sion, and interactions with other factors that confer vulnerability which will in turn depend upon the individual patient’s pro-
to depression. file of vulnerability factors and precipitants, and the manner in
We have attempted in this review to provide a clear account of which they affect different brain regions. It is to be expected that
the current state of research in the neurobiology of depression and in the future novel compounds might target specific syndromes
2358 P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371

such as rumination or anhedonia, or specific physiological vul- (hippocampus, PFC, habenula . . .) or neurotransmitters (5HT, NA,
nerabilities; but the prospect of personalized antidepressant DA, cannabinoid . . .) in order to engage effectively with the com-
medication (Holsboer, 2008) is unlikely to be realized any time plexity of the neurobiological systems that are now known to be
soon. implicated in depression.
It is also obvious that traditional ways of thinking about depres-
sion, based around a single brain region or a single neurotransmitter
References
system, are no longer adequate for dealing with the complex sys-
tem that is now known to underly depression. A potentially more Abraham, G., Milev, R., Stuart Lawson, J., 2006. T3 augmentation of SSRI resistant
appropriate conceptualization was proposed by Tanti and Belzung depression. Journal of Affective Disorders 91, 211–215.
Adachi, M., Barrot, M., Autry, A.E., Theobald, D., Monteggia, L.M., 2008. Selective loss
(2010). They describe a theoretical neural network model that has of brain-derived neurotrophic factor in the dentate gyrus attenuates antidepres-
a number of input nodes representing vulnerability and triggering sant efficacy. Biological Psychiatry 63, 642–649.
factors, a set of inter-connected intermediate nodes representing Ahearn, E.P., Jamison, K.R., Steffens, D.C., Cassidy, F., Provenzale, J.M., Lehman, A.,
Weisler, R.H., Carroll, B.J., Krishnan, K.R., 2001. MRI correlates of suicide attempt
physiological endpoints (brain regions or neurotransmitter sys- history in unipolar depression. Biological Psychiatry 50, 266–270.
tems) that are influenced to varying extents by each of the inputs, Airan, R.D., Meltzer, L.A., Roy, M., Gong, Y., Chen, H., Deisseroth, K., 2007. High-
and a range of output nodes representing cognitive and behavioural speed imaging reveals neurophysiological links to behavior in an animal model
of depression. Science 317, 819–823.
functions that are underpinned to varying degrees by the physio-
Aisa, B., Elizalde, N., Tordera, R., Lasheras, B., Del Río, J., Ramírez, M.J., 2009. Effects
logical elements of the model. The authors show, inter alia, how of neonatal stress on markers of synaptic plasticity in the hippocampus: impli-
such a model could represent different routes to the same outcome cations for spatial memory. Hippocampus 19, 1222–1231.
Aizenstein, H.J., Butters, M.A., Wu, M., Mazurkewicz, L.M., Stenger, V.A., Gianaros,
(e.g. via dysfunction either of a critical physiological function, or of
P.J., et al., 2009. Altered functioning of the executive control circuit in late-life
several functions none of which is critical in itself); how antidepres- depression: episodic and persistent phenomena. American Journal of Geriatric
sants could promote recovery of key symptoms with or without Psychiatry 17, 30–42.
a reversal of the original pathophysiological process; and how a Akiskal, H.S., 1984. A integrative view on the etiology and treatment of depres-
sion. In: Korf, J., Pepplinkuizen (Eds.), Depression: Molecular and Psychologically
change in cognition following CBT could back-propagate through Based Therapies. TGO Foundation, Drachten, pp. 980–1111.
the system to affect the functioning of the physiological compo- Albrecht, S., Buerger, E., 2009. Potential neuroprotection mechanisms in PD: focus
nents. The explanatory potential of the model is considerable, but on dopamine agonist pramipexole. Current Medical Reserach Opinion 25,
2977–2987.
this will only be realized in the future, as the model has not yet Alexopoulos, G.S., Meyers, B.S., Young, R.C., Kakuma, T., Silbersweig, D., Charlson,
been built. M., 1997. Clinically defined vascular depression. American Journal of Psychiatry
There is also a need for formal mathematical models of 154, 562–565.
Alexopoulos, G.S., Bruce, M.L., Silbersweig, D., Kalayam, B., Stern, E., 1999. Vascular
physiological components of the depression circuitry, where the depression: a new view of late-onset depression. Dialogues in Clinical Neuro-
inter-relationships are so many and complex as to defy accurate science 1, 68–80.
intuitive prediction of the effects of any perturbation. Belzung and Alexopoulos, G.S., Kiosses, D.N., Murphy, C., Heo, M., 2004. Executive dysfunction,
heart disease burden, and remission of geriatric depression. Neuropsychophar-
Billette de Villemeur (2010) have presented such a formal compu-
macology 29, 2278–2284.
tational model of the hippocampal control of the HPA axis, which Alonso, R., Griebel, G., Pavone, G., Stemmelin, J., Le, F.G., Soubrie, P., 2004. Blockade
integrates hippocampal and HPA dysfunction in depression, and of CRF(1) or V(1b) receptors reverses stress-induced suppression of neuro-
genesis in a mouse model of depression. Molecular Psychiatry 9, 278–286,
their normalization by antidepressant drugs, and enables the spec-
224.
ification of precise hypotheses that can be tested empirically. There Alloy, L.B., Abramson, L.Y., Whitehouse, W.G., Hogan, M.E., Tashman, N.A., Steinberg,
is evidently scope to develop computational models of other facets D.L., Rose, D.T., Donovan, P., 1999. Depressogenic cognitive styles: predictive
of the depression circuitry, such as the 5HT projections to forebrain validity, information processing and personality characteristics, and develop-
mental origins. Behaviour Research and Therapy 37, 503–531.
structures and the reciprocal projections from the forebrain to the Amat, J., Sparks, P.D., Matus-Amat, P., Griggs, J., Watkins, L.R., Maier, S.F., 2001. The
DRN, or the relationships between the amygdala, PFC and nucleus role of the habenular complex in the elevation of dorsal raphe nucleus serotonin
accumbens. and the changes in the behavioral responses produced by uncontrollable stress.
Brain Research 917, 118–126.
Finally, a concept that demands attention is the suggestion that Amat, J., Baratta, M.V., Paul, E., Bland, S.T., Watkins, L.R., Maier, S.F., 2005. Medial
what is important about depression is not so much its sympto- prefrontal cortex determines how stressor controllability affects behavior and
matology as its chronicity: that is, that the phenomenology of dorsal raphe nucleus. Nature Neuroscience 8, 365–371.
Amat, J., Paul, E., Zarza, C., Watkins, L.R., Maier, S.F., 2006. Prior experience with
depression is frequently a natural response to adversity, and what behavioral control over stress blocks the behavioral effects of later uncontrol-
distinguishes people who become depressed from those who do lable stress: role of the ventral medial prefrontal cortex. Journal of Neuroscience
not is that they become “stuck in a rut” and unable to recover 26, 13264–13272.
Amat, J., Paul, E., Watkins, L.R., Maier, S.F., 2008. Activation of the ventral medial
(Holtzheimer and Mayberg, 2011). In this conceptualization, the
prefrontal cortex during an uncontrollable stressor reproduces both the imme-
onset of and recovery from depression are seen as transitions into diate and long-term protective effects of behavioral control. Neuroscience 154,
and out of a stable state. As discussed above, the cognitive appraisal 1178–1186.
Amat, J., Aleksejev, R.M., Paul, E., Watkins, L.R., Maier, S.F., 2010. Behavioral control
that stress is uncontrollable appears to play a critical role in the
over shock blocks behavioral and neurochemical effects of later social defeat.
transition into depression, and a drug- or stimulation-induced neu- Neuroscience 165, 1031–1038.
rogenetic boost to the efficiency of information processing in the Ambudkar, S.V., Dey, S., Hrycyna, C.A., Ramachandra, M., Pastan, I., Gottesman, M.M.,
hippocampus appears to provide one exit route. While these phase 1999. Biochemical, cellular, and pharmacological aspects of the multidrug trans-
porter. Annual Review of Pharmacology and Toxicology 39, 361–398.
shifts may be captured by conventional connectionist or computa- American Psychiatric Association, 1994. Diagnostic and Statistical Manual of Mental
tional models such as those described above, they may also require Disorders, 4th ed. APA, Washington.
the inclusion of non-linear computational methods such as those Amrein, I., Isler, K., Lipp, H.P., 2011. Comparing adult hippocampal neurogenesis in
mammalian species and orders: influence of chronological age and life history
provided by chaos theory. stage. The European Journal of Neuroscience 34, 978–987.
So to summarize, our three key recommendations for future Anacker, C., Pariante, C.M., 2012. Stress and neurogenesis: can adult neurogene-
depression research are: greater attention to the diathesis-stress sis buffer stress responses and depressive behaviour? Molecular Psychiatry 17,
9–10.
model in clinical studies, and in particular, a rigorous separation of Anacker, C., Zunszain, P.A., Carvalho, L.A., Pariante, C.M., 2011a. The glucocorticoid
first-episode and recurrent depression; greater attention in preclin- receptor: pivot of depression and of antidepressant treatment? Psychoneuroen-
ical research to issues of vulnerability, including the development docrinology 36, 415–425.
Anacker, C., Zunszain, P.A., Cattaneo, A., Carvalho, L.A., Garabedian, M.J., Thuret, S.,
of models of recurrent depression; and the involvement of math-
Price, J., Pariante, C.M., 2011b. Antidepressants increase human hippocampal
ematicians in research teams, with the aim of improving their neurogenesis by activating the glucocorticoid receptor. Molecular Psychiatry
capacity to break out of the traditional focus on single structures 16, 738–750.
P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371 2359

Appelhopf, B.C., Huyser, J., Verweij, M., Brouwer, J.P., van Dyck, R., Fliers, Berton, O., McClung, C.A., Dileone, R.J., Krishnan, V., Renthal, W., Russo, S.J., Graham,
E., et al., 2006. Glucocorticoids and relapse of major depression D., Tsankova, N.M., Bolanos, C.A., Rios, M., Monteggia, L.M., Self, D.W., Nestler,
(dexamethasone/corticotropin-releasing hormone test in relation to relapse of E.J., 2006a. Essential role of BDNF in the mesolimbic dopamine pathway in social
major depression). Biological Psychiatry 59, 696–701. defeat stress. Science 311, 864–868.
Arango, V., Underwood, M.D., Mann, J.J., 1997. Postmortem findings in suicide vic- Berton, O., Covington 3rd, H.E., Ebner, K., Tsankova, N.M., Carle, T.L., Ulery, P., Bhonsle,
tims. Implications for in vivo imaging studies. Annals of the New York Academy A., Barrot, M., Krishnan, V., Singewald, G.M., Singewald, N., Birnbaum, S., Neve,
of Sciences 836, 269–287. R.L., Nestler, E.J., 2007. Induction of %FosB in the periaqueductal gray by stress
Arborelius, L., Owens, M.J., Plotsky, P.M., Nemeroff, C.B., 1999. The role of promotes active coping responses. Neuron 55, 289–300.
corticotropin-releasing factor in depression and anxiety disorders. The Journal Berman, R.M., Cappiello, A., Anand, A., Oren, D.A., Heninger, G.R., Charney, D.S.,
of Endocrinology 160, 1–12. Krystal, J.H., 2000. Antidepressant effects of ketamine in depressed patients.
Arias, B., Serretti, A., Mandelli, L., Gastó, C., Catalán, R., Ronchi, D.D., Faañanás, L., Biological Psychiatry 47, 351–354.
2009. Dysbindin gene (DTNBP1) in major depression: association with clini- Berton, O., McClung, C.A., Dileone, R.J., Krishnan, V., Renthal, W., Russo, S.J., Graham,
cal response to selective serotonin reuptake inhibitors. Pharmacogenetics and D., Tsankova, N.M., Bolanos, C.A., Rios, M., Monteggia, L.M., Self, D.W., Nestler,
Genomics 19, 121–128. E.J., 2006b. Essential role of BDNF in the mesolimbic dopamine pathway in social
Avagianou, P.A., Zafiropoulou, M., 2008. Parental bonding and depression: person- defeat stress. Science 311, 868–878.
ality as a mediating factor. International Journal of Adolescent Medicine and Bessa, J.M., Ferreira, D., Melo, I., Marques, F., Cerqueira, J.J., Palha, J.A., Almeida, O.F.,
Health 20, 261–269. Sousa, N., 2009. The mood-improving actions of antidepressants do not depend
Avitsur, R., Stark, J.L., Sheridan, J.F., 2001. Social stress induces glucocorticoid resis- on neurogenesis but are associated with neuronal remodeling. Molecular Psy-
tance in subordinate animals. Hormones and Behavior 39, 247–257. chiatry 14, 764–773.
Avitsur, R., Powell, N., Padgett, D.A., Sheridan, J.F., 2009. Social interactions, stress, Bewernick, B.H., Hurlemann, R., Matusch, A., Kayser, S., Grubert, C., Hadrysiewicz,
and immunity. Immunology and Allergy Clinics of North America 29, 285–293. B., Axmacher, N., Lemke, M., Cooper-Mahkorn, D., Cohen, M.X., Brockmann, H.,
Aznar, S., Klein, A.B., Santini, N.A., Knudsen, G.M., Henn, F., Gass, P., Vollmayr, B., 2010. Lenartz, D., Sturm, V., Schlaepfer, T.E., 2010. Nucleus accumbens deep brain
Aging and depression vulnerability interaction results in decreased serotonin stimulation decreases ratings of depression and anxiety in treatment-resistant
innervation associated with reduced BDNF levels in hippocampus of rats bred depression. Biological Psychiatry 67, 110–116.
for learned helplessness. Synapse 64, 561–565. Bhagwagar, Z., Hinz, R., Taylor, M., Fancy, S., Cowen, P., Grasby, P., 2006. Increased
Baghai, T.C., Blier, P., Baldwin, D.S., Bauer, M., Goodwin, G.M., Fountoulakis, K.N., 5-HT2A receptor binding in euthymic, medication-free patients recovered from
Kasper, S., Leonard, B.E., Malt, U.F., Stein, D., Versiani, M., Möller, H.J., Section depression: a positron emission study with [11C]MDL 100,907. The American
of Pharmacopsychiatry, World Psychiatric Association, 2011. General and com- Journal of Psychiatry 163, 1580–1587.
parative efficacy and effectiveness of antidepressants in the acute treatment Bhagwagar, Z., Wylezinska, M., Jezzard, P., Evans, J., Boorman, E., Matthews, M.P.,
of depressive disorders: a report by the WPA section of pharmacopsychia- Cowen, J.P., 2008. Low GABA concentrations in occipital cortex and anterior
try. European Archives of Psychiatry and Clinical Neuroscience 261 (Suppl. 3), cingulate cortex in medication-free, recovered depressed patients. The Inter-
207–245. national Journal of Neuropsychopharmacology 11, 255–260.
Baldwin, R., Jeffries, S., Jackson, A., Sutcliffe, C., Thacker, N., Scott, M., Burns, A., Bhalla, R.K., Butters, M.A., Mulsant, B.H., Begley, A.E., Zmuda, M.D., Schoderbek, B.,
2004. Treatment response in late-onset depression: relationship to neuropsy- Pollock, B.G., Reynolds 3rd., C.F., Becker, J.T., 2006. Persistence of neuropsycho-
chological, neuroradiological and vascular risk factors. Psychological Medicine logic deficits in the remitted state of late-life depression. The American Journal
34, 125–136. of Geriatric Psychiatry: Official Journal of the American Association for Geriatric
Ballard, I.C., Murty, V.P., Carter, R.M., MacInnes, J.J., Huettel, S.A., Adcock, R.A., 2011. Psychiatry 14, 419–427.
Dorsolateral prefrontal cortex drives mesolimbic dopaminergic regions to initi- Bisgaard, C.F., Jayatissa, M.N., Enghild, J.J., Sanchéz, C., Artemychyn, R., Wiborg,
ate motivated behavior. The Journal of Neuroscience: the Official Journal of the O., 2007. Proteomic investigation of the ventral rat hippocampus links DRP-
Society for Neuroscience 31, 10340–10346. 2 to escitalopram treatment resistance and SNAP to stress resilience in the
Balleine, B.W., O’Doherty, J.P., 2010. Human and rodent homologies in action control: chronic mild stress model of depression. Journal of Molecular Neuroscience 32,
corticostriatal determinants of goal-directed and habitual action. Neuropsycho- 132–144.
pharmacology 35, 48–69. Blendy, J.A., 2006. The role of CREB in depression and antidepressant treatment.
Ballmaier, M., Narr, K.L., Toga, A.W., Elderkin-Thompson, V., Thompson, P.M., Hamil- Biological Psychiatry 59, 1144–1150.
ton, L., Haroon, E., Pham, D., Heinz, A., Kumar, A., 2008. Hippocampal morphology Blier, P., 2010. The well of novel antidepressants: running dry. Journal of Psychiatry
and distinguishing late-onset from early-onset elderly depression. The American & Neuroscience 35, 219–220.
Journal of Psychiatry 165, 229–237. Blier, P., Blondeau, C., 2011. Neurobiological bases and clinical aspects of the use of
Bambico, F.R., Katz, N., Debonnel, G., Gobbi, G., 2007. Cannabinoids elicit aripiprazole in treatment-resistant major depressive disorder. Journal of Affec-
antidepressant-like behavior and activate serotonergic neurons through the tive Disorders 128 (Suppl. 1), S3–S10.
medial prefrontal cortex. The Journal of Neuroscience: the Official Journal of Bluhm, R., Williamson, P., Lanius, R., Théberge, J., Densmore, M., Bartha, R., Neufeld,
the Society for Neuroscience 27, 11700–11711. R., Osuch, E., 2009. Resting state default-mode network connectivity in early
Banasr, M., Hery, M., Printemps, R., Daszuta, A., 2004. Serotonin-induced increases depression using a seed region-of-interest analysis: decreased connectivity with
in adult cell proliferation and neurogenesis are mediated through different and caudate nucleus. Psychiatry and Clinical Neurosciences 63, 754–761.
common 5-HT receptor subtypes in the dentate gyrus and the subventricular Bodnoff, S.R., Humphreys, A.G., Lehman, J.C., Diamond, D.M., Rose, G.M., Meaney, M.J.,
zone. Neuropsychopharmacology 29, 450–460. 1995. Enduring effects of chronic corticosterone treatment on spatial learning,
Baratta, M.V., Zarza, C.M., Gomez, D.M., Campeau, S., Watkins, L.R., Maier, S.F., 2009. synaptic plasticity, and hippocampal neuropathology in young and mid-aged
Selective activation of dorsal raphe nucleus-projecting neurons in the ventral rats. The Journal of Neuroscience: the Official Journal of the Society for Neuro-
medial prefrontal cortex by controllable stress. The European Journal of Neuro- science 15, 61–69.
science 30, 1111–1116. Boes, A.D., McCormick, L.M., Coryell, W.H., Nopoulos, P., 2008. Rostral anterior cingu-
Beck, A.T., 1967. Depression: Causes and Treatment. University of Pennsylvania late cortex volume correlates with depressed mood in normal healthy children.
Press, Philadelphia. Biological Psychiatry 63, 391–397.
Beck, A.T., 1983. Cognitive therapy of depression: new perspectives. In: Clayton, Bolwig, T.G., 2011. How does electroconvulsive therapy work? Theories on its mech-
P.J., Barrett, J.S. (Eds.), Treatment of Depression: Old Controversies and New anism. Canadian Journal of Psychiatry 56, 13–18.
Approaches. Raven, New York, pp. 265–290. Boldrini, M., Underwood, M., Hen, R., Rosoklija, G., Dwork, A., John Mann, J., Arango,
Bedi, R.P., 1999. Depression: an inability to adapt to one’s perceived life distress? V., 2009. Antidepressants increase neural progenitor cells in the human hip-
Journal of Affective Disorders 54, 225–234. pocampus. Neuropsychopharmacology 34, 2376–2389.
Beer, J.S., Hughes, B.L., 2010. Neural systems of social comparison and the “above- Bondi, C.O., Jett, J.D., Morilak, D.A., 2010. Beneficial effects of desipramine on cog-
average” effect. NeuroImage 49, 2671–2679. nitive function of chronically stressed rats are mediated by alpha1-adrenergic
Beer, J.S., John, O.P., Scabini, D., Knight, R.T., 2006. Orbitofrontal cortex and social receptors in medial prefrontal cortex. Progress in Neuro-psychopharmacology
behavior: integrating self-monitoring and emotion-cognition interactions. Jour- & Biological Psychiatry 34, 913–923.
nal of Cognitive Neuroscience 18, 871–879. Bondolfi, G., Chautems, C., Rochat, B., Bertschy, G., Baumann, P., 1996. Non-response
Behrens, M., Ali, S.S., Dao, D.N., Lucero, J., Shekhtman, G., Quick, K.L., Dugan, L.L., 2007. to citalopram in depressive patients: pharmacokinetic and clinical conse-
Ketamine-induced loss of phenotype of fast-spiking interneurons is mediated quences of a fluvoxamine augmentation. Psychopharmacology 128, 421–425.
by NADPH-oxidase. Science 318, 1645–1647. Bora, E., Harrison, B.J., Davey, C.G., Yucel, M., Pantelis, C., 2011. Meta-analysis of
Belsky, J., Jonassaint, C., Pluess, M., Stanton, M., Brummett, B., Williams, R., 2009. volumetric abnormalities in cortico-striatal-pallidal-thalamic circuits in major
Vulnerability genes or plasticity genes? Molecular Psychiatry 14, 746–754. depressive disorder. Psychological Medicine 13, 1–11.
Belzung, C., Billette de Villemeur, E., 2010. The design of new antidepressants: can Branchi, I., 2011. The double edged sword of neural plasticity: increasing serotonin
formal models help? A first attempt using a model of the hippocampal control levels leads to both greater vulnerability to depression and improved capacity
over the HPA-axis based on a review from the literature. Behavioural Pharma- to recover. Psychoneuroendocrinology 36, 339–351.
cology 21, 677–689. Brezun, J.M., Daszuta, A., 1999. Depletion in serotonin decreases neurogenesis in
Bergami, M., Rimondini, R., Santi, S., Blum, R., Gotz, M., Canossa, M., 2008. the dentate gyrus and the subventricular zone of adult rats. Neuroscience 89,
Deletion of TrkB in adult progenitors alters newborn neuron integration 999–1002.
into hippocampal circuits and increases anxiety-like behavior. Proceedings Brezun, J.M., Daszuta, A., 2000. Serotonin may stimulate granule cell proliferation in
of the National Academy of Sciences of the United States of America 105, the adult hippocampus, as observed in rats grafted with foetal raphe neurons.
15570–15575. European Journal of Neuroscience 12, 391–396.
2360 P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371

Brøsen, K., 2004. Some aspects of genetic polymorphism in the biotransformation BDNF (Val66Met) polymorphism alters anxiety-related behavior. Science 314,
of antidepressants. Therpie 59, 5–12. 140–143.
Brouwer, J.P., Appelhof, B.C., van Rossum, E.F., Koper, J.W., Fliers, E., Huyser, J., Chourbaji, S., Vogt, M.A., Fumagalli, F., Sohr, R., Frasca, A., Brandwein, C., Hörtnagl,
et al., 2006. Prediction of treatment response by HPA-axis and glucocorticoid H., Riva, M.A., Sprengel, R., Gass, P., 2008. AMPA receptor subunit 1 (GluR-A)
receptor polymorphisms in major depression. Psychoneuroendocrinology 31, knockout mice model the glutamate hypothesis of depression. The FASEB Jour-
1154–1163. nal: Official Publication of the Federation of American Societies for Experimental
Brown, G.W., Harris, T., 1978. Social Origins of Depression. Tavistock, London. Biology 22, 3129–3134.
Browning, M., Grol, M., Ly, V., Goodwin, G.M., Holmes, E.A., Harmer, C.J., 2011. Christensen, M.V., Kessing, L.V., 2005. Clinical use of coping in affective disorder,
Using an experimental medicine model to explore combination effects of a critical review of the literature. Clinical Practice and Epidemiology in Mental
pharmacological and cognitive interventions for depression and anxiety. Neu- Health 1, 20.
ropsychopharmacology 36, 2689–2697. Christensen, M.V., Kessing, L.V., 2006. Do personality traits predict first onset
Browning, M., Reid, C., Cowen, P.J., Goodwin, G.M., Harmer, C., 2007. A single dose of in depressive and bipolar disorder? Nordic Journal of Psychiatry 60,
citalopram increases fear recognition in healthy subjects. Journal of Psychophar- 79–88.
macology 21, 684–690. Christensen, T., Bisgaard, C.F., Wiborg, O., 2011. Biomarkers of anehdonic-like
Broyd, S.J., Demanuele, C., Debener, S., Helps, S.K., James, C.J., Sonuga-Barke, E.J.S., behaviour, antidepressant drug refraction, and stress resilience in a rat model
2009. Default-mode brain dysfunction in mental disorders: a systematic review. of depression. Neuroscience 196, 66–79.
Neuroscience and Biobehavioral Reviews 33, 279–296. Christensen, T., Betry, C., Mnie-Filai, O., Etievant, A., Ebert, B., Haddjeri, N., Wiborg,
Bukh, J.D., Bock, C., Vinberg, M., Gether, U., Kessing, L.V., 2011. Differences between O., 2012. Synergistic antidepressant-like action of gaboxadol and escitalopram.
early and late onset adult depression. Clinical Practice and Epidemiology in European Neuropsychopharmacology 22, 751–760.
Mental Health 7, 140–147. Christianson, J.P., Thompson, B.M., Watkins, L.R., Maier, S.F., 2009. Medial prefrontal
Bush, T.G., Savidge, T.C., Freeman, T.C., Cox, H.J., Campbell, E.A., Mucke, L., John- cortical activation modulates the impact of controllable and uncontrollable
son, M.Y., Sofroniew, M.V., 1998. Fulminant jejuno-ileitis following ablation of stressor exposure on a social exploration test of anxiety in the rat. Stress 12,
enteric glia in adult transgenic mice. Cell 93, 189–201. 445–450.
Butters, M.A., Aizenstein, H.J., Hayashi, K.M., Meltzer, C.C., Seaman, J., Reynolds, C.F., Cipriani, A., Furukawa, T.A., Salanti, G., Geddes, J.R., Higgins, J.P.T., Churchill, R.,
Toga, A.W., Thompson, P.M., Becker, J.T., 2009. Three-dimensional surface map- Watanabe, N., Nakagawa, A., Omori, I.M., McGuire, H., Tansella, M., Barbui, C.,
ping of the caudate nucleus in ate-life depression. The American Journal of 2009. Comparative efficacy and acceptability of 12 new-generation antidepres-
Geriatric Psychiatry: Official Journal of the American Association for Geriatric sants: a multiple-treatments meta-analysis. Lancet 373, 746–758.
Psychiatry 17, 4–12. Clark, L., Chamberlain, S.R., Sahakian, B.J., 2009. Neurocognitive mechanisms in
Bystrisky, A., 2006. Treatment-resistant anxiety disorders. Molecular Psychiatry 11, depression: implications for treatment. Annual Review of Neuroscience 32,
805–814. 57–74.
Cador, M., 1992. Limbic-striatal interactions in reward-related processes: modula- Cole, J., Toga, A.W., Hojatkashani, C., Thompson, P., Costafreda, S.G., Cleare, A.J.,
tion by the dopaminergic system. Clinical Neuropharmacology 15 (Suppl. 1), Pt Williams, S.C., Bullmore, E.T., Scott, J.L., Mitterschiffthaler, M.T., Walsh, N.D.,
A:548A–Pt A:549A. Donaldson, C., Mirza, M., Marquand, A., Nosarti, C., McGuffin, P., Fu, C.H., 2010.
Caldecott-Hazard, S., Mazziotta, J., Phelps, M., 1988. Cerebral correlates of depressed Subregional hippocampal deformations in major depressive disorder. Journal of
behavior in rats, visualized using 14C-2-deoxyglucose autoradiography. The Affective Disorders 126, 272–277.
Journal of Neuroscience: the Official Journal of the Society for Neuroscience 8, Colla, M., Kronenberg, G., Deuschle, M., Meichel, K., Hagen, T., Bohrer, M., Heuser, I.,
1951–1961. 2007. Hippocampal volume reduction and HPA-system activity in major depres-
Callen, D.F., Baker, E., Simmers, R.N., Seshadri, R., Roninson, I.B., 1987. Localization sion. Journal of Psychiatric Research 41, 553–560.
of the human multiple drug resistance gene, MDR1, to 7q21. 1. Human Genetics Compass, B.E., Connor-Smith, J., Jaser, S.S., 2004. Temperament, stress reactivity,
77, 142–144. and coping: implications for depression in childhood and adolescence. Journal
Campbell, S., Marriott, M., Nahmias, C., MacQueen, G.M., 2004. Lower hippocampal of Clinical Child and Adolescent Psychology: the Official Journal for the Society of
volume in patients suffering from depression: a meta-analysis. The American Clinical Child and Adolescent Psychology, American Psychological Association,
Journal of Psychiatry 161, 598–607. Division 53 33, 21–31.
Camus, V., Kraehenbühl, H., Preisig, M., Büla, C.J., Waeber, G., 2004. Geriatric depres- Connolly, K.R., Thase, M.E., 2011. If at first you don’t succeed: a review of the evidence
sion and vascular diseases: what are the links? Journal of Affective Disorders 81, for antidepressant augmentation, combination and switching strategies. Drugs
1–16. 71, 43–64.
Canto, C.B., Wouterlood, F.G., Witter, M.P., 2008. What does the anatomical organi- Cooney, R.E., Joormann, J., Eugene, F., Dennis, E.L., Gotlib, I.H., 2010. Neural correlates
zation of the entorhinal cortex tell us? Neural Plasticity 2008, Article ID 381243. of rumination in depression. Cognitive, Affective & Behavioral Neuroscience 10,
Cao, L., Jiao, X., Zuzga, D.S., Liu, Y., Fong, D.M., Young, D., During, M.J., 2004. VEGF links 470–478.
hippocampal activity with neurogenesis, learning and memory. Nature Genetics Cooper, D.C., Klipec, W.D., Fowler, M.A., Ozkan, E.D., 2006. A role for the suibicu-
36, 827–835. lum in the brain motivation/reward circuitry. Behavioural Brain Research 174,
Carlson, P.J., Singh, J.B., Zarate Jr., C.A., Drevets, W.C., Manji, H.K., 2006. Neural cir- 225–231.
cuitry and neuroplasticity in mood disorders: insights for novel therapeutic Coppell, A.L., Pei, Q., Zetterstrom, T.S.C., 2003. Bi-phasic change in BDNF gene
targets. NeuroRx. 3, 22–41. expression following antidepressant drug treatment. Neuropharmacology 44,
Carney, R.M., Shelton, R.C., 2011. Agomelatine for the treatment of major depressive 903–910.
disorder. Expert Opinion on Pharmacotherapy 12, 2411–2419. Corballis, M.C., 2009. The evolution and genetics of cerebral asymmetry. Philosophi-
Carvalho, L.A., Pariante, C.M., 2008. In vitro modulation of the glucocorticoid receptor cal Transactions of the Royal Society of London. Series B, Biological Sciences 364,
by antidepressants. Stress 11, 411–424. 867–879.
Caspi, A., Sugden, K., Moffitt, T.E., Taylor, A., Craig, I.W., Harrington, H., McClay, J., Cordon-Cardo, C., O’Brien, J.P., Casals, D., Rittman-Grauer, L., Biedler, J.L., Melamed,
Mill, J., Martin, J., Braithwaite, A., Poulton, R., 2003. Influence of life stress on M.R., Bertino, J.R., 1989. Multi-resistance gene (P-glycoprotein) is expressed
depression: moderation by a polymorphism in the 5-HTT gene. Science 301, by endothelial cells at blood-brain barrier sites. Proceedings of the National
386–389. Academy of Science USA 86, 695–698.
Caspi, A., Hariri, A.R., Holmes, A., Uher, R., Moffitt, T.E., 2010. Genetic sensitivity to the Corrigan, M.H., Denahan, A.Q., Wright, C.E., Ragual, R.J., Evans, D.L., 2000. Compari-
environment: the case of the serotonin transporter gene and its implications for son of pramipexole, fluoxetine, and placebo in patients with major depression.
studying complex diseases and traits. The American Journal of Psychiatry 167, Depression and Anxiety 11, 58–65.
509–527. Corruble, E., Falissard, B., Gorwood, P., 2011. Is DSM-IV bereavement exclusion for
Castrén, E., Rantamäki, T., 2010. The role of BDNF and its receptors in depression and major depression relevant to treatment response? A case-control, prospective
antidepressant drug action: reactivation of developmental plasticity. Develop- study. The Journal of Clinical Psychiatry 72, 898–902.
mental Neurobiology 70, 289–297. Coryell, M.W., Wunsch, A.M., Haenfler, J.M., Allen, J.E., Schnizler, M., Ziemann,
Castrén, E., Võikar, V., Rantamäki, T., 2007. Role of neurotrophic factors in depression. A.E., Cook, M.N., Dunning, J.P., Price, M.P., Rainier, J.D., Liu, Z., Light, A.R.,
Current Opinion in Pharmacology 7, 18–21. Langbehn, D.R., Wemmie, J.A., 2009. Acid-sensing ion channel-1a in the amyg-
Chan, S.W., Goodwin, G.M., Harmer, C.J., 2007. Highly neurotic never-depressed stu- dala, a novel therapeutic target in depression-related behavior. The Journal
dents have negative biases in information processing. Psychological Medicine of Neuroscience: the Official Journal of the Society for Neuroscience 29,
37, 1281–1291. 5381–5388.
Chartoff, E.H., Papadopoulou, M., MacDonald, M.L., Parsegian, A., Potter, D., Konradi, Couillard-Despres, S., Wuertinger, C., Kandasamy, M., Caioni, M., Stadler, K., Aigner,
C., Carlezon Jr., W.A., 2009. Desipramine reduces stress-activated dynorphin R., Bogdahn, U., Aigner, L., 2009. Ageing abolishes the effects of fluoxetine on
expression and CREB phosphorylation in NAc tissue. Molecular Pharmacology neurogenesis. Molecular Psychiatry 14, 856–864.
75, 704–712. Covington, H.E., Maze, I., LaPlant, Q.C., Vialou, V.F., Ohnishi, Y.N., Berton, O., Fass, D.M.,
Chase, H.W., Frank, M.J., Michael, A., Bullmore, E.T., Sahakian, B.J., Robbins, T.W., Renthal, W., Rush, A.J., Wu, E.Y., Ghose, S., Krishnan, V., Russo, S.J., Tamminga, C.,
2010. Approach and avoidance learning in patients with major depression and Haggarty, S.J., Nestler, E.J., 2009. Antidepressant actions of histone deacetylase
healthy controls: relation to anhedonia. Psychological Medicine 40, 433–440. inhibitors. The Journal of Neuroscience: the Official Journal of the Society for
Chen, Z., Skolnick, P., 2007. Triple uptake inhibitors: therapeutic potential in depres- Neuroscience 29, 11451–11460.
sion and beyond. Expert Opinion on Investigational Drugs 16, 1365–1377. Covington 3rd., H.E., Vialou, V.F., LaPlant, Q., Ohnishi, Y.N., Nestler, E.J., 2011.
Chen, Z.Y., Jing, D., Bath, K.G., Ieraci, A., Khan, T., Siao, C.J., Herrera, D.G., Toth, Hippocampal-dependent antidepressant-like activity of histone deacetylase
M., Yang, C., McEwen, B.S., Hempstead, B.L., Lee, F.S., 2006. Genetic variant inhibition. Neuroscience Letters 493, 122–126.
P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371 2361

Croarkin, P.E., Levinson, A.J., Daskalakis, Z.J., 2011. Evidence for GABAergic inhibitory serotonergic consequences of decreasing or increasing hippocampus brain-
deficits in major depressive disorder. Neuroscience and Biobehavioral Reviews derived neurotrophic factor protein levels in mice. Neuropharmacology 55,
35, 818–825. 1006–1014.
Cryan, J.F., Holmes, A., 2005. The ascent of mouse: advances in modelling human Deng, W., Aimone, J.B., Gage, F.H., 2010. New neurons and new memories: how does
depression and anxiety. Nature Reviews. Drug Discovery 9, 775–790. adult hippocampal neurogenesis affect learning and memory? Nature Reviews.
Cryan, J.F., Slattery, D.A., 2010. GABAB receptors and depression. Current Status. Neuroscience 11, 339–350.
Advances in Pharmacology 58, 427–451. DeRubeis, R.J., Siegle, G.J., Hollon, S.D., 2008. Cognitive therapy versus medication
Cryan, J.F., O’Leary, O.F., Jin, S.H., Friedland, J.C., Ouyang, M., Hirsch, B.R., Page, for depression: treatment outcomes and neural mechanisms. Nature Reviews.
M.E., Dalvi, A., Thomas, S.A., Lucki, I., 2004. Norepinephrine-deficient mice Neuroscience 9, 788–796.
lack responses to antidepressant drugs, including selective serotonin reuptake DeVane, C.L., Gill, H.S., 1997. Clinical pharmacokinetics of fluvoxamine: applica-
inhibitors. Proceedings of the National Academy of Sciences of the United States tions to dosage regimen design. The Journal of Clinical Psychiatry 58 (Suppl. 5),
of America 101, 8186–8191. 7–14.
Cullinan, W.E., Ziegler, D.R., Herman, J.P., 2008. Functional role of local GABAergic Dias-Ferreira, E., Sousa, J.C., Melo, I., Morgado, P., Mesquita, A.R., Cerqueira, J.J., Costa,
influences on the HPA axis. Brain Structure & Function 213, 63–72. R.M., Sousa, N., 2009. Chronic stress causes frontostriatal reorganization and
Cybulska-Klosowicz, A., Zakrzewska, R., Kossut, M., 2009. Brain activation patterns affects decision-making. Science 325, 621–625.
during classical conditioning with appetitive or aversive UCS. Behavioural Brain Diaz, S.L., Doly, S., Narboux-Nême, N., Fernández, S., Mazot, P., Banas, S.M., Bou-
Research 204, 102–111. tourlinsky, K., Moutkine, I., Belmer, A., Roumier, A., Maroteaux, L., 2012. 5-HT(2B)
Czeh, B., Lucassen, P.J., 2007. What causes the hippocampal volume decrease in receptors are required for serotonin-selective antidepressant actions. Molecular
depression? Are neurogenesis, glial changes and apoptosis implicated? Euro- Psychiatry 17, 154–163.
pean Archives of Psychiatry and Clinical Neuroscience 257, 250–260. Diazgranados, N., Ibrahim, L., Brutsche, N.E., Newberg, A., Kronstein, P., Khalife, S.,
Czéh, B., Michaelis, T., Watanabe, T., Frahm, J., de Biurrun, G., van Kampen, M., Bar- Kammerer, W.A., Quezado, Z., Luckenbaugh, D.A., Salvadore, G., Machado-Vieira,
tolomucci, A., Fuchs, E., 2001. Stress-induced changes in cerebral metabolites, R., Manji, H.K., Zarate Jr., C.A., 2010. A randomized add-on trial of an N-methyl-
hippocampal volume, and cell proliferation are prevented by antidepressant D-aspartate antagonist in treatment-resistant bipolar depression. Archives of
treatment with tianeptine. Proceedings of the National Academy of Sciences of General Psychiatry 67, 793–802.
the United States of America 98, 12796–12801. Di Chiara, G., Loddo, P., Tanda, G., 1999. Reciprocal changes in prefrontal and limbic
Czéh, B., Welt, T., Fischer, A.K., Erhardt, A., Schmitt, W., Muller, M.B., Toschi, dopamine responsiveness to aversive and rewarding stimuli after chronic mild
N., Fuchs, E., Keck, M.E., 2002. Chronic psychosocial stress and con- stress: implications for the psychobiology of depression. Biological Psychiatry
comitant repetitive transcranial magnetic stimulation: effects on stress 46, 1624–1633.
hormone levels and adult hippocampal neurogenesis. Biological Psychiatry 52, Dichter, G.S., Felder, J.N., Smoski, M.J., 2010. The effects of brief behavioral activa-
1057–1065. tion therapy for depression on cognitive control in affective contexts: an fMRI
Dallaspezia, S., Benedetti, F., 2011. Chronobiological therapy for mood disorders. investigation. Journal of Affective Disorders 126, 236–244.
Expert Review of Neurotherapeutics 11, 961–970. Dickerson, S.S., Gable, S.L., Irwin, M.R., Aziz, N., Kemeny, M.E., 2009. Social-evaluative
David, D.J., Klemenhagen, K.C., Holick, K.A., Saxe, M.D., Mendez, I., Santarelli, threat and proinflammatory cytokine regulation: an experimental laboratory
L., Craig, D.A., Zhong, H., Swanson, C.J., Hegde, L.G., Ping, X.I., Dong, investigation. Psychological Sciences: A Journal of the American Psychological
D., Marzabadi, M.R., Gerald, C.P., Hen, R., 2007. Efficacy of the MCHR1 Society 20, 1237–1244.
antagonist N-[3-(1-{[4-(3,4-difluorophenoxy)phenyl]methyl}(4-piperidyl))-4- Dienes, K.A., Hammen, C., Henry, R.M., Cohen, A.N., Daley, S.E., 2006. The stress sen-
methylphen yl]-2-methylpropanamide (SNAP 94847) in mouse models of sitization hypothesis: understanding the course of bipolar disorder. Journal of
anxiety and depression following acute and chronic administration is inde- Affective Disorders 95, 43–49.
pendent of hippocampal neurogenesis. The Journal of Pharmacology and Dillon, D.G., Holmes, A.J., Birk, J.L., Brooks, N., Lyons-Ruth, K., Pizzagalli, D.A., 2009.
Experimental Therapeutics 321, 237–248. Childhood adversity is associated with left basal ganglia dysfunction during
David, D.J., Samuels, B.A., Rainer, Q., Wang, J.W., Marsteller, D., Mendez, I., Drew, M., reward anticipation in adulthood. Biological Psychiatry 66, 206–213.
Craig, D.A., Guiard, B.P., Guilloux, J.P., Artymyshyn, R.P., Gardier, A.M., Gerald, Di Simplicio, M., Norbury, R., Harmer, C.J., 2011. Short-term antidepressant admin-
C., Antonijevic, I.A., Leonardo, E.D., Hen, R., 2009. Neurogenesis-dependent and istration reduces negative self-referential processing in the medial prefrontal
-independent effects of fluoxetine in an animal model of anxiety/depression. cortex in subjects at risk for depression. Molecular Psychiatry 17, 502–510.
Neuron 62, 479–493. Disner, S.G., Beevers, C.G., Haigh, E.A., Beck, A.T., 2011. Neural mechanisms
Davidson, R.J., 2000. Affective style, psychopathology, and resilience: brain mecha- of the cognitive model of depression. Nature Reviews. Neuroscience 12,
nisms and plasticity. The American Psychologist 55 (11), 1196–1214. 467–477.
Davila, J., Stroud, C.B., Starr, L.R., Miller, M.R., Yoneda, A., Hershenberg, R., Dombrovski, Y., Siegle, G.J., Szanto, K., Clark, L., Reynolds, C.F., Aizensteim, H.,
2009. Romantic and sexual activities, parent-adolescent stress, and depres- 2011. The temptation of suicide: striatal gray matter, discounting of delayed
sive symptoms among early adolescent girls. Journal of Adolescence 32, rewards, and suicide attempts in late-life depression. Psychological Medicine 17,
909–924. 1–13.
de Kloet, E.R., Joels, M., Holsboer, F., 2005. Stress and the brain: from adaptation to Domschke, K., Dannlowski, U., Ohrmann, P., Lawford, B., Bauer, J., Kugel, H.,
disease. Nature Reviews. Neuroscience 6, 463–475. Heindel, W., Young, R., Morris, P., Arolt, V., Deckert, J., Suslow, T., Baune,
Delgado, P.L., Charney, D.S., Price, L.H., Aghajanian, G.K., Landis, H., Heninger, G.R., B.T., 2008. Cannabinoid receptor 1 (CNR1) gene: impact on antidepressant
1990. Serotonin function and the mechanism of antidepressant action: reversal treatment response and emotion processing in major depression. European
of antidepressant induced remission by rapid depletion of plasma tryptophan. Neuropsychopharmacology: the Journal of the European College of Neuropsy-
Archives of General Psychiatry 47, 411–418. chopharmacology 18, 751–759.
Delgado, P.L., Miller, H.M., Salomon, R.M., Licinio, J., Gelenberg, A.J., Charney, D.S., Drevets, W.C., Savitz, J., Trimble, M., 2008. The subgenual anterior cingulate cortex
1993. Monoamines and the mechanism of antidepressant action: effects of in mood disorders. CNS Spectrums 13, 663–681.
catecholamine depletion on mood in patients treated with antidepressants. Duman, R.S., Monteggia, L.M., 2006. A neurotrophic model for stress-related mood
Psychopharmacology Bulletin 29, 389–396. disorders. Biological Psychiatry 59, 1116–1127.
Delgado, P.L., Miller, H.L., Salomon, R.M., Licinio, J., Krystal, J.H., Moreno, Duman, R.S., Vaidya, V.A., 1998. Molecular and cellular actions of chronic electro-
F.A., Heninger, G.R., Charney, D.S., 1999. Tryptophan-depletion challenge in convulsive seizures. The Journal of ECT 14, 181–193.
depressed patients treated with desipramine or fluoxetine: implications for the Duman, C.H., Schlesinger, L., Kodama, M., Russell, D.S., Duman, R.S., 2007. A role for
role of serotonin in the mechanism of antidepressant action. Biological Psychi- MAP kinase signaling in behavioral models of depression and antidepressant
atry 46, 212–220. treatment. Biological Psychiatry 61, 661–670.
Delgado, Palacios, R., Campo, A., Henningsen, K., Verhoye, M., Poot, D., Dijkstra, J., Duman, R.S., Li, N., Liu, R.J., Duric, V., Aghajanian, G., 2012. Signaling pathways
Van Audekerke, J., Benveniste, H., Sijbers, J., Wiborg, O., Van der Linden, A., 2011. underlying the rapid antidepressant actions of ketamine. Neuropharmacology
Magnetic resonance imaging and spectroscopy reveal differential hippocampal 62, 35–41.
changes in anhedonic and resilient subtypes of the chronic mild stress rat model. Duncan, G.E., Breese, G.R., Criswell, H., Stumpf, W.E., Mueller, R.A., Covey, J.B.,
Biological Psychiatry 70, 449–457. 1986. Effects of antidepressant drugs injected into the amygdala on behav-
Dell’osso, B., Camuri, G., Castellano, F., Vecchi, V., Benedetti, M., Bortolussi, S., ioral responses of rats in the forced swim test. The Journal of Pharmacology
Altamura, A.C., 2011. Meta-review of Metanalytic Studies with Repetitive Trans- and Experimental Therapeutics 238, 758–762.
cranial Magnetic Stimulation (rTMS) for the treatment of major depression. Dunkin, J.J., Leuchter, A.F., Cook, I.A., Kasl-Godley, J.E., Abrams, M., Rosenberg-
Clinical Practice and Epidemiology in Mental Health 7, 167–177. Thompson, S., 2000. Executive dysfunction predicts nonresponse to fluoxetine
de Maat, S.M., Dekker, J., Schoevers, R.A., de Jonghe, F., 2007. Relative efficacy in major depression. Journal of Affective Disorders 60, 13–23.
of psychotherapy and combined therapy in the treatment of depression: a Dunn, A.J., Swiergiel, A.H., de Beaurepaire, R., 2005. Cytokines as mediators of depres-
meta-analysis. European Psychiatry: the Journal of the Association of European sion: what can we learn from animal studies? Neuroscience and Biobehavioral
Psychiatrists 22, 1–8. Reviews 29, 891–909.
de Quervain, D.J., Roozendaal, B., McGaugh, J.L., 1998. Stress and glucocorticoids Dunn, R.T., Kimbrell, T.A., Ketter, T.A., Frye, M.A., Willis, M.W., Luckenbaugh, D.A.,
impair retrieval of long-term spatial memory. Nature 394, 787–790. Post, R.M., 2002. Principal components of the Beck Depression Inventory and
de Quervain, D.J., Roozendaal, B., Nitsch, R.M., McGaugh, J.L., Hock, C., 2000. Acute regional cerebral metabolism in unipolar and bipolar depression. Biological Psy-
cortisone administration impairs retrieval of longterm declarative memory in chiatry 51, 387–399.
humans. Nature Neuroscience 3, 313–314. Duvarci, S., Pare, D., 2007. Glucocorticoids enhance the excitability of principal baso-
Deltheil, T., Guiard, B.P., Cerdan, J., David, D.J., Tanaka, K.F., Repérant, C., lateral amygdala neurons. The Journal of Neuroscience: the Official Journal of the
Guilloux, J.P., Coudoré, F., Hen, R., Gardier, A.M., 2008. Behavioral and Society for Neuroscience 27, 4482–4491.
2362 P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371

Dwyer, D.M., 2011. Lesions of the basolateral, but not central, amygdala impair Frodl, T., Schaub, A., Banac, S., Charypar, M., Jager, M., Kummler, P., Bottlender, R.,
flavour-taste learning based on fructose or quinine reinforcers. Behavioural Zetzsche, T., Born, C., Leinsinger, G., Reiser, M., Möller, H.J., Meisenzahl, E.M.,
Brain Research 220, 349–353. 2006. Reduced hippocampal volume correlates with executive dysfunctioning
Dziedzicka-Wasylewska, M., Willner, P., Papp, M., 1997. Changes in dopamine recep- in major depression. Journal of Psychiatry & Neuroscience 31, 316–323.
tor mRNA expression following chronic mild stress and chronic antidepressant Frodl, T., Jager, M., Smajstrlova, I., Born, C., Bottlender, R., Palladino, T., Reiser, M.,
treatment. Behavioural Pharmacology 8, 607–618. Möller, H.J., Meisenzahl, E.M., 2008a. Effect of hippocampal and amygdala vol-
Egner, T., Etkin, A., Gale, S., Hirsch, J., 2008. Dissociable neural systems resolve umes on clinical outcomes in major depression: a 3-year prospective magnetic
conflict from emotional versus nonemotional distracters. Cerebral Cortex 18, resonance imaging study. Journal of Psychiatry & Neuroscience 33, 423–430.
1475–1484. Frodl, T., Möller, H.J., Meisenzahl, E., 2008b. Neuroimaging genetics: new per-
Eisch, A.J., Bolanos, C.A., de Wit, J., Simonak, R.D., Pudiak, C.M., Barrot, M., Verhaagen, spectives in research on major depression? Acta Psychiatrica Scandinavica 18,
J., Nestler, E.J., 2003. Brain-derived neurotrophic factor in the ventral midbrain- 363–372.
nucleus accumbens pathway: a role in depression. Biological Psychiatry 54, Frokjaer, V.G., Mortensen, E.L., Nielsen, F.A., Haugbol, S., Pinborg, L.H., Adams,
994–1005. K.H., Svarer, C., Hasselbalch, S.G., Holm, S., Paulson, O.B., Knudsen, G.M.,
Elizalde, N., Gil-Bea, F.J., Ramírez, M.J., Aisa, B., Lasheras, B., Del Rio, J., Tordera, R.M., 2008.Frontolimbic serotonin 2A receptor binding in healthy subjects is asso-
2008. Long-lasting behavioral effects and recognition memory deficit induced ciated with personality risk factors for affective disorder. Biological Psychiatry
by chronic mild stress in mice: effect of antidepressant treatment. Psychophar- 63, 569–576.
macology 199, 1–14. Frokjaer, V.G., Vinberg, M., Erritzoe, D., Svarer, C., Baaré, W., Budtz-Joergensen, E.,
Elizalde, N., García-García, A.L., Totterdell, S., Gendive, N., Venzala, E., Ramirez, M.J., Madsen, K., Madsen, J., Kessing, L.V., Knudsen, G.M., 2009. High familial risk for
Del Rio, J., Tordera, R.M., 2010. Sustained stress-induced changes in mice as a mood disorder is associated with low dorsolateral prefrontal cortex serotonin
model for chronic depression. Psychopharmacology 210, 393–406. transporter binding. NeuroImage 46, 360–366.
Elliott, R., Sahakian, B.J., McKay, A.P., Herrod, J.J., Robbins, T.W., Paykel, E.S., Frokjaer, V.G., Vinberg, M., Erritzoe, D., Baaré, W., Holst, K.K., Mortensen, E.L., Arfan,
1996. Neuropsychological impairments in unipolar depression: the influence H., Madsen, J., Jernigan, T.L., Kessing, L.V., Knudsen, G.M., 2010. Familial risk
of perceived failure on subsequent performance. Psychological Medicine 26, for mood disorder and the personality risk factor, neuroticism, interact in their
975–989. association with frontolimbic serotonin 2A receptor binding. Neuropsychophar-
Elliott, R., Zahn, R., Deakin, J.F., Anderson, I.M., 2011. Affective cogni- macology 35, 1129–1137.
tion and its disruption in mood disorders. Neuropsychopharmacology 36, Fu, C.H., Williams, S.C., Brammer, M.J., Suckling, J., Kim, J., Cleare, A.J., Walsh, N.D.,
153–182. Mitterschiffthaler, M.T., Andrew, C.M., Pich, E.M., Bullmore, E.T., 2007. Neural
Enns, M.W., Cox, B.J., 1997. Personality dimensions and depression: review and responses to happy facial expressions in major depression following antide-
commentary. Canadian Journal of Psychiatry 42, 274–284. pressant treatment. The American Journal of Psychiatry 164, 599–607.
Epstein, J., Pan, H., Kocsis, J.H., Yang, Y., Butler, T., Chusid, J., Hochberg, H., Mur- Fujikawa, T., Yamawaki, S., Touhouda, Y., 1993. Incidence of silent cerebral infarction
rough, J., Strohmayer, E., Stern, E., Silbersweig, D.A., 2006. Lack of ventral striatal in patients with major depression. Stroke 24, 1631–1634.
response to positive stimuli in depressed versus normal subjects. The American Furman, D.J., Hamilton, J.P., Joormann, J., Gotlib, I.H., 2011. Altered timing of amyg-
Journal of Psychiatry 163, 1784–1790. dala activation during sad mood elaboration as a function of 5-HTTLPR. Social,
Eriksson, P.S., Perfilieva, E., Bjork-Eriksson, T., Alborn, A.M., Nordborg, C., Peterson, Cognitive and Affective Neuroscience 6, 270–276.
D.A., Gage, F.H., 1998. Neurogenesis in the adult human hippocampus. Nature Furr, A., Lapiz-Bluhm, M.D., Morilak, D.A., 2011. 5-HT2A receptors in the
Medicine 4, 1313–1317. orbitofrontal cortex facilitate reversal learning and contribute to the benefi-
Ernst, M., Paulus, M.P., 2005. Neurobiology of decision making: a selective review cial cognitive effects of chronic citalopram treatment in rats. The International
from a neurocognitive and clinical perspective. Biological Psychiatry 58, Journal of Neuropsychopharmacology 15, 1295–1305.
597–604. Gabbay, V., Hess, D., Liu, S., Babb, J.S., Klein, R.G., Gonen, O., 2007. Lateralized
Essex, M.J., Shirtcliff, E.A., Burk, L.R., Ruttle, P.L., Klein, M.H., Slattery, M.J., Kalin, caudate metabolic abnormalities in adolescent major depressive disorder:
N.H., Armstrong, J.M., 2011. Influence of early life stress on later hypothalamic- a proton MR spectroscopy study. The American Journal of Psychiatry 164,
pituitary-adrenal axis functioning and its covariation with mental health 1881–1889.
symptoms: a study of the allostatic process from childhood into adolescence. Galynker, I.I., Yaseen, Z.S., Katz, C., Zhang, X., Jennings-Donovan, G., Dashnaw, S.,
Development and Psychopathology 23, 1039–1058. Hirsch, J., Mayberg, H., Cohen, L.J., Winston, A., 2011. Distinct but overlapping
Estrada-Camarena, E., López-Rubalcava, C., Vega-Rivera, N., Récamier-Carballo, S., neural networks subserve depression and insecure attachment. Social, Cognitive
Fernández-Guasti, A., 2010. Antidepressant effects of estrogens: a basic approx- and Affective Neuroscience 7, 896–908.
imation. Behavioural Pharmacology 21, 451–464. Garrigou, D., Broekkamp, C.L., Lloyd, K.G., 1981. Involvement of the amygdala in the
Etkin, A., Wager, T.D., 2007. Functional neuroimaging of anxiety: a meta-analysis of effect of antidepressants on the passive avoidance deficit in bulbectomised rats.
emotional processing in PTSD, social anxiety disorder, and specific phobia. The Psychopharmacology 74, 66–70.
American Journal of Psychiatry 164, 1476–1488. Gartside, S.E., Umbers, V., Sharp, T., 1997. Inhibition of 5HT firing in the
Eugene, F., Joormann, J., Cooney, R.E., Atlas, L.Y., Gotlib, I.H., 2010. Neural correlates DRN by non-selective 5-HT reuptake inhibitors: studies on the role of 5-
of inhibitory deficits in depression. Psychiatry Research 181, 30–35. HT1A autoreceptors and noradrenergic mechanisms. Psychopharmaclogy 130,
Farb, N.A., Anderson, A.K., Bloch, R.T., Segal, Z.V., 2011. Mood-linked responses in 261–268.
medial prefrontal cortex predict relapse in patients with recurrent unipolar Gaynes, B.N., Dusetzina, S.B., Ellis, A.R., Hansen, R.A., Farley, J.F., Miller, W.C., Stürmer,
depression. Biological Psychiatry 70, 366–372. T., 2012. Treating depression after initial treatment failure: directly comparing
Farmer, A.E., McGuffin, P., 2003. Humiliation, loss and other types of life events switch and augmenting strategies in STAR*D. Journal of Clinical Psychopharma-
and difficulties: a comparison of depressed subjects, healthy controls and their cology 32, 114–119.
siblings. Psychological Medicine 33, 1169–1175. Ghose, S., Winter, M.K., McCarson, K.E., Tamminga, C.A., Enna, S.J., 2011. The GABA-B
Faure, A., Reynolds, S.M., Richard, J.M., Berridge, K.C., 2008. Mesolimbic dopamine receptor as a target for antidepressant drug action. British Journal of Pharma-
in desire and dread: enabling motivation to be generated by localized gluta- cology 162, 1–17.
mate disruptions in nucleus accumbens. The Journal of Neuroscience: the Official Gilbert, P., 2006. Evolution and depression: issues and implications. Psychological
Journal of the Society for Neuroscience 28, 7184–7192. Medicine 36, 287–297.
Fava, M., Mischoulon, D., 2009. Folate in depression: efficacy, safety, differences in Gilbody, S., Lightfoot, T., Sheldon, T., 2007. Is low folate a risk factor for depression?
formulations, and clinical issues. The Journal of Clinical Psychiatry 70 (Suppl. 5), A meta-analysis and exploration of heterogeneity. Journal of Epidemiology and
12–17. Community Health 61, 631–637.
Feyissa, A.M., Woolverton, W.L., Miguel-Hidalgo, J.J., Wang, Z., Kyle, P.B., Hasler, G., Gill, K.M., Grace, A.A., 2011. Heterogeneous processing of amygdala and hippocam-
Stockmeier, C.A., Iyo, A.H., Karolewicz, B., 2010. Elevated level of metabotropic pal inputs in the rostral and caudal subregions of the nucleus accumbens. The
glutamate receptor 2/3 in the prefrontal cortex in major depression. Progress in International Journal of Neuropsychopharmcology 14, 1301–1314.
Neuro-psychopharmacology & Biological Psychiatry 34, 279–283. Gold, M.S., Pottash, A.L., Extein, I., 1981. Hypothyroidism and depression. Evidence
Fogel, J., Eaton, W.W., Ford, D.E., 2006. Minor depression as a predictor of the first from complete thyroid function evaluation. JAMA: the Journal of the American
onset of major depressive disorder over a 15-year follow-up. Acta Psychiatrica Medical Association 245, 1919–1922.
Scandinavica 113, 36–43. Goldman-Rakic, P.S., Selemon, L.D., Schwartz, M.L., 1984. Dual pathways connecting
French, S.J., Totterdell, S., 2002. Hippocampal and prefrontal cortical inputs the dorsolateral prefrontal cortex with the hippocampal formation and parahip-
monosynaptically converge with individual projection neurons of the nucleus pocampal cortex in the rhesus monkey. Neuroscience 12, 719–743.
accumbens. The Journal of Comparative Neurology 446, 151–165. Goodwin, F.K., Ebert, M.H., Bunney, W.E., 1972. Mental effects of reserpine in man: a
French, S.J., Totterdell, S., 2003. Indidividual nucleus accumbens-projection neu- review. In: Shader, R.I. (Ed.), Psychiatric Complications of Medical Drugs. Raven,
tons receive both basolateral amygdala and ventral subicular afferences in rats. New York, pp. 73–101.
Neuroscience 119, 19–31. Gorlyn, M., Keilp, J.G., Grunebaum, M.F., Taylor, B.P., Oquendo, M.A., Bruder, G.E.,
Friedman, A., Lax, E., Dikshtein, Y., Abraham, L., Flaumenhaft, Y., Sudai, E., Ben- Stewart, J.W., Zalsman, G., Mann, J.J., 2008. Neuropsychobiological characteris-
Tzion, M., Yadid, G., 2011. Electrical stimulation of the lateral habenula produces tics as predictors of SSRI treatment response in depressed subjects. Journal of
an inhibitory effect on sucrose self-administration. Neuropharmacology 60, Neural Transmission 115, 1213–1219.
381–387. Gorzalka, B.B., Hill, M.N., 2011. Putative role of endocannabinoid signaling in
Frodl, T., Meisenzahl, E.M., Zetzsche, T., Born, C., Jäger, M., Groll, C., Bottlender, R., the etiology of depression and actions of antidepressants. Progress in Neuro-
Leinsinger, G., Möller, H.J., 2003. Larger amygdala volumes in first depressive psychopharmacology & Biological Psychiatry 35, 1575–1585.
episode as compared to recurrent major depression and healthy control subjects. Gotlib, I.H., Joormann, J., 2010. Cognition and depression: current status and future
Biological Psychiatry 53, 338–344. directions. Annual Review of Clinical Psychology 27, 285–312.
P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371 2363

Gotlib, I.H., Joormann, J., Minor, K.L., Hallmayer, J., 2008. HPA axis reactivity: a mech- Harmer, C.J., 2008. Serotonin and emotional processing: does it help explain antide-
anism underlying the associations among 5-HTTLPR, stress, and depression. pressant drug action? Neuropharmacology 55, 1023–1028.
Biological Psychiatry 63, 847–851. Harmer, C.J., Hill, S.A., Taylor, M.J., Cowen, P.J., Goodwin, G.M., 2003a. Towards a
Gould, E., 2007. How widespread is adult neurogenesis in mammals? Nature nneuropsychological theory of antidepressant drug action: potentiation of nor-
Reviews. Neuroscience 8, 481–488. epinephrine activity increases positive emotional bias. The American Journal of
Gould, E., McEwen, B.S., Tanapat, P., Galea, L.A., Fuchs, E., 1997. Neurogenesis in the Psychiatry 160, 990–992.
dentate gyrus of the adult tree shrew is regulated by psychosocial stress and Harmer, C.J., Bhagwagar, Z., Perrett, D.I., Vollm, B.A., Cowen, P.J., Goodwin, G.M.,
NMDA receptor activation. The Journal of Neuroscience: the Official Journal of 2003b. Acute SSRI administration affects the processing of social cues in healthy
the Society for Neuroscience 17, 2492–2498. volunteers. Neuropsychopharmacology 28, 148–152.
Gould, E., Tanapat, P., McEwen, B.S., Flugge, G., Fuchs, E., 1998. Proliferation of gran- Harmer, C.J., Shelley, N.C., Cowen, P.J., Goodwin, G.M., 2004. Increased positive versus
ule cell precursors in the dentate gyrus of adult monkeys is diminished by stress. negative affective perception and memory in healthy volunteers following selec-
Proceedings of the National Academy of Sciences of the United States of America tive serotonin and norepinephrine reuptake inhibition. The American Journal of
95, 3168–3171. Psychiatry 161, 1256–1263.
Goveas, J.S., Xie, C., Ward, B.D., Wu, Z., Li, W., Franczak, M., Jones, J.L., Antuono, Harmer, C.J., Mackay, C.E., Reid, C.B., Cowen, P.J., Goodwin, G.M., 2006. Antidepres-
P.G., Li, S.J., 2011a. Recovery of hippocampal network connectivity correlates sant drug treatment modifies the neural processing of nonconscious threat cues.
withcognitive improvement in mild Alzheimer’s disease patients treated with Biological Psychiatry 59, 816–820.
donepezil assessed by resting-state fMRI. Journal of Magnetic Resonance Imag- Harmer, C.J., Goodwin, G.M., Cowen, P.J., 2009a. Why do antidepressants take so
ing: JMRI 34, 764–773. long to work? A cognitive neuropsychological model of antidepressant drug
Goveas, J., Xie, C., Wu, Z., Douglas, Ward, B., Li, W., Franczak, M.B., Jones, J.L., Antuono, action. The British Journal of Psychiatry: the Journal of Mental Science 195,
P.G., Yang, Z., Li, S.J., 2011b. Neural correlates of the interactive relationship 102–108.
between memory deficits and depressive symptoms in nondemented elderly: Harmer, C.J., O’Sullivan, U., Favaron, E., Massey-Chase, R., Ayres, R., Reinecke, A.,
resting-state fMRI study. Behavioural Brain Research 219, 205–212. Goodwin, G.M., Cowen, P.J., 2009b. Effect of acute antidepressant administra-
Grasmäder, K., Verwohlt, P.L., Rietschel, M., Dragicevic, A., Müller, M., Hiemke, C., tion on negative affective bias in depressed patients. The American Journal of
et al., 2004. Impact of polymorphisms of cytochrome-P450 isoenzymes 2C9, Psychiatry 166, 1178–1184.
2C19 and 2D6 on plasma concentrations and clinical effects of antidepressants Hashimoto, K., 2011. A comprehensive review of the role of glutamate in the patho-
in a naturalistic clinical setting. European Journal of Clinical Pharmacology 60, physiology of depression and the mechanism of action of antidepressants.
329–336. The role of glutamate on the action of antidepressants. Progress in Neuro-
Green, T.A., Alibhai, I.N., Unterberg, S., Neve, R.L., Ghose, S., Tamminga, C.A., Nestler, psychopharmacology & Biological Psychiatry 35, 1558–1568.
E.J., 2008. Induction of activating transcription factors (ATFs) ATF2, ATF3, and Hasler, G., Northoff, G., 2011. Discovering imaging endophenotypes for
ATF4 in the nucleus accumbens and their regulation of emotional behavior. The major depression. Molecular Psychiatry 16, 604–619, http://dx.doi.
Journal of Neuroscience: the Official Journal of the Society for Neuroscience 28, org/10.1038/mp.2011.23.
2025–2032. Hasler, G., van der Veen, J.W., Tumonis, T., Meyers, N., Shen, J., Drevets, W.C., 2007.
Grønli, J., Bramham, C., Murison, R., Kanhema, T., Fiske, E., Bjorvatn, B., Ursin, R., Por- Reduced prefrontal glutamate/glutamine and gamma-aminobutyric acid levels
tas, C.M., 2006. Chronic mild stress inhibits BDNF protein expression and CREB in major depression determined using proton magnetic resonance spectroscopy.
activation in the dentate gyrus but not in the hippocampus proper. Pharmacol- Archives of General Psychiatry 64, 193–200.
ogy, Biochemistry, and Behavior 85, 842–849. Hasler, G., Fromm, S., Carlson, P.J., Luckenbaugh, D.A., Waldeck, T., Geraci, M.,
Gross, J.J., 2002. Emotion regulation: affective, cognitive, and social consequences. Roiser, J.P., Neumeister, A., Meyers, N., Charney, D.S., Drevets, W.C., 2008. Neu-
Psychophysiology 39 (3), 281–291. ral response to catecholamine depletion in unmedicated subjects with major
Grubert, C., Hurlemann, R., Bewernick, B.H., Kayser, S., Hadrysiewicz, B., Axmacher, depressive disorder in remission and healthy subjects. Archives of General Psy-
N., Sturm, V., Schlaepfer, T.E., 2011. Neuropsychological safety of nucleus chiatry 65, 521–531.
accumbens deep brain stimulation for major depression: effects of 12- Hasler, G., Luckenbaugh, D.A., Snow, J., Meyers, N., Waldeck, T., Geraci, M., Roiser,
month stimulation. The World Journal of Biological Psychiatry: the Official J., Knutson, B., Charney, D.S., Drevets, W.C., 2009. Reward processing after cate-
Journal of the World Federation of Societies of Biological Psychiatry 12, cholamine depletion in unmedicated, remitted subjects with major depressive
516–527. disorder. Biological Psychiatry 66, 201–205.
Gusnard, D.A., Akbudak, E., Shulman, G.L., Raichle, M.E., 2001. Medial prefrontal Heim, C., Newport, D.J., Mletzko, T., Miller, A.H., Nemeroff, C.B., 2008. The link
cortex and selfreferential mental activity: relation to a default mode of brain between childhood trauma and depression: insights from HPA axis studies in
function. Proceedings of the National Academy of Sciences of the United States humans. Psychoneuroendocrinology 33, 693–710.
of America 98, 4259–4264. Heller, A.S., Johnstone, T., Shackman, A.J., Light, S.N., Peterson, M.J., Kolden, G.G.,
Haber, S.N., 2003. The primate basal ganglia: parallel and integrative networks. Kalin, N.H., Davidson, R.J., 2009. Reduced capacity to sustain positive emotion
Journal of Chemical Neuroanatomy 26, 317–330. in major depression reflects diminished maintenance of fronto-striatal brain
Haber, S.N., Knutson, B., 2010. The reward circuit: linking primate anatomy and activation. Proceedings of the National Academy of Sciences of the United States
human imaging. Neuropsychopharmacology 35, 4–26. of America 106, 22445–22450.
Haber, S.N., Kim, K.-S., Mailly, P., Calzavara, R., 2006. Reward-related cortical Henseler, I., Falkai, P., Gruber, O., 2010. Disturbed functional connectivity within
inputs define a large striatal region in primates that interface with associa- brain networks subserving domain-specific subcomponents of working mem-
tive cortical connections, providing a substrate for incentive-based learning. The ory in schizophrenia: relation to performance and clinical symptoms. Journal of
Journal of Neuroscience: the Official Journal of the Society for Neuroscience 26, Psychiatric Research 44, 364–372.
8368–8376. Herman, J.P., Adams, D., Prewitt, C., 1995. Regulatory changes in neuroendocrine
Haeffel, G.J., 2010. When self-help is no help: traditional cognitive skills training does stress-integrative circuitry produced by a variable stress paradigm. Neuroen-
not prevent depressive symptoms in people who ruminate. Behaviour Research docrinology 61, 180–190.
and Therapy 48, 152–157. Herman, J.P., Prewitt, C.M.F., Cullinan, W.E., 1996. Neuronal circuit regulation of the
Hahn, H., Wadsak, W., Windischberger, C., Baldinger, P., Hoflich, A.S., Losack, J., hypothalamo-pituitary-adrenocortical stress axis. Critical Reviews in Neurobi-
et al., 2012. Differential modulation of the default mode network via serotonin- ology 10, 371–394.
1A receptors. Proceedings of the National Academy of Sciences, USA 109, Herman, J.P., Cullinan, W.E., Ziegler, D.R., Tasker, J.G., 2002. Role of the paraventric-
2619–2624. ular nucleus microenvironment in stress integration. The European Journal of
Hajos, M., Richards, C.D., Szekely, A.D., Sharp, T., 1998. An electrophysiological and Neuroscience 16, 381–385.
neuroanatomical study of the medial prefrontal cortical projection to the mid- Herman, J.P., Ostrander, M.M., Mueller, N.K., Figueiredo, H., 2005. Limbic system
brain raphe nuclei in the rat. Neuroscience 87, 95–108. mechanisms of stress regulation: hypothalamo-pituitary-adrenocortical axis.
Hamani, C., Diwan, M., Macedo, C.E., Brandão, M.L., Shumake, J., Gonzalez-Lima, F., Progress in Neuro-psychopharmacology & Biological Psychiatry 29, 1201–1213.
Raymond, R., Lozano, A.M., Fletcher, P.J., Nobrega, J.N., 2010. Antidepressant- Herry, C., Ciocchi, S., Senn, V., Demmou, L., Muller, C., Luthi, A., 2008. Switching on
like effects of medial prefrontal cortex deep brain stimulation in rats. Biological and off fear by distinct neuronal circuits. Nature 454, 600–606.
Psychiatry 67, 117–124. Heuser, I., Yassouridis, A., Holsboer, F., 1994. The combined dexamethasone/CRH
Hamani, C., Mayberg, H., Stone, S., Laxton, A., Haber, S., Lozano, A.M., 2011. The sub- test: a refined laboratory test for psychiatric disorders. Journal of Psychiatric
callosal cingulate gyrus in the context of major depression. Biological Psychiatry Research 28, 341–356.
69, 301–308. Hikosaka, O., 2010. The habenula: from stress-evasion to value-based decision mak-
Hamilton, J.P., Furman, D.J., Chang, C., Thomason, M.E., Dennis, E., Gotlib, I.H., 2011. ing. Nature Reviews Neuroscience 11, 503–513.
Default-mode and task-positive network activity in major depressive disorder: Hikosaka, O., Sesack, S.R., Lecourtier, L., Shepard, P.D., 2008. Habenula: crossroad
implications for adaptive and maladaptive rumination. Biological Psychiatry 70, between the basal ganglia and the limbic system. The Journal of Neuroscience:
327–333. the Official Journal of the Society for Neuroscience 28, 11825–11829.
Hammen, C., 2006. Stress generation in depression: reflections on origins, research, Hill, M.N., Gorzalka, B.B., 2009. Impairments in endocannabinoid signaling and
and future directions. Journal of Clinical Psychology 62, 1065–1082. depressive illness. JAMA: the Journal of the American Medical Association 301,
Hanson, N.D., Owens, M.J., Nemeroff, C.B., 2011. Depression, antidepressants, and 1165–1166.
neurogenesis: a critical reappraisal. Neuropsychopharmacology 36, 2589–2602. Hill, M.N., Hillard, C.J., Bambico, F.R., Patel, S., Gorzalka, B.B., Gobbi, G., 2009. The
Harkness, K.L., Monroe, S.M., 2006. Severe melancholic depression is more vulnera- therapeutic potential of the endocannabinoid system for the development
ble than non-melancholic depression to minor precipitating life events. Journal of a novel class of antidepressants. Trends in Pharmacological Sciences 30,
of Affective Disorders 91, 257–263. 484–493.
2364 P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371

Hinwood, M., Morandini, J., Day, T.A., Walker, F.R., 2011. Evidence that microglia Jenness, J.L., Hankin, B.J., Abela, J.R.Z., Young, J.F., Smolen, A., 2011. Chronic fam-
mediate the neurobiological effects of chronic psychological stress on the medial ily stress interacts with 5HTTLPR to predict prospective depressive symptoms
prefrontal cortex. Cerebral Cortex 22, 1442–1454. among youth. Depression and Anxiety 28, 1074–1080.
Hodes, G.E., Yang, L., Van, K.J., Santollo, J., Shors, T.J., 2009. Prozac during puberty: Jiang, W., Zhang, Y., Xiao, L., Van Cleemput, J., Ji, S.P., Bai, G., Zhang, X., 2005. Cannabi-
distinctive effects on neurogenesis as a function of age and sex. Neuroscience noids promote embryonic and adult hippocampus neurogenesis and produce
163, 609–617. anxiolytic- and antidepressant-like effects. The Journal of Clinical Investigation
Holick, K.A., Lee, D.C., Hen, R., Dulawa, S.C., 2008. Behavioral effects of chronic flu- 115, 3104–3116.
oxetine in BALB/cJ mice do not require adult hippocampal neurogenesis or the Jin, K., Zhu, Y., Sun, Y., Mao, X.O., Xie, L., Greenberg, D.A., 2002. Vascular endothelial
serotonin 1A receptor. Neuropsychopharmacology 33, 406–417. growth factor (VEGF) stimulates neurogenesis in vitro and in vivo. Proceedings
Holmes, A., Wellman, C.L., 2009. Stress-induced prefrontal reorganization and exec- of the National Academy of Sciences of the United States of America 99,
utive dysfunction in rodents. Neuroscience and Biobehavioral Reviews 33, 11946–11950.
773–783. Joiner, T.E., Wingate, L.R., Gencoz, T., Gencoz, F., 2005. Stress generation in
Holmes, A.J., Pizzagalli, D.A., 2008. Spatiotemporal dynamics of error processing depression: three studies on its resilience, possible mechanism, and symptom
dysfunctions in major depressive disorder. Archives of General Psychiatry 65, specificity. Journal of Social and Clinical Psychology 24, 236–253.
179–188. Jones, B.F., Groenewegen, H.J., Witter, M.P., 2005. Intrinsic connections of the cingu-
Holmes, A.J., Bogdan, R., Pizzagalli, D.A., 2010. Serotonin transporter genotype and late cortex in the rat suggest the existence of multiple functionally segregated
action monitoring dysfunction: a possible substrate underlying increased vul- networks. Neuroscience 133, 193–207.
nerability to depression. Neuropsychopharmacology 35, 1186–1197. Joormann, J., Gotlib, I.H., 2008. Updating the contents of working memory in
Holsboer, F., 2000. The corticosteroid receptor hypothesis of depression. Neuropsy- depression: interference from irrelevant negative material. Journal of Abnormal
chopharmacology 23, 477–501. Psychology 117, 182–192.
Holsboer, F., 2008. How can we realize the promise of personalized antidepressant Joormann, J., Talbot, L., Gotlib, I.H., 2007. Biased processing of emotional information
medicines? Nature Reviews. Neuroscience 9, 638–646. in girls at risk for depression. Journal of Abnormal Psychology 116, 135–143.
Holsboer, F., Ising, M., 2010. Stress hormone regulation: biological role and transla- Juhasz, G., Chase, D., Pegg, E., Downey, D., Toth, Z.G., Stones, K., Platt, H., Mekli, K.,
tion into therapy. Annual Review of Psychology 61, 81–109, C1-11. Payton, A., Elliott, R., Anderson, I.M., Deakin, J.F., 2009. CNR1 gene is associated
Holtzheimer 3rd., P.E., Mayberg, H.S., 2010. Deep brain stimulation for treatment- with high neuroticism and low agreeableness and interacts with recent negative
resistant depression. The American Journal of Psychiatry 167, 1437–1444. life events to predict current depressive symptoms. Neuropsychopharmacology
Holtzheimer, P.E., Mayberg, H.S., 2011. Stuck in a rut: rethinking depression and its 34, 2019–2027.
treatment. Trends in Neurosciences 34, 1–9. Kandel, E.R., 2001. The molecular biology of memory storage: a dialogue between
Horder, J., Cowen, P.J., Di Simplicio, M., Browning, M., Harmer, C.J., 2009. genes and synapses. Science 294, 1030–1038.
Acute administration of the cannabinoid CB1 antagonist rimonabant impairs Karege, F., Vaudan, G., Schwald, M., Perroud, N., La Harpe, R., 2005. Neurotrophin
positive affective memory in healthy volunteers. Psychopharmacology 205, levels in postmortem brains of suicide victims and the effects of antemortem
85–91. diagnosis and psychotropic drugs. Brain Research. Molecular Brain Research 136,
Horder, J., Harmer, C.J., Cowen, P.J., McCabe, C., 2010. Reduced neural response 29–37.
to reward following 7 days treatment with the cannabinoid CB(1) antagonist Karg, K., Burmeister, M., Shedden, K., Sen, S., 2011. The serotonin transporter
rimonabant in healthy volunteers. The International Journal of Neuropsycho- promoter variant (5-HTTLPR), stress, and depression meta-analysis revisited.
pharmacology 13, 1103–1113. Archives of General Psychiatry 68, 444–454.
Horvitz, A.V., Wakefield, J.C., 2007. The Loss of Sadness: How Psychiatry Transformed Kato, M., Serretti, A., 2010. Review and meta-analysis of antidepressant pharmaco-
Normal Sorrow into Depressive Disorder. Oxford University Press, New York. genetic findings in major depressive disorder. Molecular Psychiatry 15, 473–500.
Horstmann, S., Binder, E.B., 2009. Pharmacogenomics of antidepressant drugs. Phar- Kaymaz, N., van Os, J., Loonen, A.J., Nolen, W.A., 2008. Evidence that patients with
macology & Therapeutics 124, 57–73. single versus recurrent depressive episodes are differentially sensitive to treat-
Humphries, M.D., Prescott, T.J., 2010. The ventral basal ganglia, a selection mecha- ment discontinuation: a meta-analysis of placebo-controlled randomized trials.
nism at the crossroads of space, strategy, and reward. Progress in Neurobiology The Journal of Clinical Psychiatry 69, 1423–1436.
90, 385–417. Kee, N., Teixeira, C.M., Wang, A.H., Frankland, P.W., 2007. Preferential incorporation
Huuhka, K., Anttila, S., Huuhka, M., Hietala, J., Huhtala, H., Mononen, N., of adult-generated granule cells into spatial memory networks in the dentate
Lehtimaki, T., Leinonen, E., 2008. Dopamine 2 receptor C957T and catechol- gyrus. Nature Neuroscience 10, 355–362.
o-methyltransferase Val158Met polymorphisms are associated with treat- Keedwell, P.A., Andrew, C., Williams, S.C., Brammer, M.J., Phillips, M.L., 2005. The
ment response in electroconvulsive therapy. Neuroscience Letters 448, neural correlates of anhedonia in major depressive disorder. Biological Psychi-
79–83. atry 58, 843–853.
Ibarguen-Vargas, Y., Surget, A., Vourc’h, P., Leman, S., Andres, C., Gardier, A., Belzung, Kendler, K.S., Thornton, L.M., Gardner, C.O., 2000. Stressful life events and previous
C., 2009. Deficit in BDNF level does not increase vulnerability to stress but episodes in the etiolgy of major depression in women: an evaluation of the
dampens antidepressant-like effects in the unpredictable chronic mild stress. “kindling” hypothesis. The American Journal of Psychiatry 157, 1243–1251.
Behavioural Brain Research 202, 245–251. Kendler, K.S., Thornton, L.M., Gardner, C.O., 2001. Genetic risk, number of previous
Ikemoto, S., Panksepp, J., 1999. The role of nucleus accumbens dopamine in depressive episodes, and stressful life events in predicting the onset of major
motivated behavior: a unifying interpretation with special reference to reward- depression. The American Journal of Psychiatry 158, 582–586.
seeking. Brain Research. Brain Research Reviews 31, 6–41. Kendler, K.S., Gardner, C.O., Prescott, C.A., 2002. Toward a comprehensive devel-
Iosifescu, D.V., Clementi-Craven, N., Fraguas, R., Papakostas, G.I., Petersen, T., Alpert, opmental model for major depression in women. The American Journal of
J.E., Nierenberg, A.A., Fava, M., 2005. Cardiovascular risk factors may moderate Psychiatry 159, 1133–1145.
pharmacological treatment effects in major depressive disorder. Psychosomatic Kendler, K.S., Gardner, C.O., Prescott, C.A., 2006. Toward a comprehensive develop-
Medicine 67, 703–706. mental model for major depression in men. The American Journal of Psychiatry
Ising, M., Horstmann, S., Kloiber, S., Lucae, S., Binder, E.B., Kern, N., et al., 2007. Com- 163, 115–124.
bined dexamethasone/corticotropin releasing hormone test predicts treatment Kendler, K.S., Gardner, C.O., 2011. A longitudinal etiologic model for symptoms of
response in major depression – a potential biomarker? Biological Psychiatry 62, anxiety and depression in women. Psychological Medicine 41, 2035–2045.
47–54. Kennedy, S.H., Konarski, J.Z., Segal, Z.V., Lau, M.A., Bieling, P.J., McIntyre, R.S., May-
Isingrini, E., Camus, V., Le Guisquet, A.M., Pingaud, M., Devers, S., Belzung, C., 2010. berg, H.S., 2007. Differences in brain glucose metabolism between responders
Association between repeated unpredictable chronic mild stress (UCMS) proce- to CBT and venlafaxine in a 16-week randomized controlled trial. The American
dures with a high fat diet: a model of fluoxetine resistance in mice. PLoS One 5, Journal of Psychiatry 164, 778–788.
e10404. Kennedy, S.H., Young, A.H., Blier, P., 2011. Strategies to achieve clinical effectiveness:
Jacobson, L., Sapolsky, R., 1991. The role of the hippocampus in feedback regula- refining existing therapies and pursuing emerging targets. Journal of Affective
tion of the hypothalamic-pituitary-adrenocortical axis. Endocrine Reviews 12, Disorders 132 (Suppl. 1), S21–S28.
118–134. Kessler, R.C., Price, R.H., Wortman, C.B., 1985. Social factors in psychopathology:
Jalabert, M., Aston-Jones, G., Herzog, E., Manzoni, O., Georges, F., 2009. Role of the bed stress, social support, and coping processes. Annual Review of Psychology 36,
nucleus of the stria terminalis in the control of ventral tegmental area dopamine 531–572.
neurons. Progress in Neuro-psychopharmacology & Biological Psychiatry 33, Khundakar, A., Morris, C., Oakley, A., Thomas, A.J., 2011. Morphometric analysis of
1336–1346. neuronal and glial cell pathology in the caudate nucleus in late-life depression.
Jankord, R., Herman, J.P., 2008. Limbic regulation of hypothalamo-pituitary- The American Journal of Geriatric Psychiatry: Official Journal of the American
adrenocortical function during acute and chronic stress. Annals of the New York Association for Geriatric Psychiatry 19, 132–141.
Academy of Sciences 1148, 64–73. Kiecolt-Glaser, J.K., Preacher, K.J., MacCallum, R.C., Atkinson, C., Malarkey, W.B.,
Jay, T.M., Glowinski, J., Thierry, A.M., 1989. Selectivity of the hippocampal projection Glaser, R., 2003. Chronic stress and age-related increases in the proinflammatory
to the prelimbic area of the prefrontal cortex in the rat. Brain Research 505, cytokine IL-6. Proceedings of the National Academy of Sciences of the United
337–340. States of America 100, 9090–9095.
Jayatissa, M.N., Bisgaard, C.F., West, M.J., Wiborg, O., 2008. The number of granule Kim, M.J., Hamilton, J.P., Gotlib, I.H., 2008. Reduced caudate gray matter volume in
cells in rat hippocampus is reduced after chronic mild stress and re-established women with major depressive disorder. Psychiatry Research 164, 114–122.
after chronic escitalopram treatment. Neuropharmacology 54, 530–541. Kimura, M., Müller-Preuss, P., Lu, A., Wiesner, E., Flachskamm, C., Wurst, W.,
Jayatissa, M.N., Henningsen, K., Nikolajsen, G., West, M.J., Wiborg, O., 2010. A reduced Holsboer, F., Deussing, J.M., 2010. Conditional corticotropin-releasing hormone
number of hippocampal granule cells does not associate with an anhedonia-like overexpression in the mouse forebrain enhances rapid eye movement sleep.
phenotype in a rat chronic mild stress model of depression. Stress 13, 95–105. Molecular Psychiatry 15, 154–165.
P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371 2365

Kinsbourne, M., 1978. The biological determinants of functional bisymmetry and Lisiecka, D.M., Carballedo, A., Fagan, A.J., Connolly, G., Meaney, J., Frodl, T., 2011.
asymmetry. In: Kinsbourne, M. (Ed.), Asymmetrical Functions of the Brain. Altered inhibition of negative emotions in subjects at family risk of major depres-
Cambridge University Press, New York, pp. 553–565. sive disorder. Journal of Psychiatric Research 46, 181–188.
Kirby, E.D., Friedman, A.R., Covarrubias, D., Ying, C., Sun, W.G., Goosens, K.A., Sapol- Lisman, J.E., Grace, A.A., 2005. The hippocampal-VTA loop: controlling the entry of
sky, R.M., Kaufer, D., 2012. Basolateral amygdala regulation of adult hippocampal information into long-term memory. Neuron 46, 703–713.
neurogenesis and fear-related activation of newborn neurons. Molecular Psychi- Liu, R.J., Aghajanian, G.K., 2008. Stress blunts serotonin- and hypocretin-evoked
atry, http://dx.doi.org/10.1038/mp.2011.71 [Epub ahead of print]. EPSCs in prefrontal cortex: role of corticosterone-mediated apical dendritic atro-
Kito, S., Fujita, K., Koga, Y., 2008. Regional cerebral blood flow changes after phy. Proceedings of the National Academy of Sciences of the United States of
low frequency transcranial magnetic stimulation of the right dorsolateral America 105, 359–364.
prefrontal cortex in treatment-resistant depression. Neuropsychobiology 58, Lorenzetti, V., Allen, N.B., Fornito, A., Yücel, M., 2009. Structural brain abnormalities
29–36. in major depressive disorder: a selective review of recent MRI studies. Journal
Knapp, R.J., Goldenberg, R., Shuck, C., Cecil, A., Watkins, J., Miller, C., Crites, G., of Affective Disorders 117, 1–17.
Malatynska, E., 2002. Antidepressant activity of memory-enhancing drugs in Lowry, C.A., Hale, M.W., Evans, A.K., Heerkens, J., Staub, D.R., Gasser, P.J., Shekhar, A.,
the reduction of submissive behavior model. European Journal of Pharmacology 2008. Serotonergic systems, anxiety, and affective disorder: focus on the dor-
440, 27–35. somedial part of the dorsal raphe nucleus. Annals of the New York Academy of
Komossa, K., Depping, A.M., Gaudchau, A., Kissling, W., Leucht, S., 2010. Second- Sciences 1148, 86–94.
generation antipsychotics for major depressive disorder and dysthymia. Lucassen, P.J., Stumpel, M.W., Wang, Q., Aronica, E., 2010. Decreased numbers of
Cochrane Database of Systematic Reviews 12, CD008121. progenitor cells but no response to antidepressant drugs in the hippocampus of
Koo, J.W., Duman, R.S., 2009. Evidence for IL-1 receptor blockade as a therapeu- elderly depressed patients. Neuropharmacology 58, 940–949.
tic strategy for the treatment of depression. Current Opinion in Investigational Lucki, I., O’Leary, O.F., 2004. Distinguishing roles for norepinephrine and serotonin in
Drugs 10, 664–671. the behavioral effects of antidepressant drugs. The Journal of Clinical Psychiatry
Kringelbach, M.L., 2005. The human orbitofrontal cortex: linking reward to hedonic 65 (Suppl. 4), 11–24.
experience. Nature Reviews. Neuroscience 6, 691–702. Lyons, D.M., Buckmaster, P.S., Lee, A.G., Wu, C., Mitra, R., Duffey, L.M., Buckmaster,
Kringelbach, M.L., Rolls, E.T., 2004. The functional neuroanatomy of the human C.L., Her, S., Patel, P.D., Schatzberg, A.F., 2010. Stress coping stimulates hip-
orbitofrontal cortex: evidence from neuroimaging and neuropsychology. pocampal neurogenesis in adult monkeys. Proceedings of the National Academy
Progress in Neurobiology 72, 341–372. of Sciences of the United States of America 107, 14823–14827.
Krishnan, K.R., Hays, J.C., Blazer, D.G., 1997. MRI-defined vascular depression. Amer- MacQueen, G., Frodl, T., 2010. The hippocapmus in major depression: evidence for
ican Journal of Psychiatry 154, 497–501. the convergence of the bench and bedside in psychiatric research? Molecular
Krishnan, V., Nestler, E.J., 2008. The molecular neurobiology of depression. Nature Psychiatry 16, 252–264.
455, 894–902. MacQueen, G.M., Yucel, K., Taylor, V.H., Macdonald, K., Joffe, R., 2008. Posterior hip-
Krishnan, V., Han, M.H., Graham, D.L., Berton, O., Renthal, W., Russo, S.J., Laplant, Q., pocampal volumes are associated with remission rates in patients with major
Graham, A., Lutter, M., Lagace, D.C., Ghose, S., Reister, R., Tannous, P., Green, T.A., depressive disorder. Biological Psychiatry 15, 880–883.
Neve, R.L., Chakravarty, S., Kumar, A., Eisch, A.J., Self, D.W., Lee, F.S., Tamminga, Maeng, S., Zarate, C.A., Du, J., Schloesser, R.J., McCammon, J., Chen, G., Manji, H.K.,
C.A., Cooper, D.C., Gershenfeld, H.K., Nestler, E.J., 2007. Molecular adaptations 2008. Cellular mechanisms underlying the antidepressant effects of ketamine:
underlying susceptibility and resistance to social defeat in brain reward regions. role of alpha-amino-3-hydroxy-5-methylisoxazole-4-propionic acid receptors.
Cell 131, 391–404. Biological Psychiatry 63, 349–352.
Kronenberg, G., Reuter, K., Steiner, B., Brandt, M.D., Jessberger, S., Yamaguchi, M., Magarinos, A.M., McEwen, B.S., Flugge, G., Fuchs, E., 1996. Chronic psychosocial
Kempermann, G., 2003. Subpopulations of proliferating cells of the adult hip- stress causes apical dendritic atrophy of hippocampal CA3 pyramidal neurons
pocampus respond differently to physiologic neurogenic stimuli. The Journal of in subordinate tree shrews. The Journal of Neuroscience: the Official Journal of
Comparative Neurology 467, 455–463. the Society for Neuroscience 16, 3534–3540.
Kulkarni, V.A., Jha, S., Vaidya, V.A., 2002. Depletion of norepinephrine decreases the Maier, S.F., Seligman, M.E.P., 1976. Learned helplessness:theory and evidence. Jour-
proliferation, but does not influence the survival and differentiation, of granule nal of Experimental Psychology: General 105, 3–46.
cell progenitors in the adult rat hippocampus. The European Journal of Neuro- Maier, S.F., Watkins, L.R., 2010. Role of the medial prefrontal cortex in coping and
science 16, 2008–2012. resilience. Brain Research 1355, 52–60.
Lacefield, C.O., Itskov, V., Reardon, T., Hen, R., Gordon, J.A., 2012. Effects of adult- Malberg, J.E., Duman, R.S., 2003. Cell proliferation in adult hippocampus is decreased
generated granule cells on coordinated network activity in the dentate gyrus. by inescapable stress: reversal by fluoxetine treatment. Neuropsychopharma-
Hippocampus 22, 106–116. cology 28, 1562–1571.
Lakdawalla, Z., Hankin, B.L., 2008. Personality as a prospective vulnerability to Malberg, J.E., Eisch, A.J., Nestler, E.J., Duman, R.S., 2000. Chronic antidepres-
dysphoric symptoms among college students: proposed mechanisms. Journal sant treatment increases neurogenesis in adult rat hippocampus. The Journal
of Psychopathology and Behavioral Assessment 30, 121–131. of Neuroscience: the Official Journal of the Society for Neuroscience 20,
Lee, J.S., Jang, D.J., Lee, N., Ko, H.G., Kim, H., Kim, Y.S., Kim, B., Son, J., Kim, S.H., 9104–9110.
Chung, H., Lee, M.Y., Kim, W.R., Sun, W., Zhuo, M., Abel, T., Kaang, B.K., Son, H., Malykhin, N.V., Carter, R., Seres, P., Coupland, N.J., 2010. Structural changes in the
2009. Induction of neuronal vascular endothelial growth factor expression by hippocampus in major depressive disorder: contributions of disease and treat-
cAMP in the dentate gyrus of the hippocampus is required for antidepressant- ment. Journal of Psychiatry & Neuroscience 35, 337–343.
like behaviors. The Journal of Neuroscience: the Official Journal of the Society Manev, R., Uz, T., Manev, H., 2001. Fluoxetine increases the content of neurotrophic
for Neuroscience 29, 8493–8505. protein S100beta in the rat hippocampus. European Journal of Pharmacology
Lemogne, C., Delaveau, P., Freton, M., Guionnet, S., Fossati, P., 2012. Medial pre- 420, R1–R2.
frontal cortex and the self in major depression. Journal of Affective Disorders Maninger, N., Wolkowitz, O.M., Reus, V.I., Epel, E.S., Mellon, S.H., 2009. Neurobiolog-
136, e1–e11. ical and neuropsychiatric effects of dehydroepiandrosterone (DHEA) and DHEA
Lemos, J.C., Wanat, M.J., Smith, J.S., Reyes, B.A., Hollon, N.G., Van Bockstaele, E.J., sulfate (DHEAS). Frontiers in Neuroendocrinology 30, 65–91.
Chavkin, C., Phillips, P.E., 2012. Severe stress switches CRF action in the nucleus Martinowich, K., Manji, H., Lu, B., 2007. New insights into BDNF function in depres-
accumbens from appetitive to aversive. Nature 7420, 402–406. sion and anxiety. Nature Neuroscience 10, 1089–1093.
Lesser, I.M., Boone, K.B., Mehringer, C.M., Wohl, M.A., Miller, B.L., Berman, N.G., Mateus-Pinheiro, A., Pinto, L., Bessa, J.M., Morais, M., Monteiro, S., Sousa, N.N., 2011.
1996. Cognitiong and white matter hyperintensities in older depressed patients. Long-lasting mood and cognitive recovery from depression is neurogenesis-
American Journal of Psychiatry 153, 1280–1287. dependent. Society for Neuroscience Abstracts 683, 19/GG10.
Levesque, M.L., Beauregard, M., Ottenhof, K.W., Fortier, E., Tremblay, R.E., Brend- Mathews, A., MacLeod, C., 2005. Cognitive vulnerability to emotional disorders.
gen, M., Perusse, D., Dionne, G., Robaey, P., Vitaro, F., Boivin, M., Booij, L., Annual Review of Clinical Psychology 1, 167–195.
2011. Altered patterns of brain activity during transient sadness in chil- Matsumoto, M., Hikosaka, O., 2009. Representation of negative motivational value
dren at familial risk for major depression. Journal of Affective Disorders 135, in the primate lateral habenula. Nature Neuroscience 12, 77–84.
410–413. Mayberg, H.S., 2009. Targeted electrode-based modulation of neural circuits for
Li, B., Piriz, J., Mirrione, M., Chung, C., Proulx, C.D., Schulz, D., Henn, F., Malinow, R., depression. The Journal of Clinical Investigation 119, 717–725.
2011a. Synaptic potentiation onto habenula neurons in the learned helplessness Mayberg, H.S., Brannan, S.K., Mahurin, R.K., Jerabek, P.A., Brickman, J.S., Tekell, J.L.,
model of depression. Nature 470, 535–539. Silva, J.A., McGinnis, S., Glass, T.G., Martin, C.C., Fox, P.T., 1997. Cingulate func-
Li, N., Lee, B., Liu, R.J., Banasr, M., Dwyer, J.M., Iwata, M., Li, X.Y., Aghaja- tion in depression: a potential predictor of treatment response. Neuroreport 8,
nian, G., Duman, R.S., 2010. mTOR-dependent synapse formation underl- 1057–1061.
ies the rapid antidepressant effects of NMDA antagonists. Science 329, Mayberg, H.S., Liotti, M., Brannan, S.K., McGinnis, S., Mahurin, R.K., Jerabek, P.A., Silva,
959–964. J.A., Tekell, J.L., Martin, C.C., Lancaster, J.L., Fox, P.T., 1999. Reciprocal limbic-
Li, N., Liu, R.J., Dwyer, J.M., Banasr, M., Lee, B., Son, H., Li, X.Y., Aghajanian, G., cortical function and negative mood: converging PET findings in depression and
Duman, R.S., 2011b. Glutamate N-methyl-D-aspartate receptor antagonists normal sadness. The American Journal of Psychiatry 156, 675–682.
rapidly reverse behavioral and synaptic deficits caused by chronic stress expo- Mayberg, H.S., Silva, J.A., Brannan, S.K., Tekell, J.L., Mahurin, R.K., McGinnis, S.,
sure. Biological Psychiatry 69, 754–761. Jerabek, P.A., 2002. The functional neuroanatomy of the placebo effect. The
Li, Y., Luikart, B.W., Birnbaum, S., Chen, J., Kwon, C.H., Kernie, S.G., Bassel-Duby, American Journal of Psychiatry 159, 728–737.
R., Parada, L.F., 2008. TrkB regulates hippocampal neurogenesis and governs McCabe, C., Cowen, P.J., Harmer, C.J., 2009. Neural representation of reward in recov-
sensitivity to antidepressive treatment. Neuron 59, 399–412. ered depressed patients. Psychopharmacology 205, 667–677.
Licinio, J., Wong, M.L., 2011. Pharmacogenomics of antidepressant treatment effects. McCormick, L.M., Boles Ponto, L.L., Pierson, R.K., Johnson, H.J., Magnotta, V., Brumm,
Dialogues in Clinical Neuroscience 13, 63–71. M.C., 2007. Metabolic correlates of antidepressant and antipsychotic response in
2366 P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371

patients with psychotic depression undergoing electroconvulsive therapy. The Monroe, S.M., Harkness, K.L., 2005. Life stress, the “kindling” hypothesis, and the
Journal of ECT 23, 265–273. recurrence of depression: considerations from a life stress perspective. Psycho-
McEwen, B.S., 2010. Stress, sex, and neural adaptation to a changing environment: logical Review 112, 417–445.
mechanisms of neuronal remodeling. Annals of the New York Academy of Sci- Monroe, S.M., Simons, A.D., 1991. Diathesis-stress theories in the context of life-
ences 1204 (Suppl.), E38–E59. stress research: implications for the depressive disorders. Psychological Bulletin
McEwen, B.S., Gianaros, P.J., 2010. Central role of the brain in stress and adapta- 110, 406–425.
tion: links to socioeconomic status, health, and disease. Annals of the New York Monroe, S.M., Slavich, G.M., Torres, L.D., Gotlib, I.H., 2007. Severe life events predict
Academy of Sciences 1186, 192–222. specific patterns of change in cognitive biases in major depression. Psychological
McEwen, B.S., Magarinos, A.M., Reagan, L.P., 2002. Structural plasticity and tianep- Medicine 37, 863–871.
tine: cellular and molecular targets. European Psychiatry: the Journal of the Monteggia, L.M., Barrot, M., Powell, C.M., Berton, O., Galanis, V., Gemelli, T., Meuth,
Association of European Psychiatrists 17 (Suppl. 3), 318–330. S., Nagy, A., Greene, R.W., Nestler, E.J., 2004. Essential role of brain-derived neu-
McFarland, B.R., Klein, D.N., 2008. Emotional reactivity in depression: diminished rotrophic factor in adult hippocampal function. Proceedings of the National
responsiveness to anticipated reward but not to anticipated punishment or to Academy of Sciences of the United States of America 101, 10827–10832.
nonreward or avoidance. Depression and Anxiety 26, 117–122. Montgomery, S.A., Baldwin, D.S., Blier, P., Fineberg, N.A., Kasper, S., Lader, M., Lam,
McKinnon, M.C., Yucel, K., Nazarov, A., MacQueen, G.M., 2009. A meta-analysis R.W., Lépine, J.P., Möller, H.J., Nutt, D.J., Rouillon, F., Schatzberg, A.F., Thase, M.E.,
examining clinical predictors of hippocampal volume in patients with major 2007. Which antidepressants have demonstrated superior efficacy? A review of
depressive disorder. Journal of Psychiatry & Neuroscience 34, 41–54. the evidence. International Clinical Psychopharmacology 22, 323–329.
McLaughlin, R.J., Gobbi, G., 2012. Cannabinoids and emotionality: a neuroanatomical Morris, J.S., Smith, K.A., Cowen, P.J., Friston, K.J., Dolan, R.J., 1999. Covariation of
perspective. Neuroscience 204, 134–144. activity in habenula and dorsal raphé nuclei following tryptophan depletion.
Mehta, M., Gore-Langton, E., Golembo, N., Colvert, E., Williams, S., Sonuga-Barke, NeuroImage 10, 163–172.
E., 2010. Hyporesponsive reward anticipation in the basal ganglia following Morris, M.C., Ciesla, J.A., Garber, J., 2010. A prospective study of stress autonomy
severe institutional deprivation early in life. Journal of Cognitive Neuroscience versus stress sensitization in adolescents at varied risk for depression. Journal
22, 2316–2325. of Abnormal Psychology 119, 341–354.
Merkl, A., Heuser, I., Bajbouj, M., 2009. Antidepressant electroconvulsive therapy: Morrison, S.E., Salzman, C.D., 2010. Re-valuing the amygdala. Current Opinion in
mechanism of action, recent advances and limitations. Experimental Neurology Neurobiology 20, 221–230.
219, 20–26. Morrison, S.E., Salzman, C.D., 2011. Representations of appetitive and aversive infor-
Meseguer, V., Romero, M.J., Barrós-Loscertales, A., Belloch, V., Bosch-Morell, F., rmation in the primate orbitofrontal cortex. Annals of the New York Academy
Romero, J., Avila, C., 2007. Mapping the apetitive and aversive systems with emo- of Sciences 1239, 59–70.
tional pictures using a block-design fMRI procedure. Psicothema 19, 483–488. Moses-Kolko, E.L., Fraser, D., Wisner, K.L., James, J.A., Saul, A.T., Fiez, J.A., Phillips,
Meyer, J.H., 2012. Neuroimaging markers of cellular function in major depressive M.L., 2011. Rapid habituation of ventral striatal response to reward receipt in
disorder: implications for therapeutics, personalized medicine and prevention. postpartum depression. Biological Psychiatry 70, 395–399.
Clinical Pharmacology and Therupetics 91, 201–214. Munafò, M.R., Brown, S.M., Hariri, A.R., 2008. Serotonin transporter (5-HTTLPR)
Meyer, J.H., McMain, S., Kennedy, S.H., Korman, L., Brown, G.M., DaSilva, J.N., Wil- genotype and amygdala activation: a meta-analysis. Biological Psychiatry 63,
son, A.A., Blak, T., Eynan-Harvey, R., Goulding, V.S., Houle, S., Links, P., 2003. 852–857.
Dysfunctional attitudes and 5-HT2 receptors during depression and self harm. Murray, E.A., Wise, S.P., Drevets, W.C., 2011. Localization of dysfunction in major
The American Journal of Psychiatry 160, 90–99. depressive disorder: prefrontal cortex and amygdala. Biological Psychiatry 69,
Meyer, J.H., Ginovart, N., Boovariwala, A., Sagrati, S., Hussey, D., Garcia, A., Young, T., 43–54.
Praschak-Rieder, N., Wilson, A.A., Houle, S., 2006. Elevated monoamine oxidase Muscat, R., Sampson, D., Willner, P., 1990. Dopaminergic mechanism of imipramine
a levels in the brain: an explanation for the monoamine imbalance of major action in an animal model of depression. Biological Psychiatry 28, 223–229.
depression. Archives of General Psychiatry 63, 209–1216. Muscat, R., Papp, M., Willner, P., 1992. Reversal of stress-induced anhedonia by the
Meyer, J.H., Wilson, A.A., Sagrati, S., Miler, L., Rusjan, P., Bloomfield, P.M., Clark, atypical antidepressants, fluoxetine and maprotiline. Psychopharmacology 109,
M., Sacher, J., Voineskos, A.N., Houle, S., 2009. Brain monoamine oxidase A 433–438.
binding in major depressive disorder: relationship to selective serotonin Nandrino, J.L., Dodin, V., Martin, P., Henniaux, M., 2004. Emotional information
reuptake inhibitor treatment, recovery, and recurrence. Archives of General processing in first and recurrent major depressive episodes. Journal of Psychi-
Psychiatry 66, 1304–1312. atric Research 38, 475–484.
Miller, B.H., Schultz, L.E., Gulati, A., Cameron, M.D., Pletcher, M.T., 2008. Genetic Naudin, M., El-Hage, W., Gomes, M., Gaillard, P., Belzung, C., Atanasova, B., 2012.
regulation of behavioral and neuronal responses to fluoxetine. Neuropsycho- State and trait olfactory markers of major depression. PLoS One 7 (10), e46938.
pharmacology 33, 1312–1322. Nery, F.G., Stanley, J.A., Chen, H.H., Hatch, J.P., Nicoletti, M.A., Monkul, E.S., Matsuo,
Miller, H.L., Delgado, P.L., Salomon, R.M., Berman, R., Krystal, J.H., Heninger, G.R., K., Caetano, S.C., Peluso, M.A., Najt, P., Soares, J.C., 2009. Normal metabolite lev-
Charney, D.S., 1996. Clinical and biochemical effects of catecholamine deple- els in the left dorsolateral prefrontal cortex of unmedicated major depressive
tion on antidepressant-induced remission of depression. Archives of General disorder patients: a single voxel (1)H spectroscopy study. Psychiatry Research
Psychiatry 53, 117–128. 174, 177–183.
Mineur, Y.S., Picciotto, M.R., 2010. Nicotine receptors and depression: revisiting Neumeister, A., Nugent, A.C., Waldeck, T., Geraci, M., Schwarz, M., Bonne, O., Bain,
and revising the cholinergic hypothesis. Trends in Pharmacological Sciences 31, E.E., Luckenbaugh, D.A., Herscovitch, P., Charney, D.S., Drevets, W.C., 2004. Neu-
580–586. ral and behavioral responses to tryptophan depletion in unmedicated patients
Mineur, Y.S., Belzung, C., Crusio, W.E., 2007. Functional implications of decreases in with remitted major depressive disorder and controls. Archives of General Psy-
neurogenesis following chronic mild stress in mice. Neuroscience 150, 251–259. chiatry 61, 765–773.
Miracle, A.D., Brace, M.F., Huyck, K.D., Singler, S.A., Wellman, C.L., 2006. Chronic Neumeister, A., Wood, S., Bonne, O., Nugent, A.C., Luckenbaugh, D.A., Young, T.,
stress impairs recall of extinction of conditioned fear. Neurobiology of Learning Bain, E.E., Charney, D.S., Drevets, W.C., 2005. Reduced hippocampal volume in
and Memory 85, 213–218. unmedicated, remitted patients with major depression versus control subjects.
Mirescu, C., Gould, E., 2006. Stress and adult neurogenesis. Hippocampus 16, Biological Psychiatry 57, 935–937.
233–238. Newcomer, J.W., Selke, G., Melson, A.K., Hershey, T., Craft, S., Richards, K.,
Miskowiak, K.W., Vinberg, M., Harmer, C.J., Ehrenreich, H., Kessing, L.V., 2012. Alderson, A.L., 1999. Decreased memory performance in healthy humans
Erythropoietin: a candidate treatment for mood symptoms and memory dys- induced by stress-level cortisol treatment. Archives of General Psychiatry 56,
function in depression. Psychopharmacology 219, 687–698. 527–533.
Mogenson, G.J., Swanson, L.W., Yim, C.Y., 1980. From motivation to action: func- Newton, S.S., Thome, J., Wallace, T.L., Shirayama, Y., Schlesinger, L., Sakai, N.,
tional interface between the limbic system and the motor system. Progress in Chen, J., Neve, R., Nestler, E.J., Duman, R.S., 2002. Inhibition of cAMP response
Neurobiology 14, 69–97. element-binding protein or dynorphin in the nucleus accumbens produces an
Moghaddam, B., Adams, B., Verma, A., Daly, D., 1997. Activation of glutamatergic neu- antidepressant-like effect. The Journal of Neuroscience: the Official Journal of
rotransmission by ketamine: a novel step in the pathway from NMDA receptor the Society for Neuroscience 22, 10883–10890.
blockade to dopaminergic and cognitive disruptions associated with the pre- Ng, F., Berk, M., Dean, O., Bush, A.I., 2008. Oxidative stress in psychiatric disor-
frontal cortex. The Journal of Neuroscience: the Official Journal of the Society ders: evidence base and therapeutic implications. The International Journal of
for Neuroscience 17, 2921–2927. Neuropsychopharmacology 11, 851–876.
Möller, H.J., 2005. Antipsychotic and antidepressive effects of second generation Nibuya, M., Morinobu, S., Duman, R.S., 1995. Regulation of BDNF and TrkB mRNA in
antipsychotics: two different pharmacological mechanisms? European Archives rat brain by chronic electroconvulsive seizure and antidepressant drug treat-
of Psychiatry and Clinical Neuroscience 255, 190–201. ments. The Journal of Neuroscience: the Official Journal of the Society for
Molteni, R., Calabrese, F., Bedogni, F., Tongiorgi, E., Fumagalli, F., Racagni, G., Riva, Neuroscience 15, 7539–7547.
M.A., 2006. Chronic treatment with fluoxetine up-regulates cellular BDNF mRNA Nibuya, M., Nestler, E.J., Duman, R.S., 1996. Chronic antidepressant administration
expression in rat dopaminergic regions. The International Journal of Neuropsy- increases the expression of cAMP response element binding protein (CREB) in
chopharmacology 9, 307–317. rat hippocampus. The Journal of Neuroscience: the Official Journal of the Society
Monkul, E.S., Silva, L.A., Narayana, S., Peluso, M.A., Zamarripa, F., Nery, F.G., Najt, for Neuroscience 16, 2365–2372.
P., Li, J., Lancaster, J.L., Fox, P.T., Lafer, B., Soares, J.C., 2011. Abnormal res- Nolen-Hoeksema, S.N., Wisco, B.E., Lyubomirsky, S., 2008. Rethinking rumination.
ting state corticolimbic blood flow in depressed unmedicated patients with Perspectives on Psychological Science 3, 400–423.
major depression: a (15) O-H(2) O PET study. Human Brain Mapping 33 (2), Ochsner, K.N., Gross, J.J., 2005. The cognitive control of emotion. Trends in Cognitive
272–279. Sciences 9, 242–249.
P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371 2367

Ochsner, K.N., Ray, R.D., Cooper, J.C., Robertson, E.R., Chopra, S., Gabrieli, J.D., Gross, Phelps, E.A., LeDoux, J.E., 2005. Contributions of the amygdala to emotion processing:
J.J., 2004. For better or for worse: neural systems supporting the cognitive down- from animal models to human behavior. Neuron 48, 175–187.
and up-regulation of negative emotion. NeuroImage 23, 483–499. Phillips, M.L., Drevets, W.C., Rauch, S.L., Lane, R., 2003a. Neurobiology of emotion per-
Okugawa, G., Omori, K., Suzukawa, J., Fujiseki, Y., Kinoshita, T., Inagaki, C., 1999. ception I: the neural basis of normal emotion perception. Biological Psychiatry
Long-term treatment with antidepressants increases glucocorticoid receptor 54, 504–514.
binding and gene expression in cultured rat hippocampal neurones. Journal of Phillips, M.L., Drevets, W.C., Rauch, S.L., Lane, R., 2003b. Neuorbiology of emotion
Neuroendocrinology 11, 887–895. perception II: implications for major psychiatric disorders. Biological Psychiatry
Olff, M., 1999. Stress, depression and immunity: the role of defense and coping styles. 54, 515–528.
Psychiatry Research 85, 7–15. Pimontel, M.A., Culang-Reinlieb, M.E., Morimoto, S.S., Sneed, J.R., 2012. Executive
Oomen, C.A., Mayer, J.L., de Kloet, E.R., Joels, M., Lucassen, P.J., 2007. Brief treatment dysfunction and treatment response in late-life depression. International Jour-
with the glucocorticoid receptor antagonist mifepristone normalizes the reduc- nal of Geriatric Psychiatry 27, 893–899.
tion in neurogenesis after chronic stress. The European Journal of Neuroscience Pitkänen, A., Pikkarainen, M., Nurminen, N., Ylinen, A., 2000. Reciprocal connections
26, 3395–3401. between the amygdala and the hippocampal formation, perirhinal cortex, and
O’Reardon, J.P., Solvason, H.B., Janicak, P.G., Sampson, S., Isenberg, K.E., Nahas, Z., postrhinal cortex in rat. A review. Annals of the New York Academy of Sciences
McDonald, W.M., Avery, D., Fitzgerald, P.B., Loo, C., Demitrack, M.A., George, M.S., 911, 369–391.
Sackeim, H.A., 2007. Efficacy and safety of transcranial magnetic stimulation in Pittenger, C., Duman, R.S., 2008. Stress, depression, and neuroplasticity: a conver-
the acute treatment of major depression: a multisite randomized controlled gence of mechanisms. Neuropsychopharmacology 33, 88–109.
trial. Biological Psychiatry 62, 1208–1216. Pizzagalli, D.A., 2011. Frontocingulate dysfunction in depression: toward biomarkers
Ossewaarde, L., Verkes, R.J., Hermans, E.J., Kooijman, S.C., Urner, M., Tendolkar, I., van of treatment response. Neuropsychopharmacology Review 36, 183–206.
Wingen, G.A., Fernández, G., 2011. Two-week administration of the combined Pizzagalli, D., Iosifescu, D., Hallett, L.A., Ratner, K.G., Fava, M., 2008. Reduced hedonic
serotonin-noradrenaline reuptake inhibitor duloxetine augments functioning of capacity in major depressive disorder: evidence from a probabilistic reward task.
mesolimbic incentive processing circuits. Biological Psychiatry 70, 568–574. Journal of Psychiatric Research 43, 76–87.
Palomero-Gallagher, N., Vogt, B.A., Schliegher, A., Mayberg, H.S., Zilles, K., 2009. Pizzagalli, D.A., Holmes, A.J., Dillon, D.G., Goetz, E.L., Birk, J.L., Bogdan, R., Dougherty,
Receptor architecture of human cingulate cortex: evaluatin of the four-region D.D., Iosifescu, D.V., Rauch, S.L., Fava, M., 2009. Reduced caudate and nucleus
neurobiological model. Human Brain Mapping 30, 2336–2355. accumbens response to rewards in unmedicated individuals with major depres-
Pampallona, S., Bollini, P., Tibaldi, G., Kupelnick, B., Munizza, C., 2002. Patient adher- sive disorder. The American Journal of Psychiatry 166, 702–710.
ence in the treatment of depression. The British Journal of Psychiatry: the Journal Pollack, M.H., Otto, M.W., Roy-Byrne, P.P., Coplan, J.D., Rothbaum, B.O., Simon, N.M.,
of Mental Science 180, 104–109. Gorman, J.M., 2008. Novel treatment approaches for refractory anxiety disor-
Papakostas, G.I., Petersen, T., Sonawalla, S.B., Merens, W., Iosifescu, D.V., Alpert, ders. Depression and Anxiety 25 (6), 467–476.
J.E., Fava, M., Nierenberg, A.A., 2003. Serum cholesterol in treatment-resistant Polter, A.M., Li, X., 2010. 5HT1A receptor-regulated signal transduction pathways in
depression. Neuropsychobiology 47, 146–151. brain. Cellular Signalling 22, 1406–1412.
Papakostas, G.I., Ongür, D., Iosifescu, D.V., Mischoulon, D., Fava, M., 2004. Cholesterol Post, R.M., 1992. Transduction of psychosocial stress into the neurobiology of recur-
in mood and anxiety disorders: review of the literature and new hypotheses. rent affective disorder. The American Journal of Psychiatry 149, 999–1010.
European Neuropsychopharmacology 14, 135–142. Preiss, M., Kucerova, H., Lukavsky, J., Stepankova, H., Sos, P., Kawaciukova, R., 2009.
Park, S.K., Nguyen, M.D., Fischer, A., Luke, M.P., Affar el, B., Dieffenbach, P.B., Tseng, Cognitive deficits in the euthymic phase of unipolar depression. Psychiatry
H.C., Shi, Y., Tsai, L.H., 2005. Par-4 links dopamine signaling and depression. Cell Research 169, 235–239.
122, 275–287. Preskorn, S.H., 1997. Clinically relevant pharmacology of selective serotonin reup-
Paslakis, G., Heuser, I., Schweiger, U., Deuschle, M., 2010. A single DEX/CRH test in take inhibitors. An overview with emphasis on pharmacokinetics and effects on
male drug-free depressed patients is associated with the clinical response to oxidative drug metabolism. Clinical Pharmacokinetics 32 (Suppl. 1), 1–21.
fluoxetine. Journal of Psychiatric Research 44, 1154–1157. Price, J.L., Drevets, W.C., 2010. Neurocircuitry of mood disorders. Neuropsychophar-
Passingham, R.E., Bengtsson, S.L., Lau, H.C., 2010. Medial frontal cortex: from self- macology 35, 192–216.
generated action to reflection on one’s own performance. Trends in Cognitive Price, J.L., Drevets, W.C., 2012. Neural circuits underlying the pathophysiology of
Sciences 14, 16–21. mood disorders. Trends in Cognitive Sciences 16, 61–71.
Parihar, V.K., Hattiangady, B., Kuruba, R., Shuai, B., Shetty, A.K., 2011. Predictable Price, R.B., Shungu, D.C., Mao, X., Nestadt, P., Kelly, C., Collins, K.A., Murrough, J.W.,
chronic mild stress improves mood, hippocampal neurogenesis and memory. Charney, D.S., Mathew, S.J., 2009. Amino acid neurotransmitters assessed by
Molecular Psychiatry 16, 171–183. proton magnetic resonance spectroscopy: relationship to treatment resistance
Parsons, C.E., Young, K.S., Murray, L., Stein, A., Kringelbach, M.L., 2010. The functional in major depressive disorder. Biological Psychiatry 65, 792–800.
neuroanatomy of the evolving parent-infant relationship. Progress in Neurobi- Pringle, A., Browning, M., Cowen, P.J., Harmer, C.J., 2011. A cognitive neu-
ology 91, 220–241. ropsychological model of antidepressant drug action. Progress in Neuro-
Peluso, M.A., Glahn, D.C., Matsuo, K., Monkul, E.S., Najt, P., Zamarripa, F., Li, J., Lan- psychopharmacology & Biological Psychiatry 35, 1586–1592.
caster, J.L., Fox, P.T., Gao, J.H., Soares, J.C., 2009a. Amygdala hyperactivation in Post, R.M., 2004. Neurobiology of seizures and behavioral abnormalities. Epilepsia
untreated depressed individuals. Psychiatry Research 173, 158–161. 45 (Suppl. 2), 5–14.
Perry, P.J., Zeilmann, C., Arndt, S., 1994. Tricyclic antidepressant concentrations in Quera Salva, M.A., Hartley, S., Barbot, F., Alvarez, J.C., Lofaso, F., Guilleminault, C.,
plasma: an estimate of their sensitivity and specificity as a predictor of response. 2011. Circadian rhythms, melatonin and depression. Current Pharmaceutical
Journal of Clinical Psychopharmacology 14, 230–240. Design 17, 1459–1470.
Patel, J.G., Bartoszyk, G.D., Edwards, E., Ashby Jr., C.R., 2004. The highly selective Racagni, G., Riva, M.A., Molteni, R., Musazzi, L., Calabrese, F., Popoli, M., Tardito,
5-hydroxytryptamine (5-HT)2A receptor antagonist, EMD 281014, significantly D., 2011. Mode of action of agomelatine: synergy between melatonergic and 5-
increases swimming and decreases immobility in male congenital learned help- HT(2C) receptors. The World Journal of Biological Psychiatry: the Official Journal
less rats in the forced swim test. Synapse 52, 73–75. of the World Federation of Societies of Biological Psychiatry 12, 574–587.
Paton, J.J., Belova, M.A., Morrison, S.E., Salzman, C.D., 2006. The primate amygdala Radley, J.J., Jacobs, B.L., 2002. 5 HT1A receptor antagonist administration decreases
represents the positive and negative value of visual stimuli during learning. cell proliferation in dentate gyrus. Brain Research 955, 264–267.
Nature 439, 865–870. Radley, J.J., Arias, C.M., Sawchenko, P.E., 2006. Regional differentiation of the medial
Pechtel, P., Pizzagalli, D.A., 2011. Effects of early life stress on cognitive and affective prefrontal cortex in regulating adaptive responses to acute emotional stress. The
function: an integrated review of human literature. Psychopharmacology 214, Journal of Neuroscience: the Official Journal of the Society for Neuroscience 26,
55–70. 12967–12976.
Peluso, M.A., Glahn, D.C., Matsuo, K., Monkul, E.S., Najt, P., Zamarripa, F., Li, J., Lan- Raison, C.L., Miller, A.H., 2003. When not enough is too much: the role of insufficient
caster, J.L., Fox, P.T., Gao, J.H., Soares, J.C., 2009b. Amygdala hyperactivation in glucocorticoid signaling in the pathophysiology of stress-related disorders. The
untreated depressed individuals. Psychiatry Research 173 (2), 158–161. American Journal of Psychiatry 160, 1554–1565.
Perera, T.D., Dwork, A.J., Keegan, K.A., Thirumangalakudi, L., Lipira, C.M., Joyce, N., Raison, C.L., Capuron, L., Miller, H., 2006. Cytokines sing the blues: inflammation and
Lange, C., Higley, J.D., Rosoklija, G., Hen, R., Sackeim, H.A., Coplan, J.D., 2011. the pathogenesis of depression. Trends in Immunology 27, 24–31.
Necessity of hippocampal neurogenesis for the therapeutic action of antide- Rakofsky, J.J., Holtzheimer, P.E., Nemeroff, C.B., 2009. Emerging targets
pressants in adult nonhuman primates. PLoS One 6 (4), e17600. for antidepressant therapies. Current Opinion in Chemical Biology 13,
Peters, E.J., Slager, S.L., Kraft, J.B., Jenkins, G.D., Reinalda, M.S., McGrath, P.J., Hamil- 291–302.
ton, S.P., 2008. Pharmacokinetic genes do not influence response or tolerance to Rasmussen, K.G., 2009. Sham electroconvulsive therapy studies in depressive ill-
citalopram in the STAR*D sample. PLoS One 3 (4), e1872. ness: a review of the literature and consideration of the placebo phenomenon
Petrik, D., Lagace, D.C., Eisch, A.J., 2012. The neurogenesis hypothesis of affective in electroconvulsive therapy practice. The Journal of ECT 25, 54–59.
and anxiety disorders: are we mistaking the scaffolding for the building? Neu- Ray, R.D., Ochsner, K.N., Cooper, J.C., Robertson, E.R., Gabrieli, J.D., Gross, J.J., 2005.
ropharmacology 62, 21–34. Individual differences in trait rumination and the neural systems suppor-
Pezawas, L., Meyer-Lindenberg, A., Goldman, A.L., Verchinski, B.A., Chen, G., ting cognitive reappraisal. Cognitive, Affective & Behavioral Neuroscience 5,
Kolachana, B.S., Egan, M.F., Mattay, V.S., Hariri, A.R., Weinberger, D.R., 2008. 156–168.
Evidence of biologic epistasis between BDNF and SLC6A4 and implications for Read, J., Bentall, R., 2010. The effectiveness of electroconvulsive therapy: a literature
depression. Molecular Psychiatry 13, 709–716. review. Epidemiologia e Psichiatria Sociale 19, 333–347.
Pham, K., Nacher, J., Hof, P.R., McEwen, B.S., 2003. Repeated restraint stress Rentesi, G., Antoniou, K., Marselos, M., Fotopoulos, A., Alboycharali, J., Konstandi,
suppresses neurogenesis and induces biphasic PSA-NCAM expression in M., 2010. Long-term consequences of early maternal deprivation in sero-
the adult rat dentate gyrus. The European Journal of Neuroscience 17, tonergic activity and HPA function in adult rat. Neuroscience Letters 480,
879–886. 7–11.
2368 P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371

Revest, J.M., Dupret, D., Koehl, M., Funk-Reiter, C., Grosjean, N., Piazza, P.V., Abrous, Salvadore, G., van der Veen, J.W., Zhang, Y., Marenco, S., Machado-Vieira, R., Bau-
D.N., 2009. Adult hippocampal neurogenesis is involved in anxiety-related mann, J., Ibrahim, L.A., Luckenbaugh, D.A., Shen, J., Drevets, W.C., Zarate, C.A.,
behaviors. Molecular Psychiatry 14, 959–967. 2011. An investigation of amino-acid neurotransmitters as potential predictors
Revesz, D., Tjernstrom, M., Ben-Menachem, E., Thorlin, T., 2008. Effects of vagus of clinical improvement to ketamine in depression. The International Journal of
nerve stimulation on rat hippocampal progenitor proliferation. Experimental Neuropsychopharmacology 1, 1–10.
Neurology 214, 259–265. Salloway, S., Malloy, P., Kohn, R., Gillard, E., Duffy, J., Rogg, J., et al., 1996. MRI and
Reynolds, S.M., Berridge, K.C., 2008. Emotional environments retune the valence of neuropsychological differences in early- and late-life-onset geriatric depression.
appetitive versus fearful functions in nucleus accumbens. Nature Neuroscience Neurology 46, 1567–1574.
11, 423–425. Sampson, D., Muscat, R., Willner, P., 1991. Reversal of antidepressant action by
Richard, J.M., Berridge, K.C., 2011. Nucleus accumbens dopamine/glutamate interac- dopamine antagonists in an animal model of depression. Psychopharmacology
tion switches modes to generate desire versus dread: D(1) alone for appetitive 104, 491–495.
eating but D(1) and D(2) together for fear. The Journal of Neuroscience: the Samuels, B.A., Hen, R., 2011. Neurogenesis and affective disorders. The European
Official Journal of the Society for Neuroscience 31, 12866–12879. Journal of Neuroscience 33, 1152–1159.
Richardson, M.P., Strange, B.A., Dolan, R.J., 2004. Encoding of emotional memories Samuels, B.A., Leonardo, E.D., Gadient, R., Williams, A., Zhou, J., David, D.J., Gardier,
depends on amygdala and hippocampus and their interactions. Nature Neuro- A.M., Wong, E.H., Hen, R., 2011. Modeling treatment-resistant depression. Neu-
science 7, 278–285. ropharmacology 61, 408–413.
Risch, N., Herrell, R., Lehner, T., Liang, K.-L., Eaves, L., Hoh, J., et al., 2009. Interaction Sanacora, G., Zarate, C.A., Krystal, J.H., Manji, H.K., 2008. Targeting the glutamatergic
between the serotonin transporter gene (5-HTTLPR), stressful life events, and system to develop novel, improved therapeutics for mood disorders. Nature
risk of depression: a meta-analysis. Journal of the American Medical Association Reviews. Drug Discovery 7, 426–437.
301, 2462–2471. Santarelli, L., Saxe, M., Gross, C., Surget, A., Battaglia, F., Dulawa, S., Weisstaub, N.,
Rizk, P., Salazar, J., Raisman-Vozari, R., Marien, M., Ruberg, M., Colpaert, F., Debeir, T., Lee, J., Duman, R., Arancio, O., Belzung, C., Hen, R., 2003. Requirement of hip-
2006. The alpha2-adrenoceptor antagonist dexefaroxan enhances hippocampal pocampal neurogenesis for the behavioral effects of antidepressants. Science
neurogenesis by increasing the survival and differentiation of new granule cells. 301, 805–809.
Neuropsychopharmacology 31, 1146–1157. Santesso, D.L., Bogdan, R., Birk, J.L., Goetz, E.L., Holmes, A.J., Pizzagalli, D.A., 2012.
Robertson, J., Bowlby, J., 1952. Responses of young children to separation from their Neural responses to negative feedback are related to negative emotionality in
mothers. Courrier du Centre International de l’Enfance 2, 131–142. healthy adults. Social, Cognitive and Affective Neuroscience 7, 794–803.
Robins, C.J., Bagby, R.M., Rector, N.A., Lynch, R., Sidney, H., Kennedy, S.H., 1997. Santini, E., Ge, H., Ren, K., Peña de Ortiz, S., Quirk, G.J., 2004. Consolidation of fear
Sociotropy, autonomy, and patterns of symptoms in patients with major extinction requires protein synthesis in the medial prefrontal cortex. The Jour-
depression: a comparison of dimensional and categorical approaches. Cognitive nal of Neuroscience: the Official Journal of the Society for Neuroscience 24,
Therapy and Research 21, 285–300. 5704–5710.
Robinson, O.J., Sahakian, B.J., 2008. Recurrence in major depressive disorder: a neu- Sapolsky, R.M., 2000. Glucocorticoids and hippocampal atrophy in neuropsychiatric
rocognitive perspective. Psychological Medicine 38, 315–318. disorders. Archives of General Psychiatry 57, 925–935.
Robinson, O.J., Cools, R., Carlisi, C.O., Sahakian, B.J., Drevets, W.C., 2011. Ventral stri- Sartorius, A., Kiening, K.L., Kirsch, P., von Gall, C.C., Haberkorn, U., Unterberg, A.W.,
atum response during reward and punishment reversal learning in unmedicated Henn, F.A., Meyer-Lindenberg, A., 2010. Remission of major depression under
major depressive disorder. The American Journal of Psychiatry 169 (2), 152–159. deep brain stimulation of the lateral habenula in a therapy-refractory patient.
Rocha, B.A., Fleischer, R., Schaeffer, J.M., Rohrer, S.P., Hickey, G.J., 2005. 17 Beta- Biological Psychiatry 67 (2), e9–e11.
estradiol-induced antidepressant-like effect in the forced swim test is absent Savitz, J., Drevets, W.C., 2009. Bipolar and major depressive disorder: neuroimag-
in estrogen receptor-beta knockout (BERKO) mice. Psychopharmacology 179, ing the developmental-degenerative divide. Neuroscience and Biobehavioral
637–643. Reviews 33, 699–771.
Roiser, J.P., Levy, J., Fromm, S.J., Nugent, A.C., Talagala, S.L., Hasler, G., Henn, F.A., Saxena, S., Brody, A.L., Ho, M.L., Zohrabi, N., Maidment, K.M., Baxter Jr., L.R., 2003.
Sahakian, B.J., Drevets, W.C., 2009. The effects of tryptophan depletion on neural Differential brain metabolic predictors of response to paroxetine in obsessive-
responses to emotional words in remitted depression. Biological Psychiatry 66, compulsive disorder versus major depression. American Journal of Psychiatry
441–450. 160, 522–532.
Roiser, J.P., Elliott, R., Sahakian, B.J., 2012. Cognitive mechanisms of treatment in Schaefer, S.M., Jackson, D.C., Davidson, R.J., Aguirre, G.K., Kimberg, D.Y., Thompson-
depression. Neuropsychopharmacology 37, 117–136. Schill, S.L., 2002. Modulation of amygdalar activity by the conscious regulation
Roy, A., 1981. Role of past loss in depression. Archives of General Psychiatry 38, of negative emotion. Journal of Cognitive Neuroscience 14, 913–921.
301–302. Scharfman, H., Goodman, J., Macleod, A., Phani, S., Antonelli, C., Croll, S., 2005.
Rozeske, R.R., Evans, A.K., Frank, M.G., Watkins, L.R., Lowry, C.A., Maier, S.F., 2011. Increased neurogenesis and the ectopic granule cells after intrahippocampal
Uncontrollable, but not controllable, stress desensitizes 5-HT1A receptors in BDNF infusion in adult rats. Experimental Neurology 192, 348–356.
the dorsal raphe nucleus. The Journal of Neuroscience: the Official Journal of the Schlaepfer, T.E., Cohen, M.X., Frick, C., Kosel, M., Brodesser, D., Axmacher, N., Joe, A.Y.,
Society for Neuroscience 31, 14107–14115. Kreft, M., Lenartz, D., Sturm, V., 2008. Deep brain stimulation to reward circuitry
Rudy, J.W., 2009. Context representations, context functions, and the alleviates anhedonia in refractory major depression. Neuropsychopharmaco-
parahippocampal-hippocampal system. Learning & Memory 16, 573–585. logy 33, 368–377.
Ruhé, H.G., Mason, N.S., Schene, A.H., 2007. Mood is indirectly related to serotonin, Schloesser, R.J., Lehmann, M., Martinowich, K., Manji, H.K., Herkenham, M., 2010.
norepinephrine and dopamine levels in humans: a meta-analysis of monoamine Environmental enrichment requires adult neurogenesis to facilitate the recovery
depletion studies. Molecular Psychiatry 12, 331–359. from psychosocial stress. Molecular Psychiatry 15, 1152–1163.
Rush, A.J., Wisniewski, S.R., Zisook, S., Fava, M., Sung, S.C., Haley, C.L., Chan, H.N., Schmidt, H.D., Duman, R.S., 2007. The role of neurotrophic factors in adult hip-
Gilmer, W.S., Warden, D., Nierenberg, A.A., Balasubramani, G.K., Gaynes, B.N., pocampal neurogenesis, antidepressant treatments and animal models of
Trivedi, M.H., Hollon, S.D., 2011. Is prior course of illness relevant to acute or depressive-like behavior. Behavioural Pharmacology 18, 391–418.
longer-term outcomes in depressed out-patients? A STAR*D report. Psycholog- Schmidt, H.D., Duman, R.S., 2010. Peripheral BDNF produces antidepressant-
ical Medicine 19, 1–19. like effects in cellular and behavioral models. Neuropsychopharmacology 35,
Russo-Neustadt, A., Beard, R.C., Cotman, C.W., 1999. Exercise, antidepressant 2378–2391.
medications, and enhanced brain derived neurotrophic factor expression. Neu- Schoenfelder, E.N., Sandler, I.N., Wolchik, S., MacKinnon, D., 2011. Quality of social
ropsychopharmacology 21, 679–682. relationships and the development of depression in parentally-bereaved youth.
Saarelainen, T., Hendolin, P., Lucas, G., Koponen, E., Sairanen, M., MacDonald, E., Journal of Youth Adolescent 40, 85–96.
Agerman, K., et al., 2003. Activation of the TrkB neurotrophin receptor is induced Schüle, C., Baghai, T.C., Eser, D., Rupprecht, R., 2009. Hypothalamic-pituitary-
by antidepressant drugs and is required for antidepressant-induced behavioral adrenocortical system dysregulation and new treatment strategies in
effects. The Journal of Neuroscience: the Official Journal of the Society for Neu- depression. Expert Review of Neurotherapeutics 9, 1005–1019.
roscience 23, 349–357. Schulz, D., Mirrione, M.M., Henn, F.A., 2010. Cognitive aspects of congenital learned
Sabino, V., Cottone, P., Parylak, S.L., Steardo, L., Zorrilla, E.P., 2009. Sigma-1 receptor helplessness and its reversal by the monoamine oxidase (MAO)-B inhibitor
knockout mice display a depressive-like phenotype. Behavioural Brain Research deprenyl. Neurobiology of Learning and Memory 93, 291–301.
198, 472–476. Schultz, W., Dayan, P., Montague, P.R., 1997. A neural substrate of prediction and
Sahay, A., Scobie, K.N., Hill, A.S., O’Carroll, C.M., Kheirbek, M.A., Burghardt, reward. Science 275, 1593–1599.
N.S., Fenton, A.A., Dranovsky, A., Hen, R., 2011. Increasing adult hippocam- Segal, D.S., Kuczenski, R., Mandel, A.J., 1974. Theoretical implications of drug-
pal neurogenesis is sufficient to improve pattern separation. Nature 472, induced adaptive regulation for a biogenic amine hypothesis of affective
466–470. disorders. Biological Psychiatry 9, 147–159.
Sairanen, M., O’Leary, O.F., Knuuttila, J.E., Castrén, E., 2007. Chronic antidepres- Segal, Z.V., Williams, J.M., Teasdale, J.D., Gemar, M., 1996. A cognitive science per-
sant treatment selectively increases expression of plasticity-related proteins in spective on kindling and episode sensitization in recurrent affective disorder.
the hippocampus and medial prefrontal cortex of the rat. Neuroscience 144, Psychological Medicine 26, 371–380.
368–374. Seligman, M.E.P., 1975. Helplessness: On depression, Development and Death. Free-
Salamone, J.D., 1994. The involvement of nucleus accumbens dopamine in appetitive man, San Francisco.
and aversive motivation. Behavioural Brain Research 61, 117–133. Sesack, S.R., Grace, A.A., 2010. Cortico-Basal Ganglia reward network: microcircuitry.
Salome, N., Stemmelin, J., Cohen, C., Griebel, G., 2006. Differential roles of amygdaloid Neuropsychopharmacology 35, 27–47.
nuclei in the anxiolytic- and antidepressant-like effects of the V1b receptor Serretti, A., Kato, M., De Ronchi, D., Kinoshita, T., 2007. Meta-analysis of sero-
antagonist, SSR 149415, in rats. Psychopharmacology 187, 237–244. tonin transporter gene promoter polymorphism (5-HTTLPR) association with
P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371 2369

selective serotonin reuptake inhibitor efficacy in depressed patients. Molecular Souery, D., Amsterdam, J., de Montigny, C., Lecrubier, Y., Montgomery, S., Lipp,
Psychiatry 12, 247–257. O., Racagni, G., Zohar, J., Mendlewicz, J., 1999. Treatment resistant depression:
Shah, P.J., Glabus, M.F., Goodwin, G.M., Ebneier, K.P., 2002. Chronic treatment- methodological overview and operational criteria. European Neuropsychophar-
resistant depression and right frontal atrophy. British Journal of Psychiatry 180, macology: the Journal of the European College of Neuropsychopharmacology 9,
434–440. 83–91.
Shansky, R.M., Hamo, C., Hof, P.R., McEwen, B.S., Morrison, J.H., 2009. Stress-induced Souery, D., Oswald, P., Massat, I., Bailer, U., Bollen, J., Demyttenaere, K., Kasper, S.,
dendritic remodeling in the prefrontal cortex is circuit specific. Cerebral Cortex Lecrubier, Y., Montgomery, S., Serretti, A., Zohar, J., Mendlewicz, J., Group for the
19, 2479–2484. Study of Resistant Depression, 2007. Clinical factors associated with treatment
Sharot, T., Riccardi, A.M., Raio, C.M., Phelps, E.A., 2007. Neural mechanisms mediating resistance in major depressive disorder: results from a European multicenter
optimism bias. Nature 450, 102–105. study. The Journal of Clinical Psychiatry 68, 1062–1070.
Sharpley, C.F., Agnew, L.L., 2011. Cytokines and depression: findings, issues, and Stoy, M., Schlagenhauf, F., Sterzer, P., Bermpohl, F., Hägele, C., Suchotzki, K., Schmack,
treatment implications. Reviews in the Neurosciences 22, 295–302. K., Wrase, J., Ricken, R., Knutson, B., Adli, M., Bauer, M., Heinz, A., Ströhle, A.,
Sheline, Y.I., Gado, M.H., Kraemer, H.C., 2003. Untreated depression and hippocampal 2012. Hyporeactivity of ventral striatum towards incentive stimuli in unmedi-
volume loss. The American Journal of Psychiatry 160, 1516–1518. cated depressed patients normalizes after treatment with escitalopram. Journal
Sheline, Y.I., Barch, D.M., Price, J.L., Rundle, M.M., Vaishnavi, S.N., Snyder, A.Z., of Psychopharmacology 26, 677–678.
Mintun, M.A., Wang, S., Coalson, R.S., Raichle, M.E., 2009. The default mode net- Strekalova, T., Couch, Y., Kholod, N., Boyks, M., Malin, D., Leprince, P., Steinbusch,
work and self-referential processes in depression. Proceedings of the National H.M., 2011. Update in the methodology of the chronic stress paradigm: internal
Academy of Sciences of the United States of America 106, 1942–1947. control matters. Behavioral and Brain Functions 7, 9.
Sheline, Y.I., Price, J.L., Yan, Z., Mintun, M.A., 2010. Resting-state functional MRI in Strome, E.M., Zis, A.P., Doudet, D.J., 2007. Electroconvulsive shock enhances striatal
depression unmasks increased connectivity between networks via the dorsal dopamine D1 and D3 receptor binding and improves motor performance in 6-
nexus. Proceedings of the National Academy of Sciences of the United States of OHDA-lesioned rats. Journal of Psychiatry & Neuroscience 32, 193–202.
America 107, 11020–11025. Stroud, C.B., Davila, J., Moyer, A., 2008. The relationship between stress and
Sher, L., Oquendo, M.A., Galfalvy, H.C., Cooper, T.B., Mann, J.J., 2004. The number of depression in firstonsets versus recurrences: a meta-analytic review. Journal
previous depressive episodes is positively associated with cortisol response to of Abnormal Psychology 117, 206–213.
fenfluramine administration. Annals of the New York Academy of Sciences 1032, Stroud, C.B., Davila, J., Hammen, C., Vrshek-Schallhorn, S., 2011. Severe and nonse-
283–286. vere events in first onsets versus recurrences of depression: evidence for stress
Sherman, A.D., Petty, F., 1980. Neurochemical basis of the action of antidepressants sensitization. Journal of Abnormal Psychology 120, 142–154.
on learned helplessness. Behavioural Neural Biology 30, 119–134. Sullivan, R.M., 2004. Hemispheric asymmetry in stress processing in rat prefrontal
Shirayama, Y., Chen, A.C., Nakagaway, S., Russell, D.S., Duman, R.S., 2002. Brain- cortex and the role of mesocortical dopamine. Stress 7, 131–143.
derived neurotrophic factor produces antidepressant effects in behavioral Sullivan, P.F., Neale, M.C., Kendler, K.S., 2000. Genetic epidemiology of major
models of depression. The Journal of Neuroscience: the Official Journal of the depression: review and meta-analysis. The American Journal of Psychiatry 157,
Society for Neuroscience 22, 3251–3261. 1552–1562.
Shirayama, Y., Muneoka, K., Fukumoto, M., Tadokoro, S., Fukami, G., Hashimoto, Sun, Y., Jin, K., Childs, J.T., Xie, L., Mao, X.O., Greenberg, D.A., 2006. Vascular endothe-
K., Iyo, M., 2011. Infusions of allopregnanolone into the hippocampus and lial growth factor-B (VEGFB) stimulates neurogenesis: evidence from knockout
amygdala, but not into the nucleus accumbens and medial prefrontal cortex, mice and growth factor administration. Developmental Biology 289, 329–335.
produce antidepressant effects on the learned helplessness rats. Hippocampus Surget, A., Saxe, M., Leman, S., Ibarguen-Vargas, Y., Chalon, S., Griebel, G., Hen, R.,
21, 1105–1113. Belzung, C., 2008. Drug-dependent requirement of hippocampal neurogenesis
Shiroma, P.R., Geda, Y.E., Mrazek, D.A., 2010. Pharmacogenomic implications of in a model of depression and of antidepressant reversal. Biological Psychiatry
variants of monoaminergic-related genes in geriatric psychiatry. Pharmacoge- 64, 293–301.
nomics 11, 1305–1330. Surget, A., Wangj, Y., Leman, S., Ibarguen-Vargas, Y., Edgar, N., Griebel, G., Belzung, C.,
Shumake, J., Gonzalez-Lima, F., 2003. Brain systems underlying susceptibility to Sibille, E., 2009. Corticolimbic transcriptome changes are state-dependent and
helplessness and depression. Behavioral and Cognitive Neuroscience Reviews region-specific in a rodent model of depression and of antidepressant reversal.
2, 198–221. Neuropsychopharmacology 34 (6), 1363–1380.
Shumake, J., Poremba, A., Edwards, E., Gonzalez-Lima, F., 2000. Congenital helpless Surget, A., Tanti, A., Leonardo, E.D., Laugeray, A., Rainer, Q., Touma, C., Palme, R.,
rats as a genetic model for cortex metabolism in depression. Neuroreport 11, Griebel, G., Ibarguen-Vargas, Y., Hen, R., Belzung, C., 2011. Antidepressants
3793–3798. recruit new neurons to improve stress response regulation. Molecular Psychiatry
Shumake, J., Edwards, E., Gonzalez-Lima, F., 2003. Opposite metabolic changes in the 16 (12), 1177–1188.
habenula and ventral tegmental area of a genetic model of helpless behavior. Surguladze, S., Brammer, M.J., Keedwell, P., Giampietro, V., Young, A.W., Travis,
Brain Research 963, 274–281. M.J., Williams, S.C., Phillips, M.L., 2005. A differential pattern of neural response
Shumake, J., Colorado, R.A., Barrett, D.W., Gonzalez-Lima, F., 2010. Metabolic map- toward sad versus happy facial expressions in major depressive disorder. Bio-
ping of the effects of the antidepressant fluoxetine on the brains of congenitally logical Psychiatry 57, 201–209.
helpless rats. Brain Research 1343, 218–225. Swaab, D.F., Bao, A.M., Lucassen, P.J., 2005. The stress system in the human brain in
Siegle, G.J., Carter, C.S., Thase, M.E., 2006. Use of FMRI to predict recovery from depression and neurodegeneration. Ageing Research Reviews 4, 141–194.
unipolar depression with cognitive behavior therapy. The American Journal of Takase, L.F., Nogueira, M.I., Baratta, M., Bland, S.T., Watkins, L.R., Maier, S.F., Fornal,
Psychiatry 163, 735–738. C.A., Jacobs, B.L., 2004. Inescapable shock activates serotonergic neurons in all
Siegle, G.J., Thompson, W., Carter, C.S., Steinhauer, S.R., Thase, M.E., 2007. Increased raphe nuclei of rat. Behavioural Brain Research 153, 233–239.
amygdala and decreased dorsolateral prefrontal BOLD responses in unipo- Taliaz, D., Stall, N., Dar, D.E., Zangen, A., 2010. Knockdown of brain-derived neu-
lar depression: related and independent features. Biological Psychiatry 61, rotrophic factor in specific brain sites precipitates behaviors associated with
198–209. depression and reduces neurogenesis. Molecular Psychiatry 15, 80–92.
Sienaert, P., 2011. What we have learned about electroconvulsive therapy and its Tanti, A., Belzung, C., 2010. Open questions in current models of antidepressant
relevance for the practising psychiatrist. Canadian Journal of Psychiatry 56, 5–12. action. British Journal of Pharmacology 585, 1187–1200.
Slavich, G.M., O’Donovan, A., Epel, E.S., Kemeny, M.E., 2010. Black sheep get the blues: Tanda, G., Frau, R., Di Chiara, G., 1996. Chronic desipramine and fluoxetine
a psychobiological model of social rejection and depression. Neuroscience and differentially affect extracellular dopamine in the rat prefrontal cortex. Psy-
Biobehavioral Reviews 35, 39–45. chopharmacology 127, 83–87.
Slavich, G.M., Monroe, S.M., Gotlib, I.H., 2011. Early parental loss and depression Tardito, D., Perez, J., Tiraboschi, E., Musazzi, L., Racagni, G., Popoli, M., 2006. Signaling
history: associations with recent life stress in major depressive disorder. Journal pathways regulating gene expression, neuroplasticity, and neurotrophic mech-
of Psychiatric Research 45, 1146–1152. anisms in the action of antidepressants: a critical overview. Pharmacological
Sloman, L., Gilbert, P., Hasey, G., 2003. Evolved mechanisms in depression: the role Reviews 58, 115–134.
and interaction of attachment and social rank in depression. Journal of Affective Tata, D.A., Anderson, B.J., 2010. The effects of chronic glucocorticoid exposure on
Disorders 74, 107–121. dendritic length, synapse numbers and glial volume in animal models: implica-
Slotkin, T.A., Seidler, F.J., Ritchie, J.C., 1998. Effects of aging and glucocorticoid tions for hippocampal volume reductions in depression. Physiology & Behavior
treatment on monoamine oxidase subtypes in rat cerebral cortex: therapeutic 99, 186–193.
implications. Brain Research Bulletin 47, 345–348. Tata, D.A., Marciano, V.A., Anderson, B.J., 2006. Synapse loss from chronically
Snyder, J.S., Cameron, H.A., 2012. Could adult hippocampal neurogenesis be relevant elevated glucocorticoids: relationship to neuropil volume and cell num-
for human behavior? Behavioural Brain Research 227 (2), 384–390. ber in hippocampal area CA3. The Journal of Comparative Neurology 498,
Snyder, J.S., Soumier, A., Brewer, M., Pickel, J., Cameron, H.A., 2011. Adult hippocam- 363–374.
pal neurogenesis buffers stress responses and depressive behaviour. Nature 476, Taxidis, J., Coomber, B., Mason, R., Owen, M., 2010. Assessing cortico-hippocampal
458–461. functional connectivity under anesthesia and kainic acid using generalized par-
Solomon, D.A., Keller, M.B., Leon, A.C., Mueller, T.I., Lavori, P.W., Shea, M.T., Coryell, tial directed coherence. Biological Cybernetics 102, 327–340.
W., Warshaw, M., Turvey, C., Maser, J.D., Endicott, J., 2000. Multiple recurr- Taylor, S.E., Eisenberger, N.I., Saxbe, D., Lehman, B.J., Lieberman, M.D., 2006. Neu-
ences of major depressive disorder. The American Journal of Psychiatry 157, ral responses to emotional stimuli are associated with childhood family stress.
229–233. Biological Psychiatry 60, 296–301.
Sonawalla, S.B., Papakostas, G.I., Petersen, T.J., Yeung, A.S., Smith, M.M., Sickinger, Thiebaut, F., Tsuruo, T., Hamada, H., Gottesman, M.M., Pastan, I., Willingham,
A.H., Gordon, J., Israel, J.A., Tedlow, J.R., Lamon-Fava, S., Fava, M., 2002. Elevated M.C., 1987. Cellular localization of the multidrug-resistance gene product P-
cholesterol levels associated with nonresponse to fluoxetine treatment in major glycoprotein in normal human tissues. Proceedings of the National Academy of
depressive disorder. Psychosomatics 43, 310–316. Science of the USA 84, 7735–7738.
2370 P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371

Thierry, A.M., Gioanni, Y., Dégénétais, E., Glowinski, J., 2000. Hippocampo-prefrontal cells in the hippocampus and ameliorates depressive symptoms in chronically
cortex pathway: anatomical and electrophysiological characteristics. Hip- stressed rats. Neuroscience Letters 455, 178–182.
pocampus 10, 411–419. Victor, T.A., Furey, M.L., Fromm, S.J., Ohman, A., Drevets, W.C., 2010. Relationship
Todorovic, C., Sherrin, T., Pitts, M., Hippel, C., Rayner, M., Spiess, J., 2009. Suppression of emotional processing to masked faces in the amgydala to mood state and
of the MEK/ERK signaling pathway reverses depression-like behaviors of CRF2- treatment in major depressive disorder. Archives of General Psychiatry 67,
deficient mice. Neuropsychopharmacology 34, 1416–1426. 1128–1138.
Tordera, R.M., Totterdell, S., Wojcik, S.M., Brose, N., Elizalde, N., Lasheras, B., Videbech, P., Ravnkilde, B., 2004. Hippocampal volume and depression: a
Del Rio, J., 2007. Enhanced anxiety, depressive-like behaviour and impaired meta-analysis of MRI studies. The American Journal of Psychiatry 161,
recognition memory in mice with reduced expression of the vesicular glu- 1957–1966.
tamate transporter 1 (VGLUT1). The European Journal of Neuroscience 25, Vigo, D.V., Baldessarini, R.J., 2009. Anticonvulsants in the treatment of major depres-
281–290. sive disorder: an overview. Harvard Review of Psychiatry 17, 231–241.
Tordera, R.M., Garcia-García, A.L., Elizalde, N., Segura, V., Aso, E., Venzala, E., Ramírez, Viikki, M., Anttila, S., Kampman, O., Illi, A., Huuhka, M., Setälä-Soikkeli, E., Mononen,
M.J., Del Rio, J., 2011. Chronic stress and impaired glutamate function elicit N., Lehtimäki, T., Leinonen, E., 2010. Vascular endothelial growth factor (VEGF)
a depressive-like phenotype and common changes in gene expression in the polymorphism is associated with treatment resistant depression. Neuroscience
mouse frontal cortex. European Neuropsychopharmacology: the Journal of the Letters 25, 105–108.
European College of Neuropsychopharmacology 21, 23–32. Vollmayr, B., Simonis, C., Weber, S., Gass, P., Henn, F., 2003. Reduced cell prolifer-
Treadway, M.T., Zald, D.H., 2011. Reconsidering anhedonia in depression: lessons ation in the dentate gyrus is not correlated with the development of learned
from translational neuroscience. Neuroscience and Biobehavioral Reviews 35, helplessness. Biological Psychiatry 54, 1035–1040.
537–555. Vollmayr, B., Bachteler, D., Vengeliene, V., Gass, P., Spanagel, R., Henn, F.,
Tremblay, P., Blier, P., 2006. Catecholaminergic strategies for the treatment of major 2004. Rats with congenital learned helplessness respond less to sucrose but
depression. Current Drug Targets 7, 149–158. show no deficits in activity or learning. Behavioural Brain Research 150,
Treynor, W., Gonzales, R., Noel-Hoeksema, S., 2003. Rumination reconsidered: a 217–221.
psychometric analysis. Cognitive Therapeutics Research 27, 247–259. Vythilingam, M., Heim, C., Newport, D.J., Miller, A.H., Vermetten, E., Anderson, E.,
Tricomi, E.M., Delgado, M.R., Fiez, J.A., 2004. Modulation of caudate activity by action Bronen, R., Staib, L., Charney, D.S., Nemeroff, C.B., Bremner, J.D., 2002. Childhood
contingency. Neuron 41, 281–292. trauma associated with smaller hippocampal volume in women with major
Trivedi, M.H., Rush, A.J., Wisniewski, S.R., Nierenberg, A.A., Warden, D., Ritz, L., depression. The American Journal of Psychiatry 159, 2072–2080.
Norquist, G., Howland, R.H., Lebowitz, B., McGrath, P.J., Shores-Wilson, K., Biggs, Wager, T.D., Davidson, M.L., Hughes, B.L., Lindquist, M.A., Ochsner, K.N., 2008.
M.M., Balasubramani, G.K., Fava, M., STAR*D Study Team, 2006. Evaluation of Prefrontal-subcortical pathways mediating successful emotion regulation. Neu-
outcomes with citalopram for depression using measurement-based care in ron 59, 1037–1050.
STAR*D: implications for clinical practice. The American Journal of Psychiatry Wagner, G., Koch, K., Schachtzabel, C., Schultz, C.C., Sauer, H., Schlösser, R.G., 2011.
163, 28–40. Structural brain alterations in patients with major depressive disorder and high
Tronel, S., Fabre, A., Charrier, V., Oliet, S.H., Gage, F.H., Abrous, D.N., 2010. Spa- risk for suicide: evidence for a distinct neurobiological entity? NeuroImage 54,
tial learning sculpts the dendritic arbor of adult-born hippocampal neurons. 1607–1614.
Proceedings of the National Academy of Sciences of the United States of America Walf, A., Frye, C.A., 2007. Administration of estrogen receptor beta-specific selec-
107, 7963–7968. tive estrogen receptor modulators to the hippocampus decrease anxiety and
Tronel, S., Belnoue, L., Grosjean, N., Revest, J.M., Piazza, P.V., Koehl, M., Abrous, D.N., depressive behavior of ovariectomized rats. Pharmacology, Biochemistry, and
2012. Adult-born neurons are necessary for extended contextual discrimination. Behavior 86, 407–414.
Hippocampus 22, 292–298. Wallace, T.L., Stellitano, K.E., Neve, R.L., Duman, R.S., 2004. Effects of cyclic adeno-
Tsankova, N.M., Berton, O., Renthal, W., Kumar, A., Neve, R.L., Nestler, E.J., 2006. sine monophosphate response element binding protein overexpression in the
Sustained hippocampal chromatin regulation in a mouse model of depression basolateral amygdala on behavioral models of depression and anxiety. Biological
and antidepressant action. Nature Neuroscience 9, 519–525. Psychiatry 56, 151–160.
Tyring, S., Gottlieb, A., Papp, K., Gordon, K., Leonardi, C., Wang, A., Lalla, D., Woolley, Wang, D.D., Kriegstein, A.R., 2009. Defining the role of GABA in cortical development.
M., Jahreis, A., Zitnik, R., Cella, D., Krishnan, R., 2006. Etanercept and clinical Journal of Physiology 587, 1873–1879.
outcomes, fatigue, and depression in psoriasis: double-blind placebo-controlled Wang, J.W., David, D.J., Monckton, J.E., Battaglia, F., Hen, R., 2008. Chronic fluoxe-
randomised phase III trial. Lancet 367, 29–35. tine stimulates maturation and synaptic plasticity of adult born hippocampal
Uher, R., McGuffin, P., 2008. The moderation by the serotonin transporter gene of granule cells. The Journal of Neuroscience: the Official Journal of the Society for
environmental adversity in the aetiology of mental illness: review and method- Neuroscience 28, 1374–1384.
ological analysis. Molecular Psychiatry 13, 131–146. Warden, D., Rush, A.J., Trivedi, M.H., Fava, M., Wisniewski, S.R., 2007. The STAR*D
Ullsperger, M., von Cramon, D.Y., 2003. Error monitoring using external feedback: Project results: a comprehensive review of findings. Current Psychiatry Reports
specific roles of the habenular complex, the reward system, and the cingulate 9, 449–459.
motor area revealed by functional magnetic resonance imaging. The Jour- Warner-Schmidt, J.L., Duman, R.S., 2007. VEGF is an essential mediator of the neu-
nal of Neuroscience: the Official Journal of the Society for Neuroscience 23, rogenic and behavioral actions of antidepressants. Proceedings of the National
4308–4314. Academy of Sciences of the United States of America 104, 4647–4652.
Ulrich-Lai, Y.M., Herman, J.P., 2009. Neural regulation of endocrine and autonomic Watkins, E.R., Mullan, E., Wingrove, J., Rimes, K., Steiner, H., Bathurst, N., Eastman, R.,
stress responses. Nature Reviews. Neuroscience 10, 397–409. Scott, J., 2011. Rumination-focused cognitive-behavioural therapy for residual
Vago, D.R., Epstein, J., Catenaccio, E., Stern, E., 2011. Identification of neural targets depression: phase II randomised controlled trial. The British Journal of Psychia-
for the treatment of psychiatric disorders: the role of functional neuroimaging. try: the Journal of Mental Science 199, 317–322.
Neurosurgery Clinics of North America 22, 279–305. Wei, L., Meaney, M.J., Duman, R.S., Kaffman, A., 2011. Affiliative behavior requires
van Eijndhoven, P., van Wingen, G., van Oijen, K., Rijpkema, M., Goraj, B., Jan Verkes, juvenile, but not adult neurogenesis. The Journal of Neuroscience: the Official
R., Oude Voshaar, R., Fernández, G., Buitelaar, J., Tendolkar, I., 2009. Amygdala Journal of the Society for Neuroscience 31, 14335–14345.
volume marks the acute state in the early course of depression. Biological Psy- Weikop, P., Kehr, J., Scheel-Krüger, J., 2007a. Reciprocal effects of combined admin-
chiatry 65, 812–818. istration of serotonin, noradrenaline and dopamine reuptake inhibitors on
van Praag, H., Kempermann, G., Gage, F.H., 1999. Running increases cell proliferation serotonin and dopamine levels in the rat prefrontal cortex: the role of 5-HT
and neurogenesis in the adult mouse dentate gyrus. Nature Neuroscience 2, 1A receptors. Journal of Psychopharmacology 21, 795–804.
266–270. Weikop, P., Yoshitake, T., Kehr, J., 2007b. Differential effects of adjunctive
van Tol, M.J., van der Wee, N.J., van den Heuvel, O.A., Nielen, M.M., Demenescu, methylphenidate and citalopram on extracellular levels of serotonin, noradren-
L.R., Aleman, A., Renken, R., van Buchem, M.A., Zitman, F.G., Veltman, D.J., 2010. aline and dopamine in the rat brain. European Neuropsychopharmacology: the
Regional brain volume in depression and anxiety disorders. Archives of General Journal of the European College of Neuropsychopharmacology 17, 658–671.
Psychiatry 67, 1002–1011. Weisler-Frank, J., Maier, S.F., Watkins, L.R., 2005. Immune-to-brain communica-
van Wingen, G.A., van Eijndhoven, P., Cremers, H.R., Tendolkar, I., Verkes, R.J., Buite- tion dynamically modulates pain: physiological and pathological consequences.
laar, J.K., Fernández, G., 2010. Neural state and trait bases of mood-incongruent Brain, Behavior, and Immunity 19, 104–111.
memory formation and retrieval in first-episode major depression. Journal of Wetter, E.K., Hankin, B.L., 2009. Mediational pathways through which positive
Psychiatric Research 44, 527–534. and negative emotionality contribute to anhedonic symptoms of depression:
Vang, F.J., Ryding, E., Träskman-Bendz, L., van Westen, D., Lindström, M.B., 2010. Size a prospective study of adolescents. Journal of Abnormal Child Psychology 37,
of basal ganglia in suicide attempters, and its association with temperament and 507–520.
serotonin transporter density. Psychiatry Research 183, 177–179. WHO, 2008. The Global Burden of Disease: 2004 Update. WHO.
Varga, V., Kocsis, B., Sharp, T., 2003. Electrophysiological evidence for convergence Widom, C.S., DuMont, K., Czaja, S.J.A., 2007. Prospective investigation of major
of inputs from the medial prefrontal cortex and lateral habenula on single neu- depressive disorder and comorbidity in abused and neglected children grown
rons in the dorsal raphe nucleus. The European Journal of Neuroscience 17, up. Archives of General Psychiatry 64, 49–56.
280–286. Wilkinson, P.O., Goodyer, I.M., 2011. Childhood adversity and allostatic overload of
Vazquez-Borsetti, P., Celada, P., Cortes, R., Artigas, F., 2011. Simultaneous projections the hypothalamic–pituitary–adrenal axis: a vulnerability model for depressive
from prefrontal cortex to dopaminergic and serotonergic nuclei. The Interna- disorders. Development and Psychopathology 23, 1017–1037.
tional Journal of Neuropsychopharmacology 14, 289–302. Wilkinson, M.B., Xiao, G., Kumar, A., Laplant, Q., Renthal, W., Sikder, D., Kodadek,
Veena, J., Srikumar, B.N., Raju, T.R., Shankaranarayana Rao, B.S., 2009. Exposure to T.J., Nestler, E.J., 2009. Imipramine treatment and resiliency exhibit similar chro-
enriched environment restores the survival and differentiation of new born matin regulation in the mouse nucleus accumbens in depression models. The
P. Willner et al. / Neuroscience and Biobehavioral Reviews 37 (2013) 2331–2371 2371

Journal of Neuroscience: the Official Journal of the Society for Neuroscience 29, Yau, J.L., Noble, J., Thomas, S., Kerwin, R., Morgan, P.E., Lightman, S., Seckl, J.R.,
7820–7832. Pariante, C.M., 2007. The antidepressant desipramine requires the ABCB1
Willner, P., 1985. Depression: A Psychobiological Synthesis. Wiley, Chichester. (Mdr1)-type p-glycoprotein to upregulate the glucocorticoid receptor in mice.
Willner, P., 1997a. Validity, reliability and utility of the chronic mild stress (CMS) Neuropsychopharmacology 32, 2520–2529.
model of depression: a ten-year review and evaluation. Psychopharmacology Yin, H.H., Ostlund, S.B., Knowlton, B.J., Balleine, B.W., 2005. The role of the dor-
134, 319–329. somedial striatum in instrumental conditioning. The European Journal of
Willner, P., 1997b. The mesolimbic dopamine system as a substrate for (rapid?) Neuroscience 22, 513–523.
antidepressant action. International Clinical Psychopharmacology 12 (Suppl. 3), Yin, H.H., Knowlton, B.J., Balleine, B.W., 2006. Inactivation of dorsolateral striatum
S7–S14. enhances sensitivity to changes in the action-outcome contingency in instru-
Willner, P., Mitchell, P.J., 2002a. Animal models of depression: a diathesis-stress mental conditioning. Behavioural Brain Research 166, 189–196.
approach. In: D’Haenen, H., Den Boer, H., Willner, P. (Eds.), Biological Psychiatry, Yoshimizu, T., Chaki, S., 2004. Increased cell proliferation in the adult mouse hip-
vol. 2. Wiley, pp. 703–726. pocampus following chronic administration of group II metabotropic glutamate
Willner, P., Mitchell, P.J., 2002b. The validity of animal models of predisposition to receptor antagonist, MGS0039. Biochemical and Biophysical Research Commu-
depression. Behavioural Pharmacology 13, 169–188. nications 315, 493–496.
Willner, P., Scheel-Krüger, J. (Eds.), 1991. The Mesolimbic Dopamine System: From Yoshimura, S., Okamoto, Y., Onoda, K., Matsunaga, M., Ueda, K., Suzuki, S., Shige-
Motivation to Action. Wiley, Chichester, pp. 225–250. toyamawaki, 2010. Rostral anterior cingulate cortex activity mediates the
Willner, P., Towell, A., Sampson, D., Muscat, R., Sophokleous, S., 1987. Reduction relationship between the depressive symptoms and the medial prefrontal cortex
of sucrose preference by chronic mild stress and its restoration by a tricyclic activity. Journal of Affective Disorders 122, 76–85.
antidepressant. Psychopharmacology 93, 358–364. Zarate Jr., C.A., Singh, J.B., Quiroz, J.A., De Jesus, G., Denicoff, K.K., Luckenbaugh,
Willner, P., Muscat, R., Papp, M., 1992. Chronic mild stress-induced anhedonia: a D.A., Manji, H.K., Charney, D.S., 2006a. Double-blind, placebo-controlled study
realistic animal model of depression. Neuroscience and Biobehavioral Reviews of memantine in the treatment of major depression. The American Journal of
16, 525–534. Psychiatry 163, 153–155.
Willner, P., Hale, A.S., Argyropoulos, S.V., 2005. Dopaminergic mechanism of Zarate Jr., C.A., Singh, J.B., Carlson, P.J., Brutsche, N.E., Ameli, R., Luckenbaugh, D.A.,
antidepressant action in depressed patients. Journal of Affective Disorders 86, Charney, D.S., Manji, H.K., 2006b. A randomized trial of an N-methyl-D-aspartate
37–45. antagonist in treatment-resistant major depression. Archives of General Psychi-
Winter, C., Vollmayr, B., Djodari-Irani, A., Klein, J., Sartorius, A., 2011. Pharmaco- atry 63, 856–864.
logical inhibition of the lateral habenula improves depressive-like behavior in Zetterström, T.S., Pei, Q., Grahame-Smith, D.G., 1998. Repeated electroconvulsive
an animal model of treatment resistant depression. Behavioural Brain Research shock extends the duration of enhanced gene expression for BDNF in rat
216, 463–465. brain compared with a single administration. Brain Research. Molecular Brain
Wise, R.A., 2008. Dopamine and reward: the anhedonia hypothesis 30 years on. Research 57, 106–110.
Neurotoxicity Research 14, 169–183. Zhang, J., Wang, J., Wu, O., Kuang, W., Huang, X., He, Y., Gong, O., 2011. Disrupted
Wong, E.Y., Herbert, J., 2004. The corticoid environment: a determining factor for brain connectivity networks in drug-naive, first-episode major depressive dis-
neural progenitors’ survival in the adult hippocampus. The European Journal of order. Biological Psychiatry 70, 334–342.
Neuroscience 20, 2491–2498. Zhong, M., Wang, X., Xiao, J., Yi, J., Zhu, X., Liao, J., Wang, W., Yao, S., 2011. Amyg-
Woolley, C.S., Gould, E., McEwen, B.S., 1990. Exposure to excess glucocorticoids alters dala hyperactivation and prefrontal hypoactivation in subjects with cognitive
dendritic morphology of adult hippocampal pyramidal neurons. Brain Research vulnerability to depression. Biological Psychology 88, 233–242.
531, 225–231. Zisook, S., Corruble, E., Duan, N., Iglewicz, A., Karam, E.G., Lanouette, N., Lebowitz,
Wu, J., Buchsbaum, M.S., Gillin, J.C., Tang, C., Cadwell, S., Wiegand, M., et al., 1999. B., Pies, R., Reynolds, C., Seay, K., Katherine Shear, M., Simon, N., Young, I.T.,
Prediction of antidepressant effects of sleep deprivation by metabolic rates in 2012. The bereavement exclusion and DSM-5. Depression and Anxiety 29,
the ventral anterior cingulate and medial prefrontal cortex. American Journal of 425–443.
Psychiatry 156, 1149–1158. Zobel, A., Maier, W., 2010. Pharmacogenetics of antidepressive treat-
Yacoubi, M.E., Popa, D., Martin, B., Zimmer, L., Hamon, M., Adrien, J., Vaugeois, J.M., ment. European Archives of Psychiatry and Clinical Neuroscience 260,
2011. Genetic association between helpless trait and depression-related pheno- 407–417.
types: evidence from crossbreeding studies with H/Rouen and NH/Rouen mice. Zobel, A.W., Nickel, T., Sonntag, A., Uhr, M., Holsboer, F., Ising, M., 2001. Cortisol
The International Journal of Neuropsychopharmacology 3, 1–12. response in the combined dexamethasone/CRH test as predictor of relapse in
Yalcin, I., Coubard, S., Bodard, S., Chalon, S., Belzung, C., 2008. Effects patients with remitted depression: a prospective study. Journal of Psychiatric
of 5,7-dihydroxytryptamine lesion of the dorsal raphe nucleus on the Research 35, 83–94.
antidepressant-like action of tramadol in the unpredictable chronic mild stress Zorrilla, E.P., Koob, G.F., 2010. Progress in corticotropin-releasing factor-1 antagonist
in mice. Psychopharmacology 200, 497–507. development. Drug Discovery Today 15, 371–383.
Yang, L.M., Hu, B., Xia, Y.H., Zhang, B.L., Zhao, H., 2008. Lateral habenula lesions Zou, Y.F., Ye, D.Q., Feng, X.L., Su, H., Pan, F.M., Liao, F.F., 2010. Meta-analysis
improve the behavioral response in depressed rats via increasing the serotonin of BDNF Val66Met polymorphism association with treatment response in
level in dorsal raphe nucleus. Behavioural Brain Research 188, 84–90. patients with major depressive disorder. European Neuropsychopharmaco-
Yanpallewar, S.U., Fernandes, K., Marathe, S.V., Vadodaria, K.C., Jhaveri, D., Rom- logy: the Journal of the European College of Neuropsychopharmacology 20,
melfanger, K., Ladiwala, U., Jha, S., Muthig, V., Hein, L., Bartlett, P., Weinshenker, 535–544.
D., Vaidya, V.A., 2010. Alpha2-adrenoceptor blockade accelerates the neuro- Zunszain, P.A., Anacker, C., Cattaneo, A., Carvalho, L.A., Pariante, C.M.,
genic, neurotrophic, and behavioral effects of chronic antidepressant treatment. 2011. Glucocorticoids, cytokines and brain abnormalities in depres-
The Journal of Neuroscience: the Official Journal of the Society for Neuroscience sion. Progress in Neuro-psychopharmacology & Biological Psychiatry 35,
30, 1096–1109. 722–729.

You might also like