You are on page 1of 14

Neuroscience and Biobehavioral Reviews 36 (2012) 1920–1933

Contents lists available at SciVerse ScienceDirect

Neuroscience and Biobehavioral Reviews


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

Review

Deep brain stimulation for treatment-resistant depression:


Efficacy, safety and mechanisms of action
Rodney J. Anderson a , Mark A. Frye b , Osama A. Abulseoud b ,
Kendall H. Lee c,d , Jane A. McGillivray a , Michael Berk e , Susannah J. Tye a,c,f,∗
a
School of Psychology, Deakin University, Burwood, VIC, Australia
b
Department of Psychiatry, Mayo Clinic, Rochester, MN, USA
c
Department of Neurosurgery, Mayo Clinic, Rochester, MN, USA
d
Department of Physiology and Biomedical Engineering, Mayo Clinic, Rochester, MN, USA
e
School of Medicine, Deakin University, Geelong, VIC, Australia
f
Mental Health Research Institute, Melbourne, VIC, Australia

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

Article history: Deep brain stimulation (DBS), a neuromodulation therapy that has been used successfully in the treatment
Received 26 February 2012 of symptoms associated with movement disorders, has recently undergone clinical trials for individu-
Received in revised form 6 June 2012 als suffering from treatment-resistant depression (TRD). Although the small patient numbers and open
Accepted 10 June 2012
label study design limit our ability to identify optimum targets and make definitive conclusions about
treatment efficacy, a review of the published research demonstrates significant reductions in depressive
Keywords:
symptomatology and high rates of remission in a severely treatment-resistant patient group. Despite
Deep brain stimulation
these encouraging results, an incomplete understanding of the mechanisms of action underlying the
Treatment resistant depression
Subgenual cingulate gyrus
therapeutic effects of DBS for TRD is highlighted, paralleling the incomplete understanding of the neu-
Nucleus accumbens roanatomy of mood regulation and treatment resistance. Proposed mechanisms of action include short
Ventral capsule and long-term local effects of stimulation at the neuronal level, to modulation of neural network activity.
Ventral striatum
Lateral Habenula © 2012 Elsevier Ltd. All rights reserved.
Inferior thalamic peduncle
Mood

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1921
1.1. Scope of this review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1921
2. Target selection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1921
2.1. Subgenual cingulate gyrus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1923
2.2. Ventral capsule/ventral striatum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1923
2.3. Nucleus accumbens . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1923
2.4. Inferior thalamic peduncle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1923
2.5. Lateral habenula . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1923
3. Efficacy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1924
3.1. Subgenual cingulate gyrus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1924
3.2. Ventral capsule/ventral striatum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1925
3.3. Nucleus accumbens . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1925
3.4. Inferior thalamic peduncle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1925
3.5. Lateral habenula . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1925
3.6. Safety and adverse events . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1925
3.7. Efficacy evaluation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1926

∗ Corresponding author at: School of Psychology, Faculty of Health, Deakin University, Burwood, VIC 3125, Australia. Tel.: +61 3 9246 8157; fax: +61 3 9244 6858.
E-mail address: susannah.tye@deakin.edu.au (S.J. Tye).

0149-7634/$ – see front matter © 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.neubiorev.2012.06.001
R.J. Anderson et al. / Neuroscience and Biobehavioral Reviews 36 (2012) 1920–1933 1921

4. Mechanisms of action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1926


4.1. Local mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1926
4.2. Regional mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1927
5. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1929
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1929

1. Introduction 2003), anterior capsulotomy (Christmas et al., 2011), subcaudate


tractotomy (Poynton et al., 1995) and limbic leucotomy (Montoya
Major depressive disorder (MDD) is one of the most preva- et al., 2002); however, the procedures are inherently irreversible
lent mental illnesses, with lifetime prevalence rates reported to and can result in, albeit rare and often transient, serious adverse
be between 3% and 17% and annual incidence between 5% and events (Dougherty and Rauch, 2007; Greenberg et al., 2003). Vari-
7% (Andrade et al., 2003; Andrews et al., 2001; Hasin et al., 2005; ous other somatic treatments, including magnetic seizure therapy
Kessler et al., 2005a,b). A high rate of mortality is associated with (Kosel et al., 2003; Lisanby et al., 2003), epidural cortical stimulation
MDD, with an estimated overall relative risk compared to non- (Kopell et al., 2011; Nahas et al., 2010) and deep brain stimulation
depressed individuals of 1.81 (Cuijpers and Smit, 2002). Higher (DBS) (e.g., Mayberg et al., 2005) are in the experimental phase, or
suicide rates (Mykletun et al., 2007), which vary depending on pre- are undergoing clinical trials for TRD.
vious hospitalisation status (outpatients 2.2%, inpatients 4.0–8.6%, DBS involves the stereotactic implantation of electrodes into
Bostwick and Pankratz, 2000) and severity of depression (moderate specific neuroanatomical structures where continuous stimula-
3.1%, severe 13.8%, Brådvik et al., 2008), and cardiovascular dis- tion is applied via a stimulator device implanted subcutaneously
ease (Carney and Freedland, 2003; Mykletun et al., 2007; Wulsin below the clavicle (Tye et al., 2009). DBS has benefits over
et al., 1999) make significant contributions to an increased mortal- other forms of interventional somatic treatments in that it is
ity and a consequent shortened lifespan (De Hert et al., 2011). MDD non-ablative and therefore reversible, targets small specific struc-
is one of the leading causes of disability worldwide and the cause tures, has both acute and long-term effects, and can be adjusted
of more years lost due to disability than any other disease (World to attain an optimal therapeutic effect (Kennedy and Giacobbe,
Health Organisation, 2008). Over the next 10 years the disease bur- 2007; Schlaepfer et al., 2011). DBS has been used successfully
den associated with MDD is projected to increase, with estimates for the treatment of symptoms associated with movement disor-
that MDD will represent the second leading cause of disease bur- ders (Johnson et al., 2008) including Parkinson’s disease (Benabid,
den worldwide by 2020 (Murray and Lopez, 1997a,b), following 2003; The Deep-brain Stimulation for Parkinson’s Disease Study
ischemic heart disease. Group, 2001), essential tremor (Hariz et al., 2008) and dysto-
Complete remission (i.e., asymptomatic) from an episode of nia (Vidailhet et al., 2005). DBS has also been used to control
major depression is the goal of treatment, as residual or subsyn- epilepsy (Hodaie et al., 2002) and provide relief from chronic pain
dromal symptoms are associated with higher rates of relapse and (Kringelbach et al., 2007). More recently DBS has been utilised
recurrence and an overall poorer long-term prognosis (American for treatment-refractory neuropsychiatric indications including
Psychiatric Association, 2000; Bakish, 2001; Fava, 2003; Sadock Tourette’s syndrome (Kuhn et al., 2007; Maciunas et al., 2007;
and Sadock, 2007). Multiple, pharmacological and psychothera- Servello et al., 2008), obsessive–compulsive disorder (Abelson et al.,
peutic interventions have proven efficacious in the treatment of 2005; Greenberg et al., 2006; Lipsman et al., 2007) with a recent
MDD; however, approximately 30% of individuals suffering from humanitarian device exemption (HDE) by the Food and Drug
MDD fail to respond to these treatments (Warden et al., 2007). Administration in the US, and a small number of clinical trials of
This subset of individuals is often referred to as suffering from DBS for TRD have been published (see Table 1). Despite the inher-
treatment-resistant depression (TRD). Although there is debate ent methodological limitations of these early trials of DBS for TRD
over its clinical application (Hauptman et al., 2008; Rush et al., a significant reduction in depressive symptomatology has been
2003) TRD is generally defined as an inadequate clinical response to demonstrated in individuals suffering from severely treatment-
antidepressant treatments (including augmentation interventions, refractory depression.
Dodd et al., 2005) that have been administered at an effective dose
1.1. Scope of this review
for a sufficient duration (Fava, 2003; Hauptman et al., 2008; Rush
et al., 2003). The purpose of this review is to assess the efficacy and safety
Failing pharmacotherapy and psychotherapeutic interventions, of DBS for TRD and to examine the putative mechanisms of action
electroconvulsive therapy (ECT) has been used to treat TRD with underlying its therapeutic effects. This review is confined to stud-
acute antidepressant effects (Rasmussen et al., 2002) and although ies where DBS was used for the purpose of treating primary TRD,
there is a large body of evidence demonstrating efficacy (The UK ECT therefore studies were excluded where remission or improvement
Review Group, 2003), memory disturbances and a high relapse rate in depressive symptoms was achieved during DBS treatment for
(Rasmussen et al., 2002) limit its use (Kennedy and Giacobbe, 2007). other disorders (e.g., primary treatment-resistant OCD, Aouizerate
Repetitive transcranial magnetic stimulation (rTMS) has regulatory et al., 2004, 2009; and treatment of tardive diskinesia in a patient
approval for treatment of major depression that has failed to opti- with severe TRD, Kosel et al., 2007). The excluded studies however
mally respond to one prior antidepressant treatment for a current are important in improving our understanding of the neurobiol-
episode of major depression (O’Reardon et al., 2007), but its efficacy ogy of depression and treatment resistance, and also in guiding
is not comparable to ECT (Slotema et al., 2010). The maintenance selection of new DBS targets for TRD.
of response and response rates with greater treatment resistance
have yet to be investigated. Vagal nerve stimulation (VNS) also has 2. Target selection
regulatory approval for treatment of TRD (Rush et al., 2005). Opti-
mal dosing strategies and clinical predictors of response have not It could be argued that DBS, by targeting single neuroanatom-
been fully developed to optimise this treatment intervention (Rush ical structures, is an overly reductionistic approach to treating a
and Siefert, 2009). Ablative neurosurgery has also been performed psychological disorder that has been shown at the biological level
for severe TRD, including anterior cingulotomy (Dougherty et al., to involve various neuroanatomical structures, neurotransmitters,
1922
Table 1
Deep brain stimulation for treatment-resistant depression: reviewed studies.

Target and study Number of Follow-up Sex ratio Mean age Mean age Patient responsesa Primary Stimulation parametersb
(extended/follow-up study) patients period in m/f at surgery MDD onset measure
months (SD) (SD)

R.J. Anderson et al. / Neuroscience and Biobehavioral Reviews 36 (2012) 1920–1933


Response Remission V Hz PW (␮s)

Subgenual cingulate gyrus


Mayberg et al. (2005) 6 6 3/3 47.0 (8) 29.5 (12) 66.6% 33.3% HDRS17 4.0 130 60
(Lozano et al. (2008)) 20 12 9/11 47.3 (10.4) 27.1 (8.3) 55.0% 35.0%c HDRS17 3.5–5.0 130 90
(Kennedy et al. (2011)) 20 39d 9/11 47.3 (10.4) 27.1 (8.3) 64.3% 42.9% HDRS17 4.3 124.7 70.6
Neimat et al. (2008) 1e 30 0/1 55 9 – –* HDRS17 4.5 130 60
Puigdemont et al. (2012) 8 12 2/6 47.4 (11.3) 24.9 (5.3) 62.5% 50.0% HDRS17 3.5–5.0 135 120–210
Holtzheimer et al. (2012) 10 24 3/7 40.0 (9.3) 20.3 (5.6) 92.0%h 58.0%f HDRS17 4–8 mA 130 91
Ventral capsule/ventral striatum
Malone et al. (2009) 15 23g 4/11 46.3 (10.8) 25.3 (10.5) 53.3% 40.0% HDRS24 6.7 127 113
(Malone (2010)) 17 37h NR 46.3 25.3 71.0% 35.0% MADRS 2.5–8.0 100–130 NR
Nucleus accumbens
Schlaepfer et al. (2008) 3 6–23 weeks 2/1 46.7 (13.7) NR 33.3% – HDRS24 4.0 145 90
(Bewernick et al. (2010)) 10 12–36 6/4 48.6 (11.7) 31.7 (13.2) 50.0%i 30.0%i HDRS28 1.5–10.0 100–150 60–210
(Bewernick et al. (2012)) 11 12–48 7/4 48.4 (11.1) 32.6 (12.4) 45.5% 9.0% HDRS28 1.5–10.0 100–150 60–210
Inferior thalamic peduncle
Jiménez et al. (2005) 1j 24 0/1 49 29 – –* HAM-D 2.0–2.5 130 450
Lateral habenula
*
Sartorius et al. (2010) 1 12 0/1 64 18 – – HAMD21 10.5 165 60

Abbreviations: SD, standard deviation; V, volts; mA, milliamps; Hz, hertz; PW, pulse-width; NR, not reported; MDD, major depressive disorder; BD, bipolar disorder.
a
As at last follow-up, unless indicated.
b
Mean parameters for entire group during the chronic phase of stimulation.
c
Remission or within one point of remission.
d
Mean follow-up period, range = 36–72 months.
e
Previous cingulotomy.
f
Calculation includes 8 of the original 10 MDD patients and 3 BD patients.
g
Mean follow-up period, range = 6–51 months.
h
Mean follow-up period, range = 14–67 months.
i
Assessed after 12 months post-surgery.
j
Comorbid borderline personality disorder and bulimia.
*
Case studies: however all patients achieved remission.
R.J. Anderson et al. / Neuroscience and Biobehavioral Reviews 36 (2012) 1920–1933 1923

and genes (e.g., Sadock and Sadock, 2007). A host of psychological, cerebral blood flow (CBF), in depressed patients compared with
social, personality, lifestyle and environmental factors are thought healthy controls (Mayberg, 2003; Seminowicz et al., 2004) and
to contribute to the aetiology of major depressive disorder (Heim that hyperactivity in the SCG is correlated with depression symp-
and Nemeroff, 2001; Jacka et al., 2011a,b; Williams et al., 2011). tom severity (Kennedy et al., 2007). Increased activity was also
Indeed, complex interactions between biological and psychosocial observed in the SCG during activities designed to induce sadness
factors have been implicated in the aetiology of depression (e.g., in healthy volunteers (Mayberg et al., 1999). Conversely, there is
Kendler et al., 2001). The models of depression driving DBS target abundant evidence that SCG activity decreases following success-
selection consider the target neuroanatomical structures as “nodes” ful treatment of depression with pharmacotherapy (e.g., sertraline,
within dysfunctional neural circuits that modulate different aspects Drevets et al., 2002; venlafaxine, Kennedy et al., 2007; fluoxetine,
of the depressive syndrome (Mayberg, 2003, 2006; Seminowicz Mayberg et al., 2000), somatic treatments (e.g., anterior cingulo-
et al., 2004) via connections to limbic, cortical, and subcortical tomy, Dougherty et al., 2003; rTMS, Mottaghy et al., 2002; ECT,
areas (Giacobbe et al., 2009). Accumulating evidence from neu- Nobler et al., 2001) and even in placebo responders (Mayberg et al.,
roimaging, neuropathological, lesion analysis, and post-mortem 2002). Importantly it appears that functional changes in the SCG can
studies implicates specific brain networks in the pathophysiol- differentiate response to treatment versus nonresponse (Kennedy
ogy of mood disorders (Drevets et al., 2008). The dysfunction in et al., 2007; Mayberg et al., 2000) and that hyperactivity in the SCG
these neural circuits is thought to involve a failure of intrinsic com- distinguishes more treatment resistant patients (Kennedy et al.,
pensatory measures under excessive allostatic load (Kapczinski 2007; Seminowicz et al., 2004).
et al., 2008; McEwen, 2003) that maintain stable emotional states
under stress (Mayberg, 2003; McEwen, 2003), with variations 2.2. Ventral capsule/ventral striatum
in these compensatory measures thought to involve interac-
tions between biological vulnerability, environmental stressors The ventral capsule/ventral striatum (VC/VS) was selected as
and social factors (Caspi et al., 2003; Heim and Nemeroff, 2001; a DBS target after reports of elevated mood and reduced depres-
Kendler et al., 2001, 2005a,b). This de-compensation, mediated by sive symptom severity following ablative procedures for OCD and
inflammatory, and oxidative pathways, and the balance between MDD (e.g., Greenberg et al., 2003; Rück et al., 2008), and more
apoptosis and neurogenesis appears to underpin the process of recently after DBS for refractory OCD with comorbid MDD (e.g.,
neuroprogression in the disorder (Berk et al., 2011; Maes et al., Abelson et al., 2005; Aouizerate et al., 2004, 2009; Gabriels et al.,
2011). Independent and interactive effects of these factors are 2003; Greenberg et al., 2006; Van Laere et al., 2006). Further sup-
thought to underlie the heterogeneity in depressive symptomatol- porting the VC/VS as a DBS target for TRD are the findings from an
ogy (Mayberg, 2003, 2006, 2009; Seminowicz et al., 2004). These fMRI study (Epstein et al., 2006) that demonstrated that patients
networks include the limbic–cortical–striatal–pallidal–thalamic suffering from major depressive disorder showed significantly less
network, formed by connections between the orbital and medial activation bilaterally in the ventral striatum to positive stimuli
prefrontal cortex, amygdala, hippocampal subiculum, ventrome- compared to healthy controls.
dial striatum, mediodorsal and midline thalamic nuclei and ventral
pallidum (Ongür et al., 2003). 2.3. Nucleus accumbens
The orbital and medial prefrontal cortex is associated with two
extended networks, the orbital prefrontal network and the medial The nucleus accumbens (NAc), located where the head of the
prefrontal network (Drevets et al., 2008). The orbital prefrontal caudate and the putamen meet, was chosen as a DBS target pri-
network codes for affective characteristics of stimuli, including marily on the basis of its role in the anticipation and experience of
reward, aversion, and relative value (Drevets et al., 2008; Ongür pleasure and reward (Heien and Wightman, 2006; Knutson et al.,
and Price, 2000; Saleem et al., 2008). The medial prefrontal network 2001; Nestler and Carlezon, 2006; Schlaepfer et al., 2008) and goal-
is involved in introspective functions such as mood and emotion, directed behaviour (Goto and Grace, 2005), with dysfunction in the
and visceral reactions to emotional stimuli. The medial prefrontal NAc thought to play a key role in the anhedonic symptoms of MDD
network includes the dorsomedial/dorsal anterolateral prefrontal (Gorwood, 2008; Nestler and Carlezon, 2006; Tremblay et al., 2005).
cortex (e.g., BA 9), the mid- and posterior cingulate cortex, a region The dopaminergic innervation of the NAc in particular plays a key
in the anterior superior temporal gyrus and sulcus, and the entorhi- role in mood regulation and the processing of reward (Arias-Carrión
nal and posterior parahippocampal cortex (Drevets et al., 2008; et al., 2010; Malhi and Berk, 2007).
Kondo et al., 2005; Saleem et al., 2008). Elevated metabolism in
many structures within this circuitry has been reported during the 2.4. Inferior thalamic peduncle
depressive phase (Drevets et al., 2002, 2008) of MDD. Conversely, in
recovered cases that experience depressive relapse under experi- The inferior thalamic peduncle (ITP), a bundle of fibres connect-
mental conditions involving serotonin or catecholamine depletion, ing the dorsomedial thalamus to the orbital frontal cortex (Kopell
the metabolic activity increases in these regions as the depres- and Greenberg, 2008), was selected based on positron emission
sive symptoms return (Hasler et al., 2008; Neumeister et al., 2004). tomography (PET) studies showing hyperactivity in the medial
Deep brain stimulation, as well as psychotherapy, has been shown thalamus and the orbitofrontal cortex in depressed patients (e.g.,
through neuroimaging studies to normalise the dysfunction in Drevets, 2000). Surgical lesions in an area including the ITP disrupt
these neural circuits following successful treatment (Mayberg, the inhibitory action of the thalamo-orbitofrontal system, and have
2003, 2009; Mayberg et al., 2005; Ressler and Mayberg, 2007). Lev- significant antidepressant effects in preclinical models (Velasco
els of metabolic activity in these pathways additionally appear to et al., 2005).
be germane to the pathophysiology of depression (Hamani et al.,
2011). 2.5. Lateral habenula

2.1. Subgenual cingulate gyrus The lateral habenula (LHb), located on the dorsomedial surface
of the caudal thalamus, was chosen as a DBS target for TRD on the
The subgenual cingulate gyrus (SCG), a region of the cingulate basis of hypermetabolism observed in individuals suffering from
gyrus located ventral to the genu of the corpus callosum (Fig. 1), depression (Morris et al., 1999; Roiser et al., 2009) and in ani-
has shown hyperactivity, as measured by glucose metabolism and mal models of depression (Caldecott-Hazard et al., 1988; Shumake
1924 R.J. Anderson et al. / Neuroscience and Biobehavioral Reviews 36 (2012) 1920–1933

Fig. 1. Deep brain stimulation targets for treatment-resistant depression. Location of DBS targets used for TRD and the published studies associated with them in parentheses.

and Gonzalez-Lima, 2003). Recent research has demonstrated a the SCG. All patients reported acute positive effects when electrical
reduction in depressive-like behaviours in animal models following stimulation was initiated (e.g., Lozano et al., 2008) coinciding with
lesions (Yang et al., 2008), pharmacological inhibition (Winter et al., changes in positive and negative affect scores (PANAS, Watson and
2011) and deep brain stimulation (Li et al., 2011) of the LHb. Fur- Clark, 1988). During the follow-up period, a progressive reduction
ther, research in monkeys has shown that LHb neurons are excited in the severity of depression symptoms was reported, with only 10%
on the presentation of a punishment and the absence of a reward, of patients failing to show a decrease in symptom scores. Signifi-
and code for a prediction error for punishments (Hikosaka, 2010; cantly, no patients experienced a worsening of symptoms relative
Matsumoto and Hikosaka, 2007, 2008), thus possibly having a role to baseline (Lozano et al., 2008). Different subgroups of symptoms
in the motivational aspects of the depressive syndrome. improved at different rates, with mood symptoms reaching max-
imum levels of improvement, and then plateauing three months
3. Efficacy post-surgery (Lozano et al., 2008). Alternatively, somatic, sleep and
anxiety symptoms steadily improved throughout the 12-month
Concordant with custom in antidepressant trials, all DBS for period (Lozano et al., 2008). At final follow-up (3–6 years) 42.9%
TRD studies reviewed defined a clinically significant response as of patients were in remission (Kennedy et al., 2011).
a 50% reduction in the severity of depressive symptoms from Recently a second independent trial of DBS of the SCG was pub-
pre-operative baseline scores, as measured on a standardised, lished (Puigdemont et al., 2012) showing similar rates of response
clinician-rated instrument. Remission however is the ultimate goal, and remission (see Table 1). After an initial sharp decline in depres-
and best correlates with functional recovery (Berk et al., 2008). sive symptoms in the first week of stimulation where 7 of the
The primary instrument in most studies was the Hamilton Depres- 8 patients reached the response criteria and 4 patients achieved
sion Rating Scale (HDRS; Hamilton, 1960). The Montgomery-Asberg remission, an increase in depressive symptoms was observed,
Depression Scale, (MADRS: Montgomery and Asberg, 1979) was followed by a progressive decrease in HDRS scores until final
also frequently utilised. Measures of psychosocial functioning (e.g., follow-up (i.e., one year) (Puigdemont et al., 2012). A significant
Global Assessment of Function Scale (GAF)) and self-reported decrease in all subscales of the HDRS17 , that is mood, anxiety,
depression symptom severity measures (e.g., Beck Depression somatic and sleep subscales, was reported as well as overall scores
Inventory) were also commonly employed. Although inclusion (Puigdemont et al., 2012). Unlike previous attempts to explore the
criteria varied, patients in each study exhibited high degrees of relationship between therapeutic response and electrode location
treatment resistance including long histories of MDD, a current (Hamani et al., 2009) Puigdemont et al. (2012) found a significant
major depressive episode (MDE), numerous courses of antidepres- relationship between long-term response and the location of elec-
sants of different classes, augmentation/combination therapies, trodes, with responder patients having electrodes in BA24 of the
failed psychotherapy and ECT. SCG rather than BA25.
Further evidence for the antidepressant efficacy and safety of
3.1. Subgenual cingulate gyrus DBS for TRD has been provided in a subsequent independent study
of 17 patients (10 MDD; 7 BP) (Holtzheimer et al., 2012). Following
The efficacy of DBS of the SCG has been demonstrated in three a four-week single blind sham stimulation condition, all patients
independent studies (see Table 1). The first published was an ini- received open-label active stimulation for 24 weeks, followed
tial six-month study of six patients (Mayberg et al., 2005) and by a further single blind sham stimulation phase. A significant
a 12-month extension, which included an additional 14 patients reduction in depression as measured by the HDRS, Beck Depression
(Lozano et al., 2008). Recently a 3–6 year follow-up of these 20 Inventory II and an increase in function according to the Global
patients was published (Kennedy et al., 2011). Patients had elec- Assessment of Functioning (GAF) was achieved following active
trodes implanted bilaterally into white matter tracts adjacent to stimulation. An observational follow-up after two years of chronic
R.J. Anderson et al. / Neuroscience and Biobehavioral Reviews 36 (2012) 1920–1933 1925

stimulation revealed high response and remission rates (92% 3.4. Inferior thalamic peduncle
and 58%, respectively), with no recurrence of moderate or severe
episodes of depression at this time. A single patient with a 20-year history of MDD, with comor-
Further supporting the efficacy of the SCG as a DBS target for bid borderline personality disorder and bulimia, had electrodes
TRD is a case study of a patient who had undergone a cingulotomy implanted bilaterally in and around the inferior thalamic pedun-
for severe MDD, from which remission was achieved for a period cle (Jiménez et al., 2005). This study differed from the other DBS
of six months before a relapse (Neimat et al., 2008). With contin- studies in that the patient was hospitalised and all antidepressant
uous DBS of the SCG the patient achieved remission, which was medication withdrawn two months prior to surgery. Despite an
sustained throughout the 30-month duration of the study. Even initial sharp decline in depression symptom scores to remission-
though remission was significantly longer than any other treat- levels following implantation of the electrodes, after one week the
ment the patient had undergone previously, and despite being a patient experienced a relapse. The patient then attained remission
case study, the fact that the patient had undergone a cingulotomy, coincident with the initiation of stimulation (Jiménez et al., 2005).
which ablates structures connected to the SCG, makes it difficult to After eight months of stimulation, the electrodes were switched off
generalise results and efficacy of the DBS alone in this case. (double-blind) with a delay of two months before any decline was
seen in HDRS scores. When stimulation was reinstated, HDRS scores
3.2. Ventral capsule/ventral striatum returned to remission levels. As two months had elapsed before any
fluctuation in depression scores was observed and the patient did
In a multicentre trial, 17 patients who received DBS for TRD, with not suffer a MDE within 10 months following cessation of stimu-
electrodes implanted bilaterally into the VC/VS region, showed lation calls into question interpretations regarding efficacy of DBS
significant improvement in depressive symptomatology (Malone, for this patient, although relapse typically follows some months
2010; Malone et al., 2009). The greatest reduction in mean depres- after discontinuation of effective pharmacotherapy for most disor-
sion symptom severity (i.e., ∼50%) was achieved after 3 months ders. Other DBS studies for TRD have reported significant relapse of
of continuous stimulation, which was sustained at the 12-month depressive symptoms within much shorter periods of time during
period (Malone et al., 2009). Significantly, all 17 patients’ saw a blinded discontinuation studies (e.g., Schlaepfer et al., 2008).
reduction in depression rating scale scores from baseline to final
follow-up (variable follow-up periods). Additionally, a consistent 3.5. Lateral habenula
and significant improvement in GAF scores was observed through-
out the study. Suicidalilty, as measured by the MADRS suicidal A single patient with a 46-year history of MDD and demonstrat-
thoughts subscale (Montgomery and Asberg, 1979), decreased sig- ing a high level of treatment resistance over the 9 years prior to
nificantly following one month of stimulation and was sustained being considered for surgery received bilateral DBS in the stria
at the 12 month follow-up period (Malone, 2010). At final follow- medullaris, the major afferent connection to the LHb (Sartorius
up (14–67 months) 35% of the patients had achieved remission et al., 2010). The patient achieved remission after 4 months of stim-
(Malone, 2010). ulation, which was sustained at final follow-up (at ∼12 months).
A dramatic relapse of depressive symptoms was observed after an
3.3. Nucleus accumbens erroneous cessation of stimulation following treatment for a bicycle
accident; however, 12 weeks after the reinstatement of stimulation
Ten patients suffering from severe TRD had electrodes the patient again achieved remission (Sartorius et al., 2010).
implanted bilaterally into the shell and core of the NAc, and the
ventral and medial internal capsule, however only the NAc targets 3.6. Safety and adverse events
were stimulated (Bewernick et al., 2010; Schlaepfer et al., 2008).
The effects of one week of DBS in three of the patients was pub- Serious adverse events (e.g., suicide) were rare in the studies
lished first (Schlaepfer et al., 2008), with stimulation manipulated reviewed and were deemed to be unrelated to the stimulation or
in a double-blind manner (switching the voltage off and on), associ- device failure (Bewernick et al., 2010; Holtzheimer et al., 2012;
ated with a consistent and significant drop in depression symptom Kennedy et al., 2011), rather, compliance, adverse psychosocial
severity when the stimulator was on and an increase when off. events (Bewernick et al., 2010; Holtzheimer et al., 2012) and a
All three patients demonstrated decreases in symptom severity family history of completed suicide (Kennedy et al., 2011) were
during the first week of stimulation, however only one patient considered important factors. Adverse events relating to the sur-
had achieved a clinical response (Schlaepfer et al., 2008). After 12 gical procedure (e.g., localised pain, local wound infection) were
months of continuous stimulation five out of the ten patients had also reported (e.g., Holtzheimer et al., 2012; Lozano et al., 2008;
achieved a clinical response with three patients in remission for a Malone et al., 2009; Mayberg et al., 2005; Puigdemont et al., 2012;
one month period (Bewernick et al., 2010). Patients demonstrated Schlaepfer et al., 2008). In one study, electrodes were explanted
a decrease in anhedonia, as measured by a significant increase in in three patients due to infections (Lozano et al., 2008; Mayberg
activities on a list of 280 pleasant activities (Bewernick et al., 2010) et al., 2005) without a worsening of depressive symptoms. Despite
after 12 months of stimulation. Demonstrating the long-term sta- an unknown aetiology of the infections Mayberg et al. (2005)
bility of DBS of the NAc for TRD, patients that were considered changed their protocol to have the electrodes and the pulse gen-
responders at 12 months continued to show a clinical response at a erator implanted in one surgical session; following this protocol,
2 year follow-up and a final follow-up of up to 4 years (Bewernick a recent study (Puigdemont et al., 2012) reported no post-surgical
et al., 2012) without a worsening of depression symptomatology. infections.
As a group both responders and non-responders demonstrated a Acute adverse events related to the stimulation (e.g., effect;
significant drop in depression symptom severity when compared DBS target; sweating; VC/VS, anxiety; ITP, NAc, VC/VS, light-
with baseline measurements with non-responders experiencing a headedness; SCG, tachycardia; ITP, VC/VS, paresthesias; VC/VS,
sustained drop in depression symptom severity of greater than 10% NAc), generally occurring during a parameter change (e.g.,
at both 12 and 24 month follow-up periods which was sustained at Bewernick et al., 2010, 2012) or titration phase of stimulation (e.g.,
final follow-up (Bewernick et al., 2012). Significant improvements Jiménez et al., 2005; Lozano et al., 2008), were transient (Bewernick
in anxiety (as measured by the HAMA) and hedonic activities as a et al., 2010) or were ameliorated by an additional change in stimu-
group were also reported (Bewernick et al., 2012). lation parameters (e.g., Bewernick et al., 2010; Jiménez et al., 2005;
1926 R.J. Anderson et al. / Neuroscience and Biobehavioral Reviews 36 (2012) 1920–1933

Lozano et al., 2008; Malone, 2010; Malone et al., 2009; Mayberg 2008) with depression scores increasing during the off period and
et al., 2005). Hypomania was observed in one study targeting the decreasing again when stimulation was reinstated. Even acciden-
VC/VS (Malone et al., 2009), however these patients had suffered tal cessation of stimulation or battery depletion, of which patients
manic episodes prior to DBS surgery. Weak, subsyndromal symp- were unaware, resulted in an increase in depressive symptoms
toms of hypomania were experienced during a parameter change (e.g., Lozano et al., 2008; Malone et al., 2009; Mayberg et al., 2005;
by two patients receiving DBS of the NAc (Bewernick et al., 2012) Sartorius et al., 2010).
which resolved within 24 h. A relative increase in depression symp- Antidepressant medication may have presented a confounding
tom severity, although still below preoperative levels, was also factor in the interpretation of the therapeutic effects associated
reported and was most often due to battery depletion or accidental with DBS for TRD. A majority of the studies made an attempt to
cessation of stimulation (Lozano et al., 2008; Malone et al., 2009; control for medication effects by holding medications and dosage
Mayberg et al., 2005; Sartorius et al., 2010). constant throughout the study period (e.g., Lozano et al., 2008;
Results of neuropsychological testing revealed no negative Malone et al., 2009; Mayberg et al., 2005; Schlaepfer et al., 2008)
effects on cognitive function during DBS (Bewernick et al., 2010, however, this does not rule out the possibility of synergistic effects
2012; Grubert et al., 2011; Holtzheimer et al., 2012; Jiménez between DBS and antidepressant medications, even though fail-
et al., 2005; Lozano et al., 2008; Malone et al., 2009; Mayberg ure of response to pharmacotherapy is a requirement for DBS.
et al., 2005; Puigdemont et al., 2012), in fact, improvements were Interactions between DBS for Parkinson’s disease and medica-
reported in cognitive domains that were below average prior to tions (i.e., dopaminergic drugs) have been shown to occur (e.g.,
DBS (Bewernick et al., 2012; Grubert et al., 2011; Holtzheimer The Deep-brain Stimulation for Parkinson’s Disease Study Group,
et al., 2012; Jiménez et al., 2005; Malone et al., 2009; McNeely 2001). Only one study removed all antidepressant medication
et al., 2008). Interestingly the improvement in cognitive function prior to surgery (i.e., Jiménez et al., 2005), and although remission
was not associated with the improvement in mood or stimulation was achieved, it was a case study making it difficult to gener-
parameter changes (Grubert et al., 2011; McNeely et al., 2008). Neu- alise the results. Despite these limitations, DBS has demonstrated
ropsychological assessment of one group of patients (i.e., Lozano dramatic, sustained reductions in depression symptoms and signif-
et al., 2008; Mayberg et al., 2005) revealed a transient mild motor icant improvement in functioning in a large percentage of severely
slowing which had normalised by final follow-up (McNeely et al., treatment resistant patients. The mechanisms of action underlying
2008). the therapeutic effects however are still incompletely understood.

3.7. Efficacy evaluation


4. Mechanisms of action
The open-label design and uncontrolled nature of the DBS stud-
The mechanisms of action of DBS are still poorly understood
ies for TRD makes it difficult to rule out the possible influence
despite clinical efficacy across a number of disorders. The results of
of placebo and non-specific factors on the therapeutic outcomes.
preclinical research have increased our knowledge of the mecha-
Spontaneous remission is also a possibility that needs to be consid-
nisms of action of DBS for TRD. By stimulating or pharmacologically
ered, although this declines progressively with both chronicity and
activating/deactivating homologous regions in the brains of exper-
number of failed trials of therapy (Trivedi et al., 2006). The small
imental animals antidepressant-like effects have been observed
sample sizes limit the ability to generalise the results, to evaluate
(Falowski et al., 2011; Hamani et al., 2010a,b; Li et al., 2011;
the differences between responders and non-responders, and to
Scopinho et al., 2010; Slattery et al., 2011; Winter et al., 2011;
determine the superiority of one target over another. Without large,
Yang et al., 2008) utilising a number of behavioural assays. Any
randomised, placebo-controlled studies these possibilities cannot
description of the mechanisms of action of DBS for TRD will need
be ruled out, however, a number of factors can be used to argue
to be able to explain the behavioural effects of acute stimulation,
against the strong influence of a placebo effect and in favour of
the progressive improvement in depressive symptomatology, and
electrical stimulation as a major factor in the therapeutic outcomes.
CBF/metabolic changes in neuroanatomical structures distal to the
Even though some patients saw a rapid decrease in depres-
site of stimulation following months of continuous stimulation. To
sion symptomatology within weeks after initial stimulation (e.g.,
this end a number of mechanisms have been hypothesised that
Mayberg et al., 2005; Puigdemont et al., 2012; Schlaepfer et al.,
explain the effects of stimulation at the neuronal level through to
2008), typical of a placebo response, the gradual and sustained
network-wide effects.
improvement over the follow-up periods reported for all respon-
ders is inconsistent with a placebo response for MDD (Quitkin et al.,
1993). Another argument against a strong influence of a placebo 4.1. Local mechanisms
effect is the extreme level of treatment-resistance of all patients
within the studies reviewed. TRD patients have previously been A number of variables need to be taken into account when
shown to have lower placebo response rates (e.g., Rush and Thase, determining the activity at the site of stimulation, for example,
1997; Thase and Rush, 1997). Similarly, with increasing levels of the volume of neural tissue activated depends on the stimu-
treatment-resistance and longer current episodes there is a reduced lation parameters, that is voltage, pulse-width and frequency
chance of spontaneous remission (Rush and Thase, 1997; Rush et al., (Butson et al., 2007), the type of neural tissue stimulated, whether
2003). white matter or grey matter, myelinated or unmyelinated axons
Additional evidence in support of stimulation as a major factor (Perlmutter and Mink, 2006) and the orientation of the neural ele-
in the therapeutic outcomes of the DBS studies can be observed ment with respect to the electric field (McIntyre et al., 2004a). With
during the acute or titration phase of stimulation. In all stud- the electric field decaying exponentially from the electrode contact
ies reviewed, patients’ responses to stimulation were specific (Kuhn et al., 2009) distance is another factor that will determine
to certain electrode contacts and parameters, were stimulation- whether a neural element is stimulated.
locked and reproducible even though patients were blind to the Early research in the area of DBS for movement disorders sug-
stimulation parameters. Additionally, during the chronic phase gested a mechanism of action similar to that of a surgical lesion,
of stimulation, discontinuation trials were conducted single blind given similar effects were observed for both (Benabid, 2003) and
(Holtzheimer et al., 2012; Lozano et al., 2008; Mayberg et al., despite the logical fallacy (Montgomery and Gale, 2008) provided
2005) and double blinded (Jiménez et al., 2005; Schlaepfer et al., a good working hypothesis. Blockade of voltage-gated currents
R.J. Anderson et al. / Neuroscience and Biobehavioral Reviews 36 (2012) 1920–1933 1927

(Beurrier et al., 2001), membrane hyperpolarisation (Dostrovsky targets used in DBS for TRD have connections to cortical, subcorti-
and Lozano, 2002), neurotransmitter depletion (Anderson et al., cal, limbic and brainstem regions that show profound changes in
2006; Iremonger et al., 2006) and synaptic inhibition (via release of CBF/metabolism following chronic stimulation; regions which are
inhibitory neurotransmitters) (Dostrovsky et al., 2000) are mech- linked to depressive symptomatology and their amelioration.
anisms proposed to explain the lesion-like neuronal inhibition. With connections to prefrontal areas the SCG, VC/VS, NAc,
However, excitation and an increase in activation in the local area of ITP and the LHb (Fig. 2a) may be involved in the modulation of
stimulation have also been demonstrated (Hashimoto et al., 2003; the cognitive and mood symptoms of depression following DBS,
Maurice et al., 2003). Reconciling these findings, McIntyre et al. with medial, orbital and dorsolateral prefrontal areas implicated
(2004b) showed that both suppression and activation of neuronal in depressive symptomatology and recovery following successful
activity during high frequency stimulation was possible, with the treatment (Drevets, 2007; Drevets et al., 2002; Liotti and Mayberg,
decoupling of soma and axon activation, in which simultaneous 2001; Ongür and Price, 2000; Price and Drevets, 2010). FDG-PET
inhibition of the cell body and excitation of the axon can occur. imaging acquired in seven of the ten patients receiving continuous
Although, these local mechanisms of action following stimulation NAc DBS for 6 months (Bewernick et al., 2010) showed decreases
may contribute to our understanding of the therapeutic effects in metabolism, from baseline, in the orbitofrontal cortex and dor-
of DBS, they do not directly explain the gradual and differential sal medial frontal gyrus. Similarly, significant decreases in glucose
improvements observed in depressive symptomatology recorded metabolism in the medial and orbital frontal areas were reported
over months of continuous stimulation. In line with this, local neu- following 3 and 6 months of continuous DBS of the SCG (Lozano
ral activity has been shown to return to baseline within seconds et al., 2008). Decreases in CBF were observed in medial and orbital
after cessation of acute stimulation (Johnson et al., 2008). frontal cortex and increases observed in dorsolateral prefrontal
Gradual and sustained improvements in depressive symp- regions (Mayberg et al., 2005). The increase in CBF in the dorsolat-
tomatology imply long-term changes in neuronal function with eral prefrontal cortex normalised baseline abnormalities relative
continuous DBS. Support for long-term neuronal changes can be to healthy controls (Mayberg et al., 2005). Significant increases in
seen in the blinded discontinuation studies where most patients glucose metabolism were also observed in the dorsolateral and ven-
showed a level of sustained improvement in depressive symp- trolateral prefrontal areas, becoming significant after 6 months of
tomatology. Moreover, when some symptoms did return, with SCG stimulation (Lozano et al., 2008).
cessation of stimulation, they were well below pre-surgical lev- Preclinical research has supported these findings that stimu-
els (e.g., Jiménez et al., 2005; Mayberg et al., 2005). Synaptic lation at DBS for TRD targets has the ability to alter activity in
plasticity (i.e., long term potentiation and long term depres- regions of the frontal cortex involved in the symptomatology of
sion) has been suggested as one of the mechanisms producing MDD. Neuroplasticity, demonstrated by morphological changes to
long-term neural changes (Lujan et al., 2008), possibly reversing dendrites in pyramidal neurons, and changes to dopamine and
disrupted neuroplasticity implicated in the neurobiology of depres- norepinephrine concentrations have been observed in the pre-
sion (Duman, 2002; Pittenger and Duman, 2008). In support of this, frontal cortex of rats following DBS of the NAc (Falowski et al.,
high-frequency stimulation has been shown, in animal research, 2011). DBS of the NAc in rats has also demonstrated changes in the
to induce synaptic plasticity (e.g., Shen et al., 2003). Secondary power of oscillations in local field potentials in the orbitofrontal
messengers, long-term potentiation and long-term depression, cortex and medial prefrontal cortex (McCracken and Grace, 2009).
synthesis of proteins, changes in gene expression and increased The hypothalamus is involved in the regulation of sleep,
energy generation have all been suggested as mechanisms of long- appetite and circadian rhythms, and is a key structure in the
term change due to DBS (Kalbe et al., 2009; Salin et al., 2002; Schulte hypothalamic–pituitary–adrenal axis, all considered dysregulated
et al., 2006; Temperli et al., 2003). Although explaining the long- in the depressive syndrome (Nestler et al., 2002; Swaab et al., 2005).
term effects of stimulation at the neuron level, a mechanism of The SCG, VC/VS, NAc, and LHb DBS targets have direct connections
action also needs to explain the extensive effects on distal brain to the hypothalamus (Fig. 2a). A decrease in CBF from baseline was
regions during chronic stimulation, thus a neural network expla- observed in the hypothalamus of long-term responders at 3 and 6
nation is required. months of SCG DBS (Lozano et al., 2008; Mayberg et al., 2005). The
amygdala, another subcortical structure with direct connections to
4.2. Regional mechanisms the DBS for TRD targets, has demonstrated increased metabolism
and CBF in depressed individuals compared with healthy controls
The hypothesis that stimulus-evoked modulation of neuronal (Abercrombie et al., 1998; Drevets, 1999; Drevets et al., 1992; Price
network activity is responsible for the therapeutic effects of DBS and Drevets, 2010) which correlates with depressive symptom
(Lujan et al., 2008; McIntyre et al., 2004a) incorporates both the severity (Drevets, 1999, 2003) and has been shown to normalise fol-
short and long-term effects of stimulation at the neuron level and lowing successful antidepressant treatment (Drevets et al., 2002).
its impact via connections within networks thought to underlie Despite an initial increase in metabolism in the amygdala after
depression symptomatology (see Fig. 2). At the neuron level, a con- one week of NAc DBS, possibly due to acute effects arising from
sistent finding of the effects of high frequency stimulation is the implantation of the electrode (Schlaepfer et al., 2008), significant
activation of structures immediately downstream (Anderson et al., metabolic decreases in the amygdala of responders compared to
2003; Hashimoto et al., 2003; Montgomery, 2006) and upstream non-responders were observed in a volume of interest analysis
(Hashimoto et al., 2003; Maurice et al., 2003; McCracken and Grace, of FDG-PET data in a group of patients at a 6-month follow-up
2007; Montgomery, 2006) from the site of stimulation via ortho- (Bewernick et al., 2010).
dromic and antidromic activation respectively (Dostrovsky and With direct and indirect connections to brain stem areas includ-
Lozano, 2002; Grill et al., 2008; Hammond et al., 2008). Preclinical ing the locus coeruleus, a major norepinephrine producing region,
research has shown that stimulation of DBS targets can increase the dorsal raphe nuclei, a serotonergic producing area and the
neurotransmitter release (Hamani et al., 2010b), induce neuro- dopaminergic ventral tegmental area (Fig. 2a), the DBS for TRD
plastic changes (Falowski et al., 2011), and alter the power and targets have the potential to have a significant impact on these
synchronisation of local field potential activity (McCracken and monoamines that have been shown to play a significant role in
Grace, 2007, 2009) in structures distal to the source of stimulation. depression and its treatment (Krishnan and Nestler, 2008; Ressler
Data obtained from diffusion tensor imaging studies in humans, and Nemeroff, 2000; Ruhé et al., 2007). Preclinical research has
tracer analysis in nonhuman primates and rodents have shown that shown that stimulation of the LHb increases serotonin (Kalén
1928 R.J. Anderson et al. / Neuroscience and Biobehavioral Reviews 36 (2012) 1920–1933

Fig. 2. Regional mechanisms. Deep brain stimulation facilitates profound changes in areas distal the site of stimulation through connections to neuroanatomical regions
that are involved in depressive symptomatology. (a) Highly simplified, schematic representation of DBS targets and their connectivity to regions involved in depressive
symptomatology (connections from tract tracing experiments in humans and tracing studies in animals (LHb (Geisler and Trimble, 2008; Geisler and Zahm, 2005; Herkenham
and Nauta, 1977, 1979; Hikosaka et al., 2008); SCG (Chiba et al., 2001; Freedman et al., 2000; Gutman et al., 2009; Hamani et al., 2011; Johansen-Berg et al., 2008; Takagishi
and Chiba, 1991; Vertes, 2004); VC/VS (Gutman et al., 2009; Haber et al., 2006); NAc (Gutman et al., 2009; Johansen-Berg et al., 2008; Powell and Leman, 1976; Sturm et al.,
2003; Williams et al., 1977); ITP (Aggleton and Mishkin, 1984; Behrens et al., 2003; Velasco et al., 2005). (b) Key regions involved in depressive symptomatology (Brody et al.,
2001; Drevets, 2001, 2003, 2007; Fava and Kendler, 2000; Gorwood, 2008; Kringelbach, 2005; Krishnan and Nestler, 2008; Maletic et al., 2007; Mayberg et al., 1999; Nestler
et al., 2002; Ongür and Price, 2000). Abbreviations: ACG, anterior cingulate gyrus; DR, dorsal raphe; H, hypothalamus; LC, locus coeruleus; NAc, nucleus accumbens; OFC,
orbital frontal cortex; PFC, prefrontal cortex; SCG, subgenual cingulate gyrus; VTA, ventral tegmental area.
R.J. Anderson et al. / Neuroscience and Biobehavioral Reviews 36 (2012) 1920–1933 1929

et al., 1989b), dopamine (Ji and Shepard, 2007; Lecourtier et al., Andrews, G., Henderson, S., Hall, W., 2001. Prevalence, comorbidity, disability and
2008; Matsumoto and Hikosaka, 2007) and norepinephrine (Kalén service utilisation. Overview of the Australian national mental health survey.
British Journal of Psychiatry 178, 145–153.
et al., 1989a) concentrations. SCG stimulation increases hippocam- Aouizerate, B., Cuny, E., Bardinet, E., Yelnik, J., Martin-Guehl, C., Rotge, J.-Y., Rougier,
pal serotonin levels (Hamani et al., 2010b) and NAc stimulation A., Bioulac, B., Tignol, J., Mallet, L., Burbaud, P., Guehl, D., 2009. Distinct striatal
increases both dopamine and serotonin concentrations in the NAc targets in treating obsessive–compulsive disorder and major depression. Journal
of Neurosurgery 111, 775–779.
(Sesia et al., 2010) and decreases dopamine and norepinephrine Aouizerate, B., Cuny, E., Martin-Guehl, C., Guehl, D., Amieva, H., Benazzouz, A.,
levels in the prefrontal cortex (Falowski et al., 2011). Preclinical Fabrigoule, C., Allard, M., Rougier, A., Bioulac, B., Tignol, J., Burbaud, P., 2004.
research has also shown that the integrity of the serotonergic sys- Deep brain stimulation of the ventral caudate nucleus in the treatment of
obsessive–compulsive disorder and major depression. Journal of Neurosurgery
tem is important for the antidepressant-like effects observed in the
101, 682–686.
forced swim test following DBS of the ventromedial prefrontal cor- Arias-Carrión, O., Stamelou, M., Murillo-Rodríguez, E., Menéndez-González, M.,
tex, considered to be the rodent homologue of the human SCG, Pöppel, E., 2010. Dopaminergic reward system: a short integrative review. Inter-
national Archives of Medicine 3, 24.
whereas norepinephrine depleting lesions failed to influence the
Bakish, D., 2001. New standard of depression treatment: remission and full recovery.
antidepressant-like results (Hamani et al., 2010b). Journal of Clinical Psychiatry 62, 5–9.
The preceding discussion presents some of the significant work Behrens, T.E.J., Johansen-Berg, H., Woolrich, M.W., Smith, S.M., Wheeler-Kingshott,
undertaken to understand the mechanisms of action underlying the C.A.M., Boulby, P.A., Barker, G.J., Sillery, E.L., Sheehan, K., Ciccarelli, O., Thompson,
A.J., Brady, J.M., Matthews, P.M., 2003. Non-invasive mapping of connections
therapeutic effects of DBS for TRD; it also highlights limitations in between human thalamus and cortex using diffusion imaging. Nature Neuro-
our understanding. The mechanisms of action are likely to involve science 6, 750–757.
an integration of acute and long-term effects of stimulation of the Benabid, A.L., 2003. Deep brain stimulation for Parkinson’s disease. Current Opinion
in Neurobiology 13, 696–706.
target area with effects on distal brain regions within dysfunctional Berk, M., Kapczinski, F., Andreazza, A.C., Dean, O.M., Giorlando, F., Maes, M., Yücel,
neural networks. The fact that a number of different targets are M., Gama, C.S., Dodd, S., Dean, B., Magalhães, P.V.S., Amminger, P., McGorry, P.,
able to effectively ameliorate depressive symptoms, including all Malhi, G.S., 2011. Pathways underlying neuroprogression in bipolar disorder:
focus on inflammation, oxidative stress and neurotrophic factors. Neuroscience
subgroups of symptoms, in groups of severely treatment resistant and Biobehavioral Reviews 35, 804–817.
patients supports the need for a neural network explanation of Berk, M., Ng, F., Wang, W.V., Calabrese, J.R., Mitchell, P.B., Malhi, G.S., Tohen, M.,
the mechanisms of action involving a number of neuroanatomical 2008. The empirical redefinition of the psychometric criteria for remission in
bipolar disorder. Journal of Affective Disorders 106, 153–158.
regions in cortical–limbic circuits.
Beurrier, C., Bioulac, B., Audin, J., Hammond, C., 2001. High-frequency stimulation
produces a transient blockade of voltage-gated currents in subthalamic neurons.
Journal of Neurophysiology 85, 1351–1356.
5. Conclusion Bewernick, B.H., Hurlemann, R., Matusch, A., Kayser, S., Grubert, C., Hadrysiewicz,
B., Axmacher, N., Lemke, M., Cooper-Mahkorn, D., Cohen, M.X., Brockmann, H.,
Lenartz, D., Sturm, V., Schlaepfer, T.E., 2010. Nucleus accumbens deep brain
Although a limited number of studies have been conducted, stimulation decreases ratings of depression and anxiety in treatment-resistant
involving only small patient numbers, DBS for TRD has shown depression. Biological Psychiatry 67, 110–116.
very encouraging results, demonstrating significant and sustained Bewernick, B.H., Kayser, S., Sturm, V., Schlaepfer, T.E., 2012. Long-term effects of
nucleus accumbens deep brain stimulation in treatment-resistant depression:
improvements in depressive symptomatology with remission evidence for sustained efficacy. Neuropsychopharmacology, 1–11 (Epub ahead
achieved in a large percentage of cases. Large randomised placebo- of print).
controlled studies will be required for more robust interpretations Bostwick, J.M., Pankratz, V.S., 2000. Affective disorders and suicide risk: a reexami-
nation. American Journal of Psychiatry 157, 1925–1932.
of efficacy. Despite the impressive therapeutic effects of DBS for Brådvik, L., Mattisson, C., Bogren, M., Nettelbladt, P., 2008. Long-term suicide risk of
TRD, the mechanisms of action are still incompletely understood. depression in the lundby cohort 1947–1997 – severity and gender. Acta Psychi-
Research posits mechanisms operating at the neuron level through atrica Scandinavica 117, 185–191.
Brody, A.L., Saxena, S., Mandelkern, M.A., Fairbanks, L.A., Ho, M.L., Baxter, L.R., 2001.
to modulation of dysfunctional neural networks. Unpicking the Brain metabolic changes associated with symptom factor improvement in major
mechanisms and pathways operative in the clinical efficacy of DBS depressive disorder. Biological Psychiatry 50, 171–178.
has the additional potential to shed light on underlying pathophys- Butson, C.R., Cooper, S.E., Henderson, J.M., McIntyre, C.C., 2007. Patient-specific anal-
ysis of the volume of tissue activated during deep brain stimulation. Neuroimage
iology, with broader implications for the development of therapies.
34, 661–670.
The impact of DBS on biomarkers and on other potential aetiologi- Caldecott-Hazard, S., Mazziotta, J., Phelps, M., 1988. Cerebral correlates of depressed
cal factors involved in the pathophysiology of depression remains behavior in rats, visualized using 14c-2-deoxyglucose autoradiography. Journal
of Neuroscience 8, 1951–1961.
a goal for future investigation.
Carney, R.M., Freedland, K.E., 2003. Depression, mortality, and medical mor-
bidity in patients with coronary heart disease. Biological Psychiatry 54,
241–247.
References Caspi, A., Sugden, K., Moffitt, T.E., Taylor, A., Craig, I.W., Harrington, H., Mcclay,
J., Mill, J., Martin, J., Braithwaite, A., Poulton, R., 2003. Influence of life stress
Abelson, J.L., Curtis, G.C., Sagher, O., Albucher, R.C., Harrigan, M., Taylor, on depression: moderation by a polymorphism in the 5-htt gene. Science 301,
S.F., Martis, B., Giordani, B., 2005. Deep brain stimulation for refractory 386–389.
obsessive–compulsive disorder. Biological Psychiatry 57, 510–516. Chiba, T., Kayahara, T., Nakano, K., 2001. Efferent projections of infralimbic and pre-
Abercrombie, H.C., Schaefer, S.M., Larson, C.L., Oakes, T.R., Lindgren, K.A., Holden, limbic areas of the medial prefrontal cortex in the Japanese monkey, macaca
J.E., Perlman, S.B., Turski, P.A., Krahn, D.D., Benca, R.M., Davidson, R.J., 1998. fuscata. Brain Research 888, 83–101.
Metabolic rate in the right amygdala predicts negative affect in depressed Christmas, D., Eljamel, M.S., Butler, S., Hazari, H., MacVicar, R., Steele, J.D., Liv-
patients. Neuroreport 9, 3301–3307. ingstone, A., Matthews, K., 2011. Long term outcome of thermal anterior
Aggleton, J.P., Mishkin, M., 1984. Projections of the amygdala to the thalamus in the capsulotomy for chronic, treatment refractory depression. Journal of Neurology,
cynomolgus monkey. Journal of Comparative Neurology 222, 56–68. Neurosurgery and Psychiatry 82, 594–600.
American Psychiatric Association, 2000. Practice guideline for the treatment of Cuijpers, P., Smit, F., 2002. Excess mortality in depression: a meta-
patients with major depressive disorder (revision). American Journal of Psy- analysis of community studies. Journal of Affective Disorders 72,
chiatry 157, 1–45. 227–236.
Anderson, M.E., Postupna, N., Ruffo, M., 2003. Effects of high-frequency stimulation De Hert, M., Correll, C.U., Bobes, J., Cetkovich-Bakmas, M., Cohen, D., Asai, I., Detraux,
in the internal globus pallidus on the activity of thalamic neurons in the awake J., Gautam, S., Möller, H.-J., Ndetei, D.M., Newcomer, J.W., Uwakwe, R., Leucht,
monkey. Journal of Neurophysiology 89, 1150–1160. S., 2011. Physical illness in patients with severe mental disorders. I. Preva-
Anderson, T.R., Hu, B., Iremonger, K., Kiss, Z.H.T., 2006. Selective attenuation lence, impact of medications and disparities in health care. World Psychiatry 10,
of afferent synaptic transmission as a mechanism of thalamic deep brain 52–77.
stimulation-induced tremor arrest. Journal of Neuroscience 26, 841–850. Dodd, S., Horgan, D., Malhi, G.S., Berk, M., 2005. To combine or not to combine? A
Andrade, L., Caraveo-Anduaga, J.J., Berglund, P., Bijl, R.V., De Graaf, R., Vollebergh, W., literature review of antidepressant combination therapy. Journal of Affective
Dragomirecka, E., Kohn, R., Keller, M., Kessler, R.C., Kawakami, N., Kiliç, C., Offord, Disorders 89, 1–11.
D., Ustun, T.B., Wittchen, H.-U., 2003. The epidemiology of major depressive Dostrovsky, J.O., Levy, R., Wu, J.P., Hutchison, W.D., Tasker, R.R., Lozano, A.M., 2000.
episodes: results from the international consortium of psychiatric epidemiology Microstimulation-induced inhibition of neuronal firing in human globus pal-
(icpe) surveys. International Journal of Methods in Psychiatric Research 12, 3–21. lidus. Journal of Neurophysiology 84, 570–574.
1930 R.J. Anderson et al. / Neuroscience and Biobehavioral Reviews 36 (2012) 1920–1933

Dostrovsky, J.O., Lozano, A.M., 2002. Mechanisms of deep brain stimulation. Move- prefrontal cortex deep brain stimulation in rats. Journal of Psychiatric Research
ment Disorders 17, S63–S68. 44, 683–687.
Dougherty, D.D., Rauch, S.L., 2007. Somatic therapies for treatment-resistant depres- Hamani, C., Diwan, M., Macedo, C.E., Brandão, M.L., Shumake, J., Gonzalez-
sion: new neurotherapeutic interventions. Psychiatric Clinics of North America lima, F., Raymond, R., Lozano, A.M., Fletcher, P.J., Nobrega, J.N., 2010b.
30, 31–37. Antidepressant-like effects of medial prefrontal cortex deep brain stimulation
Dougherty, D.D., Weiss, A.P., Cosgrove, G.R., Alpert, N.M., Cassem, E.H., Nieren- in rats. Biological Psychiatry 67, 117–124.
berg, A.A., Price, B.H., Mayberg, H.S., Fischman, A.J., Rauch, S.L., 2003. Cerebral Hamani, C., Mayberg, H., Snyder, B., Giacobbe, P., Kennedy, S., Lozano, A.M., 2009.
metabolic correlates as potential predictors of response to anterior cin- Deep brain stimulation of the subcallosal cingulate gyrus for depression:
gulotomy for treatment of major depression. Journal of Neurosurgery 99, anatomical location of active contacts in clinical responders and a suggested
1010–1017. guideline for targeting. Journal of Neurosurgery 111, 1209–1215.
Drevets, W.C., 1999. Prefrontal cortical-amygdalar metabolism in major depression. Hamani, C., Mayberg, H., Stone, S., Laxton, A., Haber, S., Lozano, A.M., 2011. The sub-
Annals of the New York Academy of Sciences 877, 614–637. callosal cingulate gyrus in the context of major depression. Biological Psychiatry
Drevets, W.C., 2000. Neuroimaging studies of mood disorders. Biological Psychiatry 69, 301–308.
48, 813–829. Hamilton, M., 1960. A rating scale for depression. Journal of Neurology, Neurosurgery
Drevets, W.C., 2001. Neuroimaging and neuropathological studies of depression: and Psychiatry 23, 56–62.
implications for the cognitive-emotional features of mood disorders. Current Hammond, C., Ammari, R., Bioulac, B., Garcia, L., 2008. Latest view on the mechanism
Opinion in Neurobiology 11, 240–249. of action of deep brain stimulation. Movement Disorders 23, 2111–2121.
Drevets, W.C., 2003. Neuroimaging abnormalities in the amygdala in mood disor- Hariz, G.M., Blomstedt, P., Koskinen, L.O., 2008. Long-term effect of deep brain stimu-
ders. Annals of the New York Academy of Sciences 985, 420–444. lation for essential tremor on activities of daily living and health-related quality
Drevets, W.C., 2007. Orbitofrontal cortex function and structure in depression. of life. Acta Neurologica Scandinavica 118, 387–394.
Annals of the New York Academy of Sciences 1121, 499–527. Hashimoto, T., Elder, C.M., Okun, M.S., Patrick, S.K., Vitek, J.L., 2003. Stimulation of the
Drevets, W.C., Bogers, W., Raichle, M.E., 2002. Functional anatomical correlates of subthalamic nucleus changes the firing pattern of pallidal neurons. The Journal
antidepressant drug treatment assessed using pet measures of regional glucose of Neuroscience 23, 1916–1923.
metabolism. European Neuropsychopharmacology 12, 527–544. Hasin, D.S., Goodwin, R.D., Stinson, F.S., Grant, B.F., 2005. Epidemiology of major
Drevets, W.C., Price, J.L., Furey, M.L., 2008. Brain structural and functional abnormal- depressive disorder. Archives of General Psychiatry 62, 1097–1106.
ities in mood disorders: implications for neurocircuitry models of depression. Hasler, G., Fromm, S., Carlson, P.J., Luckenbaugh, D.A., Waldeck, T., Geraci, M.,
Brain Structure and Function 213, 93–118. Roiser, J.P., Neumeister, A., Meyers, N., Charney, D.S., Drevets, W.C., 2008. Neu-
Drevets, W.C., Videen, T.O., Price, J.L., Preskorn, S.H., Carmichael, S.T., Raichle, M.E., ral response to catecholamine depletion in unmedicated subjects with major
1992. A functional anatomical study of unipolar depression. Journal of Neuro- depressive disorder in remission and healthy subjects. Archives of General Psy-
science 12, 3628–3641. chiatry 65, 521–531.
Duman, R.S., 2002. Pathophysiology of depression: the concept of synaptic plasticity. Hauptman, J.S., DeSalles, A.A., Espinoza, R., Sedrak, M., Ishida, W., 2008. Potential
European Psychiatry 17 (Suppl. 3), 306–310. surgical targets for deep brain stimulation in treatment-resistant depression.
Epstein, J., Pan, H., Kocsis, J.H., Yang, Y., Butler, T., Chusid, J., Hochberg, H., Mur- Neurosurgical Focus 25, 1–9.
rough, J., Strohmayer, E., Stern, E., Silbersweig, D.A., 2006. Lack of ventral striatal Heien, M.L.A.V., Wightman, R.M., 2006. Phasic dopamine signaling during behav-
response to positive stimuli in depressed versus normal subjects. American ior, reward, and disease states. CNS & Neurological Disorders – Drug Targets 5,
Journal of Psychiatry 163, 1784–1790. 99–108.
Falowski, S.M., Sharan, A., Reyes, B.A.S., Sikkema, C., Szot, P., Van Bockstaele, E.J., Heim, C., Nemeroff, C.B., 2001. The role of childhood trauma in the neurobiology of
2011. An evaluation of neuroplasticity and behavior after deep brain stimulation mood and anxiety disorders: preclinical and clinical studies. Biological Psychi-
of the nucleus accumbens in an animal model of depression. Neurosurgery 69, atry 49, 1023–1039.
1281–1290. Herkenham, M., Nauta, W.J., 1977. Afferent connections of the habenular nuclei in
Fava, M., 2003. Diagnosis and definition of treatment-resistant depression. Biological the rat. A horseradish peroxidase study, with a note on the fiber-of-passage
Psychiatry 53, 649–659. problem. Journal of Comparative Neurology 173, 123–146.
Fava, M., Kendler, K.S., 2000. Major depressive disorder. Neuron 28, 335–341. Herkenham, M., Nauta, W.J., 1979. Efferent connections of the habenular nuclei in
Freedman, L.J., Insel, T.R., Smith, Y., 2000. Subcortical projections of area 25 (sub- the rat. Journal of Comparative Neurology 187, 19–47.
genual cortex) of the macaque monkey. Journal of Comparative Neurology 421, Hikosaka, O., 2010. The habenula: from stress evasion to value-based decision-
172–188. making. Nature Reviews Neuroscience 11, 503–513.
Gabriels, L., Cosyns, P., Nuttin, B., Demeulemeester, H., Gybels, J., 2003. Deep Hikosaka, O., Sesack, S.R., Lecourtier, L., Shepard, P.D., 2008. Habenula: crossroad
brain stimulation for treatment-refractory obsessive–compulsive disorder: between the basal ganglia and the limbic system. Journal of Neuroscience 28,
psychopathological and neuropsychological outcome in three cases. Acta Psy- 11825–11829.
chiatrica Scandinavica 107, 275–282. Hodaie, M., Wennberg, R.A., Dostrovsky, J.O., Lozano, A.M., 2002. Chronic anterior
Geisler, S., Trimble, M., 2008. The lateral habenula: no longer neglected. CNS Spec- thalamus stimulation for intractable epilepsy. Epilepsia 43, 603–608.
trums 13, 484–489. Holtzheimer, P.E., Kelley, M.E., Gross, R.E., Filkowski, M.M., Garlow, S.J., Barrocas,
Geisler, S., Zahm, D.S., 2005. Afferents of the ventral tegmental area in the A., Wint, D., Craighead, M.C., Kozarsky, J., Chismar, R., Moreines, J.L., Mewes, K.,
rat-anatomical substratum for integrative functions. Journal of Comparative Posse, P.R., Gutman, D.A., Mayberg, H.S., 2012. Subcallosal cingulate deep brain
Neurology 490, 270–294. stimulation for treatment-resistant unipolar and bipolar depression. Archives
Giacobbe, P., Mayberg, H.S., Lozano, A.M., 2009. Treatment resistant depression as a of General Psychiatry 69, 150–158.
failure of brain homeostatic mechanisms: implications for deep brain stimula- Iremonger, K.J., Anderson, T.R., Hu, B., Kiss, Z.H.T., 2006. Cellular mechanisms
tion. Experimental Neurology 219, 44–52. preventing sustained activation of cortex during subcortical high-frequency
Gorwood, P., 2008. Neurobiological mechanisms of anhedonia. Dialogues in Clinical stimulation. Journal of Neurophysiology 96, 613–621.
Neuroscience 10, 291–299. Jacka, F.N., Kremer, P.J., Berk, M., de Silva-Sanigorski, A.M., Moodie, M., Leslie, E.R.,
Goto, Y., Grace, A.A., 2005. Dopaminergic modulation of limbic and cortical drive of Pasco, J.A., Swinburn, B.A., 2011a. A prospective study of diet quality and mental
nucleus accumbens in goal-directed behavior. Nature Neuroscience 8, 805–812. health in adolescents. PLoS One 6, e24805.
Greenberg, B., Price, L., Rauch, S., Friehs, G., Noren, G., Malone, D., Carpenter, L., Rezai, Jacka, F.N., Pasco, J.A., Williams, L.J., Leslie, E.R., Dodd, S., Nicholson, G.C., Kotowicz,
A., Rasmussen, S., 2003. Neurosurgery for intractable obsessive–compulsive dis- M.A., Berk, M., 2011b. Lower levels of physical activity in childhood asso-
order and depression: critical issues. Neurosurgery Clinics of North America 14, ciated with adult depression. Journal of Science and Medicine in Sport 14,
199–212. 222–226.
Greenberg, B.D., Malone, D.A., Friehs, G.M., Rezai, A.R., Kubu, C.S., Malloy, P.F., Ji, H., Shepard, P.D., 2007. Lateral habenula stimulation inhibits rat midbrain
Salloway, S.P., Okun, M.S., Goodman, W.K., Rasmussen, S.A., 2006. Three-year dopamine neurons through a gaba(a) receptor-mediated mechanism. Journal
outcomes in deep brain stimulation for highly resistant obsessive–compulsive of Neuroscience 27, 6923–6930.
disorder. Neuropsychopharmacology 31, 2384–2393. Jiménez, F., Velasco, F., Salin-Pascual, R., Hernández, J.A., Velasco, M., Criales, J.L.,
Grill, W.M., Cantrell, M.B., Robertson, M.S., 2008. Antidromic propagation of action Nicolini, H., 2005. A patient with a resistant major depression disorder treated
potentials in branched axons: implications for the mechanisms of action of deep with deep brain stimulation in the inferior thalamic peduncle. Neurosurgery 57,
brain stimulation. Journal of Computational Neuroscience 24, 81–93. 585–593.
Grubert, C., Hurlemann, R., Bewernick, B.H., Kayser, S., Hadrysiewicz, B., Axmacher, Johansen-Berg, H., Gutman, D.A., Behrens, T.E., Matthews, P.M., Rushworth, M.F.S.,
N., Sturm, V., Schlaepfer, T.E., 2011. Neuropsychological safety of nucleus Katz, E., Lozano, A.M., Mayberg, H.S., 2008. Anatomical connectivity of the
accumbens deep brain stimulation for major depression: effects of 12-month subgenual cingulate region targeted with deep brain stimulation for treatment-
stimulation. World Journal of Biological Psychiatry 12, 516–527. resistant depression. Cerebral Cortex 18, 1374–1383.
Gutman, D.A., Holtzheimer, P.E., Behrens, T.E.J., Johansen-berg, H., Mayberg, H.S., Johnson, M.D., Miocinovic, S., McIntyre, C.C., Vitek, J.L., 2008. Mechanisms and tar-
2009. A tractography analysis of two deep brain stimulation white matter targets gets of deep brain stimulation in movement disorders. Neurotherapeutics 5,
for depression. Biological Psychiatry 65, 276–282. 294–308.
Haber, S.N., Kim, K.-S., Mailly, P., Calzavara, R., 2006. Reward-related cortical inputs Kalbe, E., Voges, J., Weber, T., Haarer, M., Baudrexel, S., Klein, J.C., Kessler, J., Sturm,
define a large striatal region in primates that interface with associative corti- V., Heiss, W.D., Hilker, R., 2009. Frontal fdg-pet activity correlates with cognitive
cal connections, providing a substrate for incentive-based learning. Journal of outcome after stn-dbs in Parkinson disease. Neurology 72, 42–49.
Neuroscience 26, 8368–8376. Kalén, P., Lindvall, O., Björklund, A., 1989a. Electrical stimulation of the lateral
Hamani, C., Diwan, M., Isabella, S., Lozano, A.M., Nobrega, J.N., 2010a. Effects of dif- habenula increases hippocampal noradrenaline release as monitored by in vivo
ferent stimulation parameters on the antidepressant-like response of medial microdialysis. Experimental Brain Research 76, 239–245.
R.J. Anderson et al. / Neuroscience and Biobehavioral Reviews 36 (2012) 1920–1933 1931

Kalén, P., Strecker, R.E., Rosengren, E., Björklund, A., 1989b. Regulation of striatal Lozano, A.M., Mayberg, H.S., Giacobbe, P., Hamani, C., Craddock, R.C., Kennedy, S.H.,
serotonin release by the lateral habenula–dorsal raphe pathway in the rat as 2008. Subcallosal cingulate gyrus deep brain stimulation for treatment-resistant
demonstrated by in vivo microdialysis: role of excitatory amino acids and gaba. depression. Biological Psychiatry 64, 461–467.
Brain Research 492, 187–202. Lujan, J.L., Chaturvedi, A., Mcintyre, C.C., 2008. Tracking the mechanisms of deep
Kapczinski, F., Vieta, E., Andreazza, A.C., Frey, B.N., Gomes, F.A., Tramontina, J., brain stimulation for neuropsychiatric disorders. Frontiers in Bioscience 13,
Kauer-Sant’anna, M., Grassi-Oliveira, R., Post, R.M., 2008. Allostatic load in bipo- 5892–5904.
lar disorder: implications for pathophysiology and treatment. Neuroscience & Maciunas, R.J., Maddux, B.N., Riley, D.E., Whitney, C.M., Schoenberg, M.R., Ogrocki,
Biobehavioral Reviews 32, 675–692. P.J., Albert, J.M., Gould, D.J., 2007. Prospective randomized double-blind trial
Kendler, K.S., Kuhn, J.W., Vittum, J., Prescott, C.A., Riley, B., 2005a. The interac- of bilateral thalamic deep brain stimulation in adults with tourette syndrome.
tion of stressful life events and a serotonin transporter polymorphism in the Journal of Neurosurgery 107, 1004–1014.
prediction of episodes of major depression. Archives of General Psychiatry 62, Maes, M., Leonard, B., Fernandez, A., Kubera, M., Nowak, G., Veerhuis, R., Gard-
529–535. ner, A., Ruckoanich, P., Geffard, M., Altamura, C., Galecki, P., Berk, M., 2011.
Kendler, K.S., Myers, J., Prescott, C.A., 2005b. Sex differences in the relationship (Neuro)inflammation and neuroprogression as new pathways and drug targets
between social support and risk for major depression: a longitudinal study of in depression: from antioxidants to kinase inhibitors. Progress in Neuro-
opposite-sex twin pairs. American Journal of Psychiatry 162, 250–256. Psychopharmacology and Biological Psychiatry 35, 659–663.
Kendler, K.S., Thornton, L.M., Gardner, C.O., 2001. Genetic risk, number of previ- Maletic, V., Robinson, M., Oakes, T., Iyengar, S., Ball, S.G., Russell, J., 2007. Neurobi-
ous depressive episodes, and stressful life events in predicting onset of major ology of depression: an integrated view of key findings. International Journal of
depression. American Journal of Psychiatry 158, 582–586. Clinical Practice 61, 2030–2040.
Kennedy, S.H., Giacobbe, P., 2007. Treatment resistant depression – advances in Malhi, G.S., Berk, M., 2007. Does dopamine dysfunction drive depression? Acta Psy-
somatic therapies. Annals of Clinical Psychiatry 19, 279–287. chiatrica Scandinavica. Supplementum 115, 116–124.
Kennedy, S.H., Giacobbe, P., Rizvi, S.J., Placenza, F.M., Nishikawa, Y., Mayberg, H.S., Malone, D.A., 2010. Use of deep brain stimulation in treatment-resistant depression.
Lozano, A.M., 2011. Deep brain stimulation for treatment-resistant depression: Cleveland Clinic Journal of Medicine 77, S77–S80.
follow-up after 3 to 6 years. American Journal of Psychiatry 168, 502–510. Malone, D.A., Dougherty, D.D., Rezai, A.R., Carpenter, L.L., Friehs, G.M., Eskandar,
Kennedy, S.H., Konarski, J.Z., Segal, Z.V., Lau, M.A., Bieling, P.J., Mcintyre, R.S., May- E.N., Rauch, S.L., Rasmussen, S.A., Machado, A.G., Kubu, C.S., Tyrka, A.R., Price,
berg, H.S., 2007. Differences in brain glucose metabolism between responders to L.H., Stypulkowski, P.H., Giftakis, J.E., Rise, M.T., Malloy, P.F., Salloway, S.P.,
cbt and venlafaxine in a 16-week randomized controlled trial. American Journal Greenberg, B.D., 2009. Deep brain stimulation of the ventral capsule/ventral
of Psychiatry 164, 778–788. striatum for treatment-resistant depression. Biological Psychiatry 65,
Kessler, R.C., Berglund, P., Demler, O., Jin, R., Merikangas, K.R., Walters, E.E., 2005a. 267–275.
Lifetime prevalence and age-of-onset distributions of dsm-iv disorders in the Matsumoto, M., Hikosaka, O., 2007. Lateral habenula as a source of negative reward
national comorbidity survey replication. Archives of General Psychiatry 62, signals in dopamine neurons. Nature 447, 1111–1115.
593–602. Matsumoto, M., Hikosaka, O., 2008. Representation of negative motivational value
Kessler, R.C., Chiu, W.T., Demler, O., Walters, E.E., 2005b. Prevalence, severity, and in the primate lateral habenula. Nature Neuroscience 12, 77–84.
comorbidity of 12-month dsm-iv disorders in the national comorbidity survey Maurice, N., Thierry, A.-M., Glowinski, J., Deniau, J.-M., 2003. Spontaneous and
replication. Archives of General Psychiatry 62, 617–627. evoked activity of substantia nigra pars reticulata neurons during high-
Knutson, B., Adams, C.M., Fong, G.W., Hommer, D., 2001. Anticipation of increasing frequency stimulation of the subthalamic nucleus. Journal of Neuroscience 23,
monetary reward selectively recruits nucleus accumbens. Journal of Neuro- 9929–9936.
science 21, 1–5. Mayberg, H.S., 2003. Modulating dysfunctional limbic – cortical circuits in depres-
Kondo, H., Saleem, K.S., Price, J.L., 2005. Differential connections of the perirhinal sion: towards development of brain-based algorithms for diagnosis and
and parahippocampal cortex with the orbital and medial prefrontal networks optimised treatment. British Medical Bulletin 65, 193–207.
in macaque monkeys. Journal of Comparative Neurology 493, 479–509. Mayberg, H.S., 2006. Defining neurocircuits in depression. Psychiatric Annals 36,
Kopell, B.H., Greenberg, B.D., 2008. Anatomy and physiology of the basal ganglia: 259–269.
implications for dbs in psychiatry. Neuroscience & Biobehavioral Reviews 32, Mayberg, H.S., 2009. Targeted electrode-based modulation of neural circuits for
408–422. depression. Journal of Clinical Investigation 119, 717–725.
Kopell, B.H., Halverson, J., Butson, C.R., Dickinson, M., Bobholz, J., Harsch, H., Rainey, Mayberg, H.S., Brannan, S.K., Tekell, J.L., Silva, J.A., Mahurin, R.K., Mcginnis, S., Jer-
C., Kondziolka, D., Howland, R., Eskandar, E., Evans, K.C., Dougherty, D.D., 2011. abek, P.A., 2000. Regional metabolic effects of fluoxetine in major depression:
Epidural cortical stimulation of the left dorsolateral prefrontal cortex for refrac- serial changes and relationship to clinical response. Biological Psychiatry 48,
tory major depressive disorder. Neurosurgery 69, 1015–1029 (discussion 1029). 830–843.
Kosel, M., Frick, C., Lisanby, S.H., Fisch, H.-u., Schlaepfer, T.E., 2003. Magnetic seizure Mayberg, H.S., Liotti, M., Brannan, S.K., Mcginnis, S., Mahurin, R.K., Jerabek, P.A., Silva,
therapy improves mood in refractory major depression. Neuropsychopharma- J.A., Tekell, J.L., Martin, C.C., Lancaster, J.L., Fox, P.T., 1999. Reciprocal limbic-
cology 28, 2045–2048. cortical function and negative mood: converging pet findings in depression and
Kosel, M., Sturm, V., Frick, C., Lenartz, D., Zeidler, G., Brodesser, D., Schlaepfer, T.E., normal sadness. American Journal of Psychiatry 156, 675–682.
2007. Mood improvement after deep brain stimulation of the internal globus Mayberg, H.S., Lozano, A.M., Voon, V., McNeely, H.E., Seminowicz, D.A., Hamani,
pallidus for tardive dyskinesia in a patient suffering from major depression. C., Schwalb, J.M., Kennedy, S.H., 2005. Deep brain stimulation for treatment-
Journal of Psychiatric Research 41, 801–803. resistant depression. Neuron 45, 651–660.
Kringelbach, M.L., 2005. The human orbitofrontal cortex: linking reward to hedonic Mayberg, H.S., Silva, J.A., Brannan, S.K., Tekell, J.L., Mahurin, R.K., McGinnis, S., Jer-
experience. Nature Reviews Neuroscience 6, 691–702. abek, P.A., 2002. The functional neuroanatomy of the placebo effect. American
Kringelbach, M.L., Jenkinson, N., Green, A.L., Owen, S.L.F., Hansen, P.C., Cornelis- Journal of Psychiatry 159, 728–737.
sen, P.L., Holliday, I.E., Stein, J., Aziz, T.Z., 2007. Deep brain stimulation for McCracken, C.B., Grace, A.A., 2007. High-frequency deep brain stimulation of the
chronic pain investigated with magnetoencephalography. Neuroreport 18, nucleus accumbens region suppresses neuronal activity and selectively modu-
223–228. lates afferent drive in rat orbitofrontal cortex in vivo. Journal of Neuroscience
Krishnan, V., Nestler, E.J., 2008. The molecular neurobiology of depression. Nature 27, 12601–12610.
455, 894–902. McCracken, C.B., Grace, A.A., 2009. Nucleus accumbens deep brain stimulation pro-
Kuhn, J., Gaebel, W., Klosterkoetter, J., Woopen, C., 2009. Deep brain stimulation duces region-specific alterations in local field potential oscillations and evoked
as a new therapeutic approach in therapy-resistant mental disorders: ethical responses in vivo. Journal of Neuroscience 29, 5354–5363.
aspects of investigational treatment. European Archives of Psychiatry and Clin- McEwen, B.S., 2003. Mood disorders and allostatic load. Biological Psychiatry 54,
ical Neuroscience 259, S135–S141. 200–207.
Kuhn, J., Lenartz, D., Mai, J.K., Huff, W., Lee, S.-H., Koulousakis, A., Klosterkoet- McIntyre, C.C., Grill, W.M., Sherman, D.L., Thakor, N.V., 2004a. Cellular effects of deep
ter, J., Sturm, V., 2007. Deep brain stimulation of the nucleus accumbens and brain stimulation: model-based analysis of activation and inhibition. Journal of
the internal capsule in therapeutically refractory tourette-syndrome. Journal of Neurophysiology 91, 1457–1469.
Neurology 254, 963–965. McIntyre, C.C., Savasta, M., Goff L.K.-l. Vitek, J.L., 2004b. Uncovering the mech-
Lecourtier, L., Defrancesco, A., Moghaddam, B., 2008. Differential tonic influence of anism(s) of action of deep brain stimulation: activation, inhibition, or both.
lateral habenula on prefrontal cortex and nucleus accumbens dopamine release. Clinical Neurophysiology 115, 1239–1248.
European Journal of Neuroscience 27, 1755–1762. McNeely, H.E., Mayberg, H.S., Lozano, A.M., Kennedy, S.H., 2008. Neuropsychologi-
Li, B., Piriz, J., Mirrione, M., Chung, C., Proulx, C.D., Schulz, D., Henn, F., Malinow, R., cal impact of cg25 deep brain stimulation for treatment-resistant depression:
2011. Synaptic potentiation onto habenula neurons in the learned helplessness preliminary results over 12 months. Journal of Nervous and Mental Disease 196,
model of depression. Nature 470, 535–539. 405–410.
Liotti, M., Mayberg, H.S., 2001. The role of functional neuroimaging in the neuropsy- Montgomery, E.B., 2006. Effects of gpi stimulation on human thalamic neuronal
chology of depression. Journal of Clinical and Experimental Neuropsychology activity. Clinical Neurophysiology 117, 2691–2702.
23, 121–136. Montgomery, E.B., Gale, J.T., 2008. Mechanisms of action of deep brain stimulation
Lipsman, N., Neimat, J.S., Lozano, A.M., 2007. Deep brain stimulation for treatment- (dbs). Neuroscience & Biobehavioral Reviews 32, 388–407.
refractory obsessive–compulsive disorder: the search for a valid target. Montgomery, S.A., Asberg, M., 1979. A new depression scale designed to be sensitive
Neurosurgery 61, 1–13. to change. British Journal of Psychiatry 134, 382–389.
Lisanby, S.H., Luber, B., Schlaepfer, T.E., Sackeim, H.A., 2003. Safety and feasibility of Montoya, A., Weiss, A.P., Price, B.H., Cassem, E.H., Dougherty, D.D., Nierenberg,
magnetic seizure therapy (mst) in major depression: randomized within-subject A.A., Rauch, S.L., Cosgrove, G.R., 2002. Magnetic resonance imaging-guided
comparison with electroconvulsive therapy. Neuropsychopharmacology 28, stereotactic limbic leukotomy for treatment of intractable psychiatric disease.
1852–1865. Neurosurgery 50, 1043–1049 (discussion 1049–1052).
1932 R.J. Anderson et al. / Neuroscience and Biobehavioral Reviews 36 (2012) 1920–1933

Morris, J.S., Smith, K.A., Cowen, P.J., Friston, K.J., Dolan, R.J., 1999. Covariation of Rush, A.J., Thase, M.E., 1997. Strategies and tactics in the treatment of chronic depres-
activity in habenula and dorsal raphé nuclei following tryptophan depletion. sion. Journal of Clinical Psychiatry 58, 14–22.
Neuroimage 10, 163–172. Rush, A.J., Thase, M.E., Dube, S., 2003. Research issues in the study of difficult-to-treat
Mottaghy, F.M., Keller, C.E., Gangitano, M., Ly, J., Thall, M., Parker, J.A., Pascual-leone, depression. Biological Psychiatry 53, 743–753.
A., 2002. Correlation of cerebral blood flow and treatment effects of repetitive Sadock, B., Sadock, V., 2007. Kaplan & Sadock’s Synopsis of Psychiatry: Behavioral
transcranial magnetic stimulation in depressed patients. Psychiatry Research: Sciences/Clinical Psychiatry, 10th ed. Lippincott Williams & Wilkins Publishers,
Neuroimaging 115, 1–14. Philadelphia, PA, USA.
Murray, C.J.L., Lopez, A.D., 1997a. Alternative projections of mortality and disability Saleem, K.S., Kondo, H., Price, J.L., 2008. Complementary circuits connecting the
by cause 1990–2020: global burden of disease study. Lancet 349, 1498–1504. orbital and medial prefrontal networks with the temporal, insular, and oper-
Murray, C.J.L., Lopez, A.D., 1997b. Global mortality, disability, and the contribution cular cortex in the macaque monkey. Journal of Comparative Neurology 506,
of risk factors: global burden of disease study. Lancet 349, 1436–1442. 659–693.
Mykletun, A., Bjerkeset, O., Dewey, M., Prince, M., Overland, S., Stewart, R., 2007. Anx- Salin, P., Manrique, C., Forni, C., Goff, L.K.-l., 2002. High-frequency stimulation of
iety, depression, and cause-specific mortality: the hunt study. Psychosomatic the subthalamic nucleus selectively reverses dopamine denervation-induced
Medicine 69, 323–331. cellular defects in the output structures of the basal ganglia in the rat. Journal
Nahas, Z., Anderson, B.S., Borckardt, J., Arana, A.B., George, M.S., Reeves, S.T., Takacs, of Neuroscience 22, 5137–5148.
I., 2010. Bilateral epidural prefrontal cortical stimulation for treatment-resistant Sartorius, A., Kiening, K.L., Kirsch, P., von Gall, C.C., Haberkorn, U., Unterberg, A.W.,
depression. Biological Psychiatry 67, 101–109. Henn, F.A., Meyer-Lindenberg, A., 2010. Remission of major depression under
Neimat, J.S., Hamani, C., Giacobbe, P., Merskey, H., Kennedy, S.H., Mayberg, H.S., deep brain stimulation of the lateral habenula in a therapy-refractory patient.
Lozano, A.M., 2008. Neural stimulation successfully treats depression in patients Biological Psychiatry 67, e9–e11.
with prior ablative cingulotomy. American Journal of Psychiatry 165, 687–693. Schlaepfer, T.E., Bewernick, B., Kayser, S., Lenz, D., 2011. Modulating affect, cogni-
Nestler, E.J., Barrot, M., Dileone, R.J., Eisch, A.J., Gold, S.J., Monteggia, L.M., 2002. tion, and behavior – prospects of deep brain stimulation for treatment-resistant
Neurobiology of depression. Neuron 34, 13–25. psychiatric disorders. Frontiers in Integrative Neuroscience 5, 29.
Nestler, E.J., Carlezon, W.A., 2006. The mesolimbic dopamine reward circuit in Schlaepfer, T.E., Cohen, M.X., Frick, C., Kosel, M., Brodesser, D., Axmacher, N., Joe, A.Y.,
depression. Biological Psychiatry 59, 1151–1159. Kreft, M., Lenartz, D., Sturm, V., 2008. Deep brain stimulation to reward circuitry
Neumeister, A., Nugent, A.C., Waldeck, T., Geraci, M., Schwarz, M., Bonne, O., Bain, alleviates anhedonia in refractory major depression. Neuropsychopharmacol-
E.E., Luckenbaugh, D.A., Herscovitch, P., Charney, D.S., Drevets, W.C., 2004. Neu- ogy 33, 368–377.
ral and behavioral responses to tryptophan depletion in unmedicated patients Schulte, T., Brecht, S., Herdegen, T., Illert, M., Mehdorn, H.M., Hamel, W., 2006. Induc-
with remitted major depressive disorder and controls. Archives of General Psy- tion of immediate early gene expression by high-frequency stimulation of the
chiatry 61, 765–773. subthalamic nucleus in rats. Neuroscience 138, 1377–1385.
Nobler, M.S., Oquendo, M.A., Kegeles, L.S., Malone, K.M., Campbell, C., Sackeim, H.A., Scopinho, A.A., Scopinho, M., Lisboa, S.F., Correa, F.M.D.A., Guimarães, F.S., Joca,
Mann, J.J., 2001. Decreased regional brain metabolism after ect. American Journal S.R.L., 2010. Acute reversible inactivation of the ventral medial prefrontal cor-
of Psychiatry 158, 305–308. tex induces antidepressant-like effects in rats. Behavioural Brain Research 214,
O’Reardon, J.P., Solvason, H.B., Janicak, P.G., Sampson, S., Isenberg, K.E., Nahas, Z., 437–442.
Mcdonald, W.M., Avery, D., Fitzgerald, P.B., Loo, C., Demitrack, M.A., George, M.S., Seminowicz, D.A., Mayberg, H.S., Mcintosh, A.R., Goldapple, K., Kennedy, S., Segal, Z.,
Sackeim, H.A., 2007. Efficacy and safety of transcranial magnetic stimulation in Rafi-Tari, S., 2004. Limbic–frontal circuitry in major depression: a path modeling
the acute treatment of major depression: a multisite randomized controlled metanalysis. Neuroimage 22, 409–418.
trial. Biological Psychiatry 62, 1208–1216. Servello, D., Porta, M., Sassi, M., Brambilla, A., Robertson, M.M., 2008. Deep brain
Ongür, D., Ferry, A.T., Price, J.L., 2003. Architectonic subdivision of the human orbital stimulation in 18 patients with severe gilles de la tourette syndrome refractory
and medial prefrontal cortex. Journal of Comparative Neurology 460, 425–449. to treatment: the surgery and stimulation. Journal of Neurology, Neurosurgery
Ongür, D., Price, J.L., 2000. The organization of networks within the orbital and and Psychiatry 79, 136–142.
medial prefrontal cortex of rats, monkeys and humans. Cerebral Cortex 10, Sesia, T., Bulthuis, V., Tan, S., Lim, L.W., Vlamings, R., Blokland, A., Steinbusch, H.W.M.,
206–219. Sharp, T., Visser-Vandewalle, V., Temel, Y., 2010. Deep brain stimulation of
Perlmutter, J.S., Mink, J.W., 2006. Deep brain stimulation. Annual Review of Neuro- the nucleus accumbens shell increases impulsive behavior and tissue levels of
science 29, 229–257. dopamine and serotonin. Experimental Neurology 225, 302–309.
Pittenger, C., Duman, R.S., 2008. Stress, depression, and neuroplasticity: a conver- Shen, K.-Z., Zhu, Z.-T., Munhall, A., Johnson, S.W., 2003. Synaptic plasticity in rat sub-
gence of mechanisms. Neuropsychopharmacology 33, 88–109. thalamic nucleus induced by high-frequency stimulation. Synapse 50, 314–319.
Powell, E.W., Leman, R.B., 1976. Connections of the nucleus accumbens. Brain Shumake, J., Gonzalez-Lima, F., 2003. Brain systems underlying susceptibility to
Research 105, 389–403. helplessness and depression. Behavioral and Cognitive Neuroscience Reviews
Poynton, A.M., Kartsounis, L.D., Bridges, P.K., 1995. A prospective clinical study of 2, 198–221.
stereotactic subcaudate tractotomy. Psychological Medicine 25, 763–770. Slattery, D.A., Neumann, I.D., Cryan, J.F., 2011. Transient inactivation of the
Price, J.L., Drevets, W.C., 2010. Neurocircuitry of mood disorders. Neuropsychophar- infralimbic cortex induces antidepressant-like effects in the rat. Journal of Psy-
macology 35, 192–216. chopharmacology 25, 1295–1303.
Puigdemont, D., Pérez-Egea, R., Portella, M.J., Molet, J., de Diego-Adeliño, J., Gironell, Slotema, C.W., Blom, J.D., Hoek, H.W., Sommer, I.E.C., 2010. Should we expand
A., Radua, J., Gómez-Anson, B., Rodríguez, R., Serra, M., de Quintana, C., Artigas, the toolbox of psychiatric treatment methods to include repetitive transcra-
F., Alvarez, E., Pérez, V., 2012. Deep brain stimulation of the subcallosal cingulate nial magnetic stimulation (rtms)? A meta-analysis of the efficacy of rtms in
gyrus: further evidence in treatment-resistant major depression. International psychiatric disorders. Journal of Clinical Psychiatry 71, 873–884.
Journal of Neuropsychopharmacology 15, 121–133. Sturm, V., Lenartz, D., Koulousakis, A., Treuer, H., Herholz, K., Klein, J.C., Klosterkötter,
Quitkin, F.M., Stewart, J.W., McGrath, P.J., Nunes, E., Ocepeck-Welikson, K., Tricamo, J., 2003. The nucleus accumbens: a target for deep brain stimulation in obsessive-
E., Rabkin, J.G., Klein, D.F., 1993. Further evidence that a placebo response to compulsive- and anxiety-disorders. Journal of Chemical Neuroanatomy 26,
antidepressants can be identified. American Journal of Psychiatry 150, 566–570. 293–299.
Rasmussen, K.G., Sampson, S.M., Rummans, T.A., 2002. Electroconvulsive therapy Swaab, D.F., Bao, A.-M., Lucassen, P.J., 2005. The stress system in the human brain in
and newer modalities for the treatment of medication-refractory mental illness. depression and neurodegeneration. Ageing Research Reviews 4, 141–194.
Mayo Clinic Proceedings 77, 552–556. Takagishi, M., Chiba, T., 1991. Efferent projections of the infralimbic (area 25) region
Ressler, K.J., Mayberg, H.S., 2007. Targeting abnormal neural circuits in mood and of the medial prefrontal cortex in the rat. An anterograde tracer pha-l study.
anxiety disorders: from the laboratory to the clinic. Nature Neuroscience 10, Brain Research 566, 26–39.
1116–1125. Temperli, P., Ghika, J., Villemure, J.G., Burkhard, P.R., Bogousslavsky, J., Vingerhoets,
Ressler, K.J., Nemeroff, C.B., 2000. Role of serotonergic and noradrenergic systems F.J., 2003. How do parkinsonian signs return after discontinuation of subthalamic
in the pathophysiology of depression and anxiety disorders. Depression and dbs? Neurology 60, 78–81.
Anxiety 12, 2–19. Thase, M.E., Rush, A.J., 1997. When at first you don’t succeed: sequential strate-
Roiser, J.P., Levy, J., Fromm, S.J., Nugent, A.C., Talagala, S.L., Hasler, G., Henn, F.A., gies for antidepressant nonresponders. Journal of Clinical Psychiatry 58,
Sahakian, B.J., Drevets, W.C., 2009. The effects of tryptophan depletion on neu- 23–29.
ral responses to emotional words in remitted depression supplemental data. The Deep-brain Stimulation for Parkinson’s Disease Study Group, 2001. Deep-
Biological Psychiatry 66, 441–450. brain stimulation of the subthalamic nucleus or the pars interna of the globus
Rück, C., Karlsson, A., Steele, J.D., Edman, G., Meyerson, B.A., Ericson, K., Nyman, H., pallidus in Parkinson’s disease. The New England Journal of Medicine 345,
Asberg, M., Svanborg, P., 2008. Capsulotomy for obsessive–compulsive disorder. 956–963.
Archives of General Psychiatry 65, 914–922. The UK ECT Review Group, 2003. Efficacy and safety of electroconvulsive therapy
Ruhé, H.G., Mason, N.S., Schene, A.H., 2007. Mood is indirectly related to serotonin, in depressive disorders: a systematic review and meta-analysis. Lancet 361,
norepinephrine and dopamine levels in humans: a meta-analysis of monoamine 799–808.
depletion studies. Molecular Psychiatry 12, 331–359. Tremblay, L.K., Naranjo, C.A., Graham, S.J., Herrmann, N., Mayberg, H.S., Hevenor, S.,
Rush, A.J., Marangell, L.B., Sackeim, H.A., George, M.S., Brannan, S.K., Davis, S.M., Busto, U.E., 2005. Functional neuroanatomical substrates of altered reward pro-
Howland, R., Kling, M.A., Rittberg, B.R., Burke, W.J., Rapaport, M.H., Zajecka, J., cessing in major depressive disorder revealed by a dopaminergic probe. Archives
Nierenberg, A.A., Husain, M.M., Ginsberg, D., Cooke, R.G., 2005. Vagus nerve of General Psychiatry 62, 1228–1236.
stimulation for treatment-resistant depression: a randomized, controlled acute Trivedi, M.H., Fava, M., Wisniewski, S.R., Thase, M.E., Quitkin, F., Warden, D., Ritz, L.,
phase trial. Biological Psychiatry 58, 347–354. Nierenberg, A.A., Lebowitz, B., Biggs, M.M., Luther, J.F., Shores-Wilson, K., Rush,
Rush, A.J., Siefert, S.E., 2009. Clinical issues in considering vagus nerve stimulation A.J., 2006. Medication augmentation after the failure of ssris for depression. New
for treatment-resistant depression. Experimental Neurology 219, 36–43. England Journal of Medicine 354, 1243–1252.
R.J. Anderson et al. / Neuroscience and Biobehavioral Reviews 36 (2012) 1920–1933 1933

Tye, S.J., Miller, A.D., Blaha, C.D., 2009. Differential corticosteroid receptor regulation Watson, D., Clark, L.A., 1988. Development and validation of brief measures of pos-
of mesoaccumbens dopamine efflux during the peak and nadir of the circadian itive and negative affect: the panas scales. Journal of Personality and Social
rhythm: a molecular equilibrium in the midbrain? Synapse 63, 982–990. Psychology 6, 1063–1070.
Van Laere, K., Nuttin, B., Gabriels, L., Dupont, P., Rasmussen, S., Greenberg, B.D., Williams, D.J., Crossman, A.R., Slater, P., 1977. The efferent projections of the nucleus
Cosyns, P., 2006. Metabolic imaging of anterior capsular stimulation in refractory accumbens in the rat. Brain Research 130, 217–227.
obsessive–compulsive disorder: a key role for the subgenual anterior cingulate Williams, L.J., Brennan, S.L., Henry, M.J., Berk, M., Jacka, F.N., Nicholson, G.C., Kotow-
and ventral striatum. Journal of Nuclear Medicine 47, 740–747. icz, M.A., Pasco, J.A., 2011. Area-based socioeconomic status and mood disorders:
Velasco, F., Velasco, M., Jiménez, F., Velasco, A.L., Salin-pascual, R., 2005. Neu- cross-sectional evidence from a cohort of randomly selected adult women.
robiological background for performing surgical intervention in the inferior Maturitas 69, 173–178.
thalamic peduncle for treatment of major depression disorders. Neurosurgery Winter, C., Vollmayr, B., Djodari-Irani, A., Klein, J., Sartorius, A., 2011. Pharmaco-
57, 439–448. logical inhibition of the lateral habenula improves depressive-like behavior in
Vertes, R.P., 2004. Differential projections of the infralimbic and prelimbic cortex in an animal model of treatment resistant depression. Behavioural Brain Research
the rat. Synapse 51, 32–58. 216, 463–465.
Vidailhet, M., Vercueil, L., Houeto, J., Krystkowiak, P., Benabid, A.L., Cornu, P., World Health Organisation, 2008. The Global Burden of Disease 2004 Update. World
Lagrange, C., Montcel, S.T.D., Dormont, D., Grand, S., Blond, S., Detante, O., Pillon, Health Organisation, Geneva, Switzerland.
B., Ardouin, C., Agid, Y., Destée, A., Pollak, P., 2005. Bilateral deep-brain stimula- Wulsin, L.R., Vaillant, G.E., Wells, V.E., 1999. A systematic review of the mortality of
tion of the globus pallidus in primary generalized dystonia. New England Journal depression. Psychosomatic Medicine 61, 6–17.
of Medicine 352, 459–467. Yang, L.-M., Hu, B., Xia, Y.-H., Zhang, B.-L., Zhao, H., 2008. Lateral habenula
Warden, D., Rush, A.J., Trivedi, M.H., Fava, M., Wisniewski, S.R., 2007. The star*d lesions improve the behavioral response in depressed rats via increasing
project results: a comprehensive review of findings. Current Psychiatry Reports the serotonin level in dorsal raphe nucleus. Behavioural Brain Research 188,
9, 449–459. 84–90.

You might also like