ATDEmegent Nemeroff 2014

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

G Model

AJP-644; No. of Pages 10

Asian Journal of Psychiatry xxx (2014) xxx–xxx

Contents lists available at ScienceDirect

Asian Journal of Psychiatry


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

Review

Emerging antidepressants to treat major depressive disorder


Samantha G. Block a,*, Charles B. Nemeroff b,1
a
Department of Psychiatry and Behavioral Sciences, Jackson Memorial Hospital, University of Miami Hospital, Leonard M. Miller School of Medicine,
1695 N.W. 9th Avenue, Miami, FL 33136, USA
b
Department of Psychiatry and Behavioral Sciences, Center on Aging, Jackson Memorial Hospital, University of Miami Hospital, Leonard M. Miller School of
Medicine, University of Miami, 1120 Northwest 14 Street, Suite 1455, Miami, FL 33136, USA

A R T I C L E I N F O A B S T R A C T

Article history: Depression is a common disorder with an annual risk of a depressive episode in the United States of 6.6%.
Received 8 June 2014 Only 30–40% of patients remit with antidepressant monotherapy, leaving 60–70% of patients who do not
Received in revised form 3 September 2014 optimally respond to therapy. Unremitted depressive patients are at increased risk for suicide.
Accepted 6 September 2014
Considering the prevalence of treatment resistant depression and its consequences, treatment
Available online xxx
optimization is imperative. This review summarizes the latest treatment modalities for major
depressive disorder including pharmacotherapy, electroconvulsive therapy, repetitive transcranial
Keywords:
magnetic stimulation and psychotherapy. Through advancements in research to better understand the
Antidepressants
Major depressive disorder
pathophysiology of depression, advances in treatment will be realized.
Treatment ß 2014 Published by Elsevier B.V.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 000
2. Pharmacology currently used to treat depression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 000
2.1. Tricyclic antidepressants (TCAs) and monoamine oxidase inhibitors (MAOIs) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 000
2.2. Selective serotonin reuptake inhibitors (SSRIs) and serotonin norepinephrine reuptake inhibitors (SNRIs) . . . . . . . . . . . . . . . . . . . . 000
2.3. Atypical antidepressants. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 000
2.4. Nefazodone . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 000
2.5. Vortioxetine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 000
3. Antidepressants approved for use outside the United States. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 000
3.1. Tianeptine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 000
3.2. Moclobemide. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 000
3.3. Reboxetine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 000
3.4. Agomelatine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 000
4. N-methyl-D-aspartate (NMDA) receptors and depression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 000
5. The HPA axis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 000
5.1. Monotherapy targets of the hypothalamic pituitary axis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 000
5.2. Augmentation therapy with hypothalamic–pituitary–adrenal axis targets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 000
6. Omega-3 fatty acids. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 000
7. Inflammation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 000
8. Atypical antipsychotics (olanzapine, quetiapine, aripiprazole, and risperidone) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 000
8.1. Lithium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 000
8.2. Thyroid hormone augmentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 000
8.3. Buspirone augmentation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 000

* Corresponding author. Tel.: +1 305 355 8264.


E-mail addresses: sblock@med.miami.edu (S.G. Block), CNemeroff@med.miami.edu (C.B. Nemeroff).
1
Tel.: +1 305 243 6400/+1 305 243 8532.

http://dx.doi.org/10.1016/j.ajp.2014.09.001
1876-2018/ß 2014 Published by Elsevier B.V.

Please cite this article in press as: Block, S.G., Nemeroff, C.B., Emerging antidepressants to treat major depressive disorder. Asian J.
Psychiatry (2014), http://dx.doi.org/10.1016/j.ajp.2014.09.001
G Model
AJP-644; No. of Pages 10

2 S.G. Block, C.B. Nemeroff / Asian Journal of Psychiatry xxx (2014) xxx–xxx

8.4. Stimulant augmentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............... . 000


8.5. Atypical antipsychotics as augmenting agents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............... . 000
8.6. Other augmenting strategies: Pindolol, yohimbine, testosterone/estrogens, and lamotrigine. . . . . . . . . . . . . . . . ............... . 000
8.6.1. Pindolol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............... . 000
8.6.2. Yohimbine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............... . 000
8.6.3. Testosterone/estrogens . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............... . 000
8.6.4. Lamotrigine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............... . 000
8.6.5. Mianserin, mirtazapine, and desipramine: Combination therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............... . 000
9. Substance P . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............... . 000
10. Electroconvulsive therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............... . 000
11. Focal brain stimulation: Vagus nerve stimulation, transcranial magnetic stimulation, magnetic seizure therapy, deep brain stimulation . 000
11.1. Vagus nerve stimulation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............... . 000
11.2. Transcranial magnetic stimulation (TMS) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............... . 000
11.3. Theta burst stimulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............... . 000
11.4. Magnetic seizure therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............... . 000
11.5. Deep brain stimulation (DBS). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............... . 000
11.6. Transcranial direct current stimulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............... . 000
12. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............... . 000
13. Expert opinion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............... . 000
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............... . 000

1. Introduction 2. Pharmacology currently used to treat depression

Depression is a common disorder affecting nearly 16% of the 2.1. Tricyclic antidepressants (TCAs) and monoamine oxidase
population over the course of a lifetime with an annual risk of a inhibitors (MAOIs)
depressive episode of 6.6%. The condition is nearly twice as
common in women as in men. Only 30–40% remit with a single Tricyclic antidepressants were the first medication class used to
antidepressant treatment leaving nearly 60–70% of patients treat depression. Their primary mechanism of action is inhibition
who do not optimally respond (Kato and Chang, 2013a). Even of the reuptake of norepinephrine and/or serotonin. Monoamine
more problematic is the fact that the disease tends to recur, oxidase inhibitors block the enzymatic degradation of serotonin,
with greater than 75% of patients experiencing more than 1 norepinephrine and dopamine resulting in increased levels of
episode in a 10 year period. It is also responsible for the these neurotransmitters within the synaptic cleft. A large number
majority of almost 40,000 suicides per year in the United of randomized controlled clinical trials and meta-analyses confirm
States. Because of a lack of any validated predictors of response that both of these classes of drugs are effective in treating
to any of the evidence-based treatments for depression depression. However, the TCAs anticholinergic and antihistami-
including various approved antidepressants and psychothera- nergic side effect profiles limit their use (Holtzheimer and
pies, there is a trial and error approach to treating major Nemeroff, 2006) the TCAs can be lethal at 10 times or less the
depressive disorder. This concatenation of facts has led to a daily recommended dose, and can cause cardiotoxicity or seizures
large number of patients who have failed one or more FDA- as well (Schatzberg and Nemeroff, 2009). Additionally, problem-
approved treatments for depression, so-called treatment atic to the MAOIs, are their interaction with tyramine (and other
refractory depression (TRD). sympathomimetic amines) containing products. The selegiline
Older treatment modalities focus on three monoamine patch, a MAOI, is also FDA approved for the treatment of
neurotransmitter systems: serotonin, dopamine, and norepineph- depression. Unfortunately, as a class, these drugs have not shown
rine. Tricyclic antidepressants, selective serotonin reuptake increased efficacy in treatment refractory depression compared to
inhibitors (SSRIs), serotonin-norepinephrine reuptake inhibitors first-line therapy and their side effect profile may limit their use.
(SNRIs), and monoamine oxidase inhibitors (MAOIs) are all However, there is some evidence that clomipramine may be more
effective antidepressants that increase the availability of one or effective than the SSRIs (Duval et al., 2006).
more of these monoamines. Although pharmacotherapy is often
chosen based on which symptoms predominate in an individual 2.2. Selective serotonin reuptake inhibitors (SSRIs) and serotonin
including changes in sleep, appetite, interest, guilt, energy, norepinephrine reuptake inhibitors (SNRIs)
concentration, psychomotor behavior, and suicidality, there is
little evidence to suggest that such symptoms predict response to Currently first-line treatments of depression include the SSRIs
one antidepressant versus another. and SNRIs. The SSRIs (including fluoxetine, sertraline, paroxetine,
More recent attempts have sought to identify genetic citalopram, and escitalopram) were introduced into the United
predictors of response or nonresponse to individual treatment. States between the years of 1987–2003, while the SNRIs
For example, the serotonin transporter gene (5HTTLPR) may (duloxetine, venlafaxine and desvenlafaxine) were introduced
predict response to some SSRIs and the gene coding a between 1995 and 2008. As the name suggests, their mechanism of
glucocorticoid co-chaperone protein, FKBPS, may predict re- action involves inhibiting the reuptake of serotonin, or serotonin
sponse to antidepressant treatment (Binder et al., 2004). This and norepinephrine, respectively, increasing the concentration of
review will not summarize the genetics of depression treatment these neurotransmitters within the synaptic cleft. Some (fluoxe-
which is reviewed elsewhere (Ozomaro et al., 2013), but instead tine and paroxetine) are potent inhibitors of the cytochrome P450
focuses on the newest treatment modalities for major depressive IID6 isoenzyme. Both classes of these drugs in men and women are
disorder including pharmacotherapy, electroconvulsive therapy, associated with sexual dysfunction and similar to all other
repetitive transcranial magnetic stimulation (TMS), and psycho- antidepressants take 6 to 8 weeks to become effective (Berton
therapy. and Nestler, 2006). The STAR*D trial was an effectiveness trial

Please cite this article in press as: Block, S.G., Nemeroff, C.B., Emerging antidepressants to treat major depressive disorder. Asian J.
Psychiatry (2014), http://dx.doi.org/10.1016/j.ajp.2014.09.001
G Model
AJP-644; No. of Pages 10

S.G. Block, C.B. Nemeroff / Asian Journal of Psychiatry xxx (2014) xxx–xxx 3

which assessed the response to various treatments after failure to agonist at 5HT1A and 5HT1B receptors, has recently been approved
achieve remission with citalopram. Remarkably, less than a third of by the FDA for the treatment of major depression. Several studies
depressed patients in this open trial achieved remission after doses have been conducted assessing its efficacy at different dosages.
of up to 40 mg of the SSRI. One question of interest is whether a While not always statistically effective at under 5 mg (n = 611)
nonresponder to an SSRI, would respond to an SNRI? The literature (Mahableshwarkar et al., 2013; Henigsberg et al., 2012; Katona
is meager. The ARGOS and the STAR*D trials assessed this issue. et al., 2012), at dosages between 10 and 20 mg, the drug
ARGOS showed an 8% advantage in achieving remission when significantly reduced MADRS scores. There is evidence that
switching to a SNRI (venlafaxine), and the STAR*D trial showed a vortioxetine may have cognitive enhancing effects (Boulenger
7% advantage when using a SNRI; however neither of these et al., 2014). Side effects include nausea, headache, diarrhea, dry
differences were statistically significant (Baldomero et al., 2005; mouth and dizziness.
Rush et al., 2006). It is important to note that venlafaxine has been
reported to have the highest rate of lethality after overdose among 3. Antidepressants approved for use outside the United States
all of these agents, as well as a number of cardiovascular side
effects including hypertension and decreased heart rate variability. 3.1. Tianeptine

2.3. Atypical antidepressants Tianeptine, approved for the treatment of depression in Europe
and Asia, but not in the United States, has an unknown mechanism
The atypical antidepressants, a heterogeneous group, also of action, though it reportedly increases serotonin reuptake and in
reportedly affect monoamine neurotransmitters, receptors, or animal models restores neuroplasticity by increasing gene
transporters. These include bupropion, mirtazapine, nefazodone expression of BDNF and nerve growth factor in the hippocampus
and trazodone, as well as several others not approved for use in the and the amygdala. It also has been shown to decrease dendritic
US, but approved for use in other countries (mianserin, reboxetine, atrophy while normalizing glutamatergic neurotransmission
and tianeptine). The mechanism of action of bupropion remains (Kasper and McEwen, 2008).
obscure, as it is a very weak dopamine and norepinephrine Tianeptine is as effective as amitriptyline, clomipramine,
reuptake inhibitor. It has the advantage of lack of sexual imipramine and paroxetine in treating depression. Not only is
dysfunction. It is also approved for smoking cessation and has a this drug beneficial in treating depression, but it has anti-anxiety
relatively favorable side effect profile, though unlike the SSRIs and properties as well (Wagstaff et al., 2001).
SNRIs it is not effective in treating syndromal anxiety disorders A meta-analysis involving two studies showed no adverse
(Schatzberg and Nemeroff, 2009). Mirtazapine increases the effects on libido. Adverse effects most commonly included
release of noradrenaline, is a mild reuptake inhibitor and blocks headache, nausea, constipation, abdominal pain, dizziness, dream
a-2 and 5HT receptor subtypes. It is approved for use in depression alterations, and dry mouth. The side effect profile is clearly
and is reportedly particularly effective in elderly patients. superior to that of TCAs (Wagstaff et al., 2001).
Mirtazapine has also shown efficacy in the treatment of dysthymia,
PTSD, social phobia, OCD, sleep disorders and chronic or recurrent 3.2. Moclobemide
pain (Schatzberg and Nemeroff, 2009). A related drug, mianserin,
has a similar mechanism of action, but is not approved for use in Moclobemide, a reversible inhibitor of monoamine oxidase A,
the US. Several meta-analyses, demonstrated the efficacy of both increases the concentration of monoamines in the synaptic cleft. A
mirtazapine and mianserin (Lopes Rocha et al., 2013); the major meta-analysis found moclobemide (OR = 6.98) to be significantly
drawback of mirtazapine is weight gain and sedation. Trazodone more effective than placebo. Additionally, pairwise comparison
both antagonizes the 5HT2 receptor and mildly inhibits the found moclobemide superior to fluoxetine (OR = 2.38) (Kriston
reuptake of serotonin or noradrenaline. Side effects of trazodone et al., 2014). Four large comparative trials have been published of
include sedation, orthostatic hypotension, and priapism, but the moclobemide treatment in depression. Silverstone (2001) demon-
latter is a rare event (Holtzheimer and Nemeroff, 2006; Schatzberg strated its efficacy in bipolar depression Additionally, bipolar
and Nemeroff, 2009). patients treated with moclobemide were less likely than those
treated with imipramine to switch to a manic episode. Similar to
2.4. Nefazodone other MAOIs, moclobemide was shown to be effective in patients
with atypical depression (Sogaard et al., 1999). It has also been
Nefazodone is approved for the treatment of depression in the shown to be effective in the treatment of dysthymia; 57% of the
United States by virtue of its efficacy in numerous placebo- patients responded after 8 weeks of treatment (Versiani et al.,
controlled trials. Similar to trazodone, it has potent selective 1997). Side effects include dry mouth and tachycardia, but the drug
5HT2 antagonistic properties and weakly inhibits the reuptake of appears to be well tolerated in the majority of patients (Tanghe
serotonin and norepinephrine (Feiger et al., 1999). A study that et al., 1997).
assessed nefazodone, psychotherapy, as well as a combination of
psychotherapy and nefazodone in chronically depressed patients 3.3. Reboxetine
revealed response rates of 55% in the nefazodone group, 52% in the
psychotherapy group, and 85% in the combined-treatment group Reboxetine is a potent and selective inhibitor of the norepi-
(Keller et al., 2000). Side effects include nausea, headache, dizziness, nephrine transporter, markedly inhibiting norepinephrine reup-
asthenia, dry mouth, constipation and somnolence (Schatzberg and take. It has been approved for use in several European countries
Nemeroff, 2009; Baldwin et al., 2001). More serious rare adverse including the United Kingdom and Germany since 1997. Several
events include hepatotoxicity (Park and Ishino, 2013), which may meta-analyses have shown it to be effective in the treatment of
be fatal and resulted in an FDA black box warning. major depression. Papakostas et al. (2008) reported similar
response rates between SSRIs (63.9%) and reboxetine (59.2%). A
2.5. Vortioxetine 2013 Cochrane review suggested that reboxetine is less tolerable
than SSRIs (fluoxetine) and had a higher dropout rate in clinical
Vortioxetine, a serotonin reuptake inhibitor, as well as an trials (Magni et al., 2013). Remarkably, a 2010 meta-analysis
antagonist at 5HT1D, 5HT3A and 5HT7 receptors and a partial combined both published and unpublished data and questioned its

Please cite this article in press as: Block, S.G., Nemeroff, C.B., Emerging antidepressants to treat major depressive disorder. Asian J.
Psychiatry (2014), http://dx.doi.org/10.1016/j.ajp.2014.09.001
G Model
AJP-644; No. of Pages 10

4 S.G. Block, C.B. Nemeroff / Asian Journal of Psychiatry xxx (2014) xxx–xxx

efficacy. Analysis of 13 trials that were placebo controlled, SSRI Although ketamine has been shown in small placebo-
controlled, or both, revealed no significant difference in remission controlled trials and meta-analyses to improve symptoms of
rates between placebo and drug (odds ratio 1.17, 95% confidence depression short-term, some concerns regarding the drug
interval 0.91 to 1.51; P = 0.216). Additionally, the published include: safety, optimal dose, route of delivery, its effects on
literature, the authors suggested, overestimated the effect of drug perception and cognition (likely secondary to NMDA and AMPA’s
by 115% compared to placebo, suggesting significant bias (Eyding involvement in cognition), and duration of its antidepressant
et al., 2010). Reboxetine’s major side effects are constipation, effect (Mathew et al., 2008; Sanacora et al., 2012). Adverse effects
difficulty with urination in men, tachycardia and insomnia. may be serious including damage to the kidneys, liver, and
neurotoxicity. There is significant potential for abuse and not all
3.4. Agomelatine patients are responders. Some non-responders carry the Val66-
Met single nucleotide polymorphism (SNP) associated with the
Because depression is associated with alterations in circadian brain derived neurotrophic factor (BDNF) gene (Murrough et al.,
rhythm and sleep disruption, agents that modify biological 2013). Schatzberg (2014) has recently provided a thoughtful
rhythms have been suggested as novel antidepressants. Agome- commentary on the extant database, the lack of true ‘‘blindness’’
latine is an agonist at melatonin MT(1) and MT(2) receptors and an in the clinical trials and the concerns of creating a new epidemic of
antagonist at serotonin 5-HT2C receptors. It is approved for the drug abuse. Despite the problems with the drug, the NMDA
treatment of depression in Europe. Agomelatine has been shown to receptor shows promise as a novel target for antidepressant
improve sleep latency and sleep efficiency over a six week period development. Therefore other antagonists are under investigation
compared to sertraline (Kasper et al., 2010). A recent meta-analysis (including memantine, AZD6765, traxopridil, MK/0657, GLYX-13,
suggests that agomelatine was comparable in efficacy to other riluzole, RO4917523, and N-acetylcysteine) some of which will be
antidepressants including SSRIs and venlafaxine (Guaiana et al., described below. Memantine exhibited no antidepressant effect
2013). However, agomelatine was superior to placebo in the acute in a large double-blind trial despite its similar mechanism of
treatment of depression (nine studies) (Koesters et al., 2013). action to ketamine (Zarate et al., 2006). Whether this is because of
Another meta-analysis involving six studies and a total of 1997 the higher affinity of ketamine for the NMDA receptor, or another
patients (1001 agomelatine; 996 SSRI/SNRI) found a significant yet obscure pharmacological effect or the fact that it is
advantage for agomelatine compared to SSRIs and SNRIs in administered intravenously, is unknown and requires further
HAM-D total scores and the Clinical Global Impression- research (Mathew et al., 2008). CP-101,606, which blocks the
Improvement scale (Kasper et al., 2013). Relapse prevention is receptor-gated ion channel, was studied in 30 patients that were
an important advantage of antidepressants. During a 6-month randomized in a parallel study to receive paroxetine titrated up to
period, relapse was significantly lower in patients on continued 40 mg with either IV placebo or IV treatment medication. Patients
agomelatine therapy compared to placebo (Goodwin et al., were found to have a greater decrease in depression with active
2009). Additionally, this drug has less sexual side effects than drug (60% response rate) from baseline compared to placebo (20%
SSRIs (Montejo et al., 2010). However, problematic are the response rate) (Preskorn et al., 2008). Additionally, MK-0657, a
transient elevations in liver enzymes, which is one reason why selective NR2B antagonist, is being studied in patients who have
agomelatine is not approved for use in the US. Nausea and failed two other antidepressant treatment modalities (clinical
vomiting are also less of a problem with agomelatine compared trials 2008). It has been analyzed in terms of efficacy for the motor
to SSRIs (Guaiana et al., 2013). symptoms of Parkinson’s disease, with no benefit, and has been
shown to cause blood-pressure variability (Addy et al., 2009).
Independent of its efficacy in treating depression, there are
4. N-methyl-D-aspartate (NMDA) receptors and depression concerns with its safety profile. In summary, although there are
several drugs that target the NMDA receptor, much additional
Glutamate, a neurotransmitter present in more than 80% of research is necessary.
neurons, binds to NMDA, AMPA and kainate receptors. There is
evidence for a role of glutamate in the pathophysiology of 5. The HPA axis
depression (Mathews et al., 2012). Ketamine, an NMDA receptor
antagonist has been used as an anesthetic for more than 5.1. Monotherapy targets of the hypothalamic pituitary axis
40 years. It was originally created as a similar but safer drug to
phencyclidine (Howland, 2013). Murrough et al. (2013) assessed Corticotropin releasing factor (CRF) is secreted from the
the effect of intravenously administered ketamine on patients hypothalamus and enters the hypothalamic–hypophyseal portal
with treatment resistant depression; 64% of patients responded, circulation where it stimulates the release of ACTH from the
with one-third (NNT = 2.8) responding specifically to ketamine. anterior pituitary gland. ACTH in turn causes the release of cortisol
However approximately half of those treated relapsed within (Schatzberg and Nemeroff, 2009; Currier and Nemeroff, 2010). A
one week of treatment (Murrough et al., 2013). Additionally, 17% substantial number (40–60%) of severely depressed patients have
of those treated with ketamine had dissociative symptoms. elevated cortisol levels and a disruption of the normal diurnal
Ketamine has also been reported to be effective for bipolar cortisol rhythm. In addition to the elevated cortisol levels,
depression. In a double blind, placebo controlled, crossover study, depressed patients also have elevated concentrations of CRF in
depression improved 40 min after infusion. 71% responded to cerebrospinal fluid (CSF) (Nemeroff et al., 1984) much derived
ketamine compared to 6% to placebo. Ketamine’s effect was from extra-hypothalamic sources. Thus, any strategy that reduces
greatest at day 2 (Diazgranados et al., 2010). Shiroma et al. (2014) HPA axis activity has been suggested to possess promise as a novel
reported that infusions of ketamine over a 12-day period resulted antidepressant approach. Therefore, several anti-glucocorticoids
in a 91.6% response rate and 66.6% remission rate in 12 patients (metyrapone, aminogluthemide, mifepristone, ketoconazole)
with treatment resistant depression. This was an open study with have been studied (Schatzberg and Nemeroff, 2009; Kling et al.,
no placebo or active comparison group. Recently, Naughton et al. 2009).
(2014) have reviewed the literature and suggested that there was A Cochrane review included nine randomized controlled trials
evidence for efficacy but needed additional study (Naughton et al., that studied antiglucocorticoid (mifepristone, ketoconazole,
2014). metyrapone or DHEA) treatment of depression. Five trials assessed

Please cite this article in press as: Block, S.G., Nemeroff, C.B., Emerging antidepressants to treat major depressive disorder. Asian J.
Psychiatry (2014), http://dx.doi.org/10.1016/j.ajp.2014.09.001
G Model
AJP-644; No. of Pages 10

S.G. Block, C.B. Nemeroff / Asian Journal of Psychiatry xxx (2014) xxx–xxx 5

non-psychotic depression, and there was a significant difference in responsive to antidepressant treatment (Musselman et al.,
favor of antiglucocorticoid treatment. Psychotic major depression 2001).
was not significantly affected by antiglucocorticoid treatment. Raison et al. (2013) studied the pharmacological inhibition of
Although a study with mifepristone using plasma concentrations TNF-a with infliximab and whether it possesses antidepressant
to assess drug exposure have been positive (Blasey et al., 2009), properties in SSRI-refractory patients. There was no significant
several studies with CRH R1 receptor antagonists as monotherapy overall difference in the change of HAM-D scores with infliximab
have failed (Holtzheimer and Nemeroff, 2008). versus placebo. However, there was a significant interaction when
assessing treatment, time, and log baseline CRP concentrations. In
5.2. Augmentation therapy with hypothalamic–pituitary–adrenal fact, CRP concentrations greater than 5 mg/L at baseline was
axis targets associated with a positive treatment response (50% reduction in
HAM-D score at any point during treatment) of 62% (8 of 13
Additionally, medications that affect the HPA axis have been patients) with infliximab compared to 33% (3 of 9 patients) in
studied as augmentation agents in the treatment of depression. placebo-treated patients (Raison et al., 2013). Likewise, the TNF
Patients treated with either nefazodone or fluvoxamine received antagonist, etanercept, previously shown to possess antidepres-
either placebo or metyrapone. More patients responded to sant effects in patients with psoriasis (Tyring et al., 2006) has
treatment when metyrapone was added ((day 21: 23 of 33 recently been shown to possess anti-anxiety and antidepressant
patients) and day 35: 19 of 33 patients) compared with placebo effects in a rat model. Thus, rats treated with etanercept exhibited
(day 21: 13 of 30 patients; Fisher exact P = 0.031; day 35: 10 of 30 decreased immobility time in the rat model of despair (Bayr-
patients; Fisher exact P = 0.047). Additionally, patients treated amgurler et al., 2013). Another study assessed celecoxib, a Cox2
with drug, had a faster response starting in the first week (Kaplan- inhibitor, as an augmenting agent to reboxetine. Superior results
Meier analysis; log-rank test P < 0.006) (Jahn et al., 2004). were found using celecoxib compared to placebo (Currier and
Remarkably, no studies using CRHR1 antagonists added to SSRI Nemeroff, 2010).
non-responders have been conducted.
8. Atypical antipsychotics (olanzapine, quetiapine,
6. Omega-3 fatty acids aripiprazole, and risperidone)

Studies have suggested that cultures with diets higher in Although antipsychotics primarily target the dopamine D2
omega-3 fatty acids have lower rates of depression, whereas receptor, second-generation antipsychotics also target the seroto-
cultures with diets lower in omega-6 fatty acids have higher rates nin 5HT1A and 5HT2 receptors. Through these and other
of depression. Meta-analyses studying omega-3 fatty acids and mechanisms, atypical antipsychotics have been postulated to
their efficacy on depression have been questioned because of bias possess antidepressant properties. The largest meta-analysis of
and lack of consistency between studies. Additionally, some second-generation antipsychotics for major depressive disorder
smaller studies have found no significant difference between and dysthymia found that quetiapine monotherapy was more
omega-3 fatty acid treatment and placebo (Rakofsky et al., 2009). effective than placebo in treating depression symptomatology,
In 13 randomized, placebo-controlled trials studying omega-3 whereas olanzapine, compared to placebo had no benefit in
fatty acids (n = 731 participants), a meta-analysis of major depressive symptom treatment of psychotic depression (n = 184,
depressive disorder showed no significant benefit of omega-3 2 RCTs, MD 2.86; 95% CI 6.19 to 0.47). Risperidone has not been
fatty acid treatment compared to placebo. Additionally, there studied as monotherapy for the treatment of depression (Komossa
was heterogeneity and publication bias between studies (Bloch et al., 2010).
and Hannestad, 2012). Freeman and Rapaport (2011) has
discussed this heterogeneity citing the short duration, varying 8.1. Lithium
dosages, and fatty acids being adjunctive to other medications as
confounding variables. Some suggest genetic polymorphisms In 1981, lithium was demonstrated to be effective as an
(FADS1 and FADS2) may influence the rate at which omega-3 augmenting agent in the treatment of depression in patients who
fatty acids are metabolized and that these genes can predict failed to respond to TCAs. Although better studied in patients
response to omega-3 fatty acids in the treatment of depression. treated with TCAs, meta-analyses have shown lithium to be
Additionally, there is evidence that high ratios of EPA:DHA may effective when combined with SSRIs as well. In a meta-analysis of
be associated with response to omega-3 fatty acids in major 9 placebo controlled trials, lithium was effective as an
depressive disorder and bipolar depression. The lack of consis- augmenting agent for conventional antidepressant therapy
tent results between studies, the small number of studies, compared to placebo (OR = 3.31) (Bauer and Dopfmer, 1999).
and the fact that the studies themselves have a small number Another, more recent meta-analysis included 10 randomized
of subjects suggest that further research is necessary before placebo-controlled trials and showed a higher response rate with
definitive recommendations can be made (Freeman and Rapa- lithium augmentation (41.2%) compared to placebo (14.4%)
port, 2011; Freeman, 2009). (OR = 3.11, NNT = 5). Additionally, there have been suggestions
that patients with the —50T/C SNP of the GSK3 gene may be more
7. Inflammation responsive to lithium augmentation (Bauer et al., 2010). Part of
the difficulty prescribing lithium is the need to titrate its dosage,
There is considerable evidence of a link between inflamma- the need for monitoring of blood levels and its side effect profile.
tory cytokines and depression. TNF, IL-1, and IL-6 have all been It often acts quickly but may take up to six weeks to become
shown to be elevated in plasma of depressed patients and effective. Dosages start between 600 and 900 mg/day with a
associated with nonresponse to conventional antidepressants. blood level of at least 0.5 mEq/L required to be effective (Chang
Therefore, anti-inflammatory agents have been hypothesized to et al., 2013). The drug has a long-term effect on renal and thyroid
represent a novel treatment for depression. Interferon-a function and these should be monitored regularly. However, of
increases TNF, IL-1, and IL-6 plasma concentrations and 40– all of the augmentation strategies, lithium has perhaps the
50% of interferon-a treated patients exhibit depressive symp- strongest scientific evidence as an augmentation agent in the
toms and develop syndromal major depressive disorder which is treatment of depression.

Please cite this article in press as: Block, S.G., Nemeroff, C.B., Emerging antidepressants to treat major depressive disorder. Asian J.
Psychiatry (2014), http://dx.doi.org/10.1016/j.ajp.2014.09.001
G Model
AJP-644; No. of Pages 10

6 S.G. Block, C.B. Nemeroff / Asian Journal of Psychiatry xxx (2014) xxx–xxx

8.2. Thyroid hormone augmentation randomized controlled trials to augment the effects of SSRIs
in MDD.
Several studies have assessed the use of tri-iodothyronine (T3)
as an augmentation agent to TCAs and SSRIs for the treatment of 8.5. Atypical antipsychotics as augmenting agents
depression. A 2009 meta-analysis with 444 patients combined
results of 4 studies in which T3 combined with SSRIs was not As noted above, atypical antipsychotics have been studied in
superior in efficacy to SSRI monotherapy. However, these patients depression both as monotherapy and as augmentation agents.
did not fulfill criteria for treatment resistant depression (TRD) Aripiprazole was the first of the second generation antipsychotics
(Papakostas et al., 2009). The STAR*D trial, in contrast, compared to obtain approval by the FDA for antidepressant augmentation.
lithium and thyroid hormone as augmenting agents for patients Since then, both quetiapine XR and olanzapine, when added to
who failed 2 treatment regimens. Thyroid hormone was superior fluoxetine, have been approved for the acute treatment of
to lithium, achieving remission in 24.7% of patients compared to depression (Kato and Chang, 2013b). Two major meta-analyses
15.9% treated with lithium (Nierenberg et al., 2006). Our more have addressed the efficacy of second generation antipsychotics as
recent trial sought to determine whether T3 accelerated or augmentation therapy with antidepressants (Nelson and Papa-
enhanced the effect of sertraline in 153 patients without treatment kostas, 2009; Komossa et al., 2010). Nelson and Papakostas
refractory depression (Garlow et al., 2012). There was no added included 16 studies and calculated a pooled OR for response rate of
benefit of thyroid hormone to sertraline, perhaps because the 1.69 [95% confidence interval (CI) 1.46 to 1.95, P < 0.00001], with a
sertraline dosage was increased to achieve optimal benefit. There NNT equal to nine. For all antipsychotics studied there was an
are few side effects of T3 in the doses utilized (25–50 mg/day). improvement in response rate: Aripiprazole OR = 2.07, olanzapine
OR = 1.39, quetiapine OR = 1.60, and risperidone OR = 1.83. Side
8.3. Buspirone augmentation effects of second-generation antipsychotics include sedation,
dizziness, insomnia, akathisia, weight gain, Q–T prolongation
Buspirone acts as a partial agonist at the 5HT1A receptor and and metabolic syndrome.
has been studied as both monotherapy and augmentation therapy
in the treatment of depression. In the STAR-D study, patients with 8.6. Other augmenting strategies: Pindolol, yohimbine, testosterone/
non-psychotic major depression treated with citalopram who did estrogens, and lamotrigine
not achieve remission were augmented with buspirone or
bupropion. HRSD-17 remission was achieved in 30.1% of patients 8.6.1. Pindolol
treated with buspirone (Trivedi et al., 2006). Despite the Pindolol, a b-adrenoreceptor and 5HT1A somatodendritic
improvement in depression, buspirone had a higher dropout rate autoreceptor, antagonist has been studied as an augmenting agent
than bupropion. Other studies have suggested that although in depression. Patients treated with pindolol and citalopram had a
buspirone may accelerate the antidepressant response resulting in higher likelihood of continued remission (OR = 5.00, 95% CI 1.191–
a significantly greater reduction in the MADRS score after 1 week of 20.989) and exhibited an earlier response, 65% less (22 days vs. 30
treatment, there is no overall effect on depression compared to days) compared to placebo (Portella et al., 2011). These results
placebo (Appelberg et al., 2001). Side effects of buspirone include were largely confirmed with meta-analysis (Portella et al., 2011)
gastrointestinal symptoms, dizziness, insomnia, palpitations, though negative findings of pindolol combined with paroxetine
paresthesias and sweating (Kishi et al., 2013). have been reported (Geretsegger et al., 2008).

8.4. Stimulant augmentation 8.6.2. Yohimbine


Because a-2 adrenoreceptor blockade increases the firing rate
Anhedonia, the inability to experience pleasure, is a cardinal of 5HT neurons, this drug has been posited to augment the action of
feature of depression. The main neurotransmitter involved in the antidepressants. Patients (n = 25/group) were assigned to receive
brain’s reward system is dopamine (DA). Stimulants, approved for either fluoxetine 20 mg plus placebo (F/P) or fluoxetine 20 mg plus
the treatment of attention deficit disorder, increase the release of yohimbine (F/Y), the latter titrated based on blood pressure
DA and 5HT as well as weakly block the reuptake of these amines. changes for 6 weeks. Clinical global impression ratings and the
For this reason, researchers have suggested that stimulants may Hamilton depression rating scales were used to monitor treatment
have some utility in the treatment of depression. In one study, response. Response was achieved faster with F/Y compared to F/P
patients who continued to have depressive symptoms after and the MDD patients had higher response rates; 18/26 (69%) vs.
therapy were assigned to receive either placebo or OROS 10/24 (42%) in the yohimbine group using the CGI79. This suggests
methylphenidate. There was no significant effect on MADRS that a-2 antagonism may serve some utility for the treatment of
scores. However, OROS methylphenidate was more effective than depression. Surprisingly, other a-2 antagonists such as dazoxan,
placebo in improving apathy and fatigue (Ravindran et al., 2008). have not been studied in this area.
Two large randomized controlled trials of lisdexamfetamine in the
treatment of MDD in SSRI non-remitters revealed no efficacy of the 8.6.3. Testosterone/estrogens
psychostimulant (Shrie, 2014). A double blind, placebo-controlled In a pioneering study, Pope et al. (2010) reported that men with
study assessed adjunctive modafinil or placebo (to an SSRI) and depression who had undiagnosed hypogonadism, when treated
found that modafinil did not have any benefit (Dunlop et al., 2007). with testosterone, experienced a clear improvement of
A Cochrane review published in 2008 included 13 trials for meta- depression. A meta-analysis of seven studies found a significant
analysis. Three trials (62 participants) suggested that psychosti- positive effect of testosterone therapy on HAM-D response when
mulant monotherapy significantly decreased depressive symp- compared with placebo (z = 4.04, P < 0.0001) (Zarrouf et al., 2009).
toms compared to placebo. However, modafinil and placebo were There is clear evidence for a role for estrogen deficiency in
not different. Additionally, the literature is primarily comprised of depression especially in the premenopausal and postmenopausal
small trials over short time periods. Atomoxetine, a norepineph- period (Soares, 2008; Soares and Frey, 2010). Indeed our group was
rine (NE) reuptake inhibitor, used in the treatment ADHD, also has involved in one of the first studies demonstrating a higher response
shown no benefit in augmentation of sertraline. Similarly, rate to fluoxetine in postmenopausal depressed women on
edivoxetine, another NE reuptake inhibitor, failed in three hormone replacement therapy (HRT) (Schneider et al., 1997).

Please cite this article in press as: Block, S.G., Nemeroff, C.B., Emerging antidepressants to treat major depressive disorder. Asian J.
Psychiatry (2014), http://dx.doi.org/10.1016/j.ajp.2014.09.001
G Model
AJP-644; No. of Pages 10

S.G. Block, C.B. Nemeroff / Asian Journal of Psychiatry xxx (2014) xxx–xxx 7

There is also evidence that estrogen may have a role in the 11. Focal brain stimulation: Vagus nerve stimulation,
treatment of postpartum depression (Gregoire et al., 1996). transcranial magnetic stimulation, magnetic seizure therapy,
deep brain stimulation
8.6.4. Lamotrigine
An inhibitor of voltage sensitive sodium channels, lamotrigine 11.1. Vagus nerve stimulation
has recently been approved for the prevention of episodes of
bipolar depression. There is no evidence that it is effective in the Vagus nerve stimulation (VNS) was first performed in the 1980s
treatment of unipolar MDD or in the acute management of mania and was FDA-approved for the treatment of pharmaco-resistant
(Schatzberg and Nemeroff, 2009). epilepsy. Patients with epilepsy treated with VNS were noted to
exhibit improvements in mood. It was approved by the FDA in
8.6.5. Mianserin, mirtazapine, and desipramine: Combination therapy 2005 for the treatment of depression in patients over the age of
A combination approach to antidepressant therapy in non- 18 who failed two or more antidepressant treatment regimens. An
responders has been suggested on the basis of combining electrode is surgically attached to the left vagus nerve and a
antidepressants with different mechanisms of action. There are generator that delivers a pulse is implanted into the chest wall.
two methods for this approach. Either two antidepressants can be Side effects include hoarseness, difficulty swallowing, or cough.
started on initiation of treatment, or a second agent can be added Part of the criticism of the vagus nerve stimulation data is the lack
after the first was unsuccessful in achieving remission. Lopes of RCTs. In fact, a 10 week randomized controlled trial including
Rocha et al. (2013) attempted to perform a meta-analysis on trials 235 patients did not show a significant effect for vagus nerve
using dual antidepressant therapy (Sanacora et al., 2004). They stimulation compared to placebo (OR = 1.61; P = 0.25) (Rush et al.,
reported only two small studies that suggested a benefit of treating 2005). However, a meta-analysis conducted in 2012 showed a
depression with an antidepressant combination (Ferreri et al., significant reduction in Hamilton Depression Rating Scale scores,
2001; Carpenter et al., 2002). The notion of combination therapy with 31.8% of subjects becoming responders ([23.2% to 41.8%],
was supported by Nelson et al. (2004) when his team found that P < 0.001) (Martin and Martin-Sanchez, 2012).
desipramine (dosed according to plasma adjusted level) plus
fluoxetine (20 mg/day) was more effective than either drug alone 11.2. Transcranial magnetic stimulation (TMS)
(53.8%, 0% and 7.1% remission rates, respectively). Blier et al.
published two supportive studies. One assessed mirtazapine Transcranial magnetic stimulation (TMS) is FDA approved for
(30 mg/day), paroxetine (20 mg/day), or a combination of both depression in patients who have failed to respond to one, but not
mirtazapine and paroxetine. At day 42, there was a 10 point more than one, antidepressant trial. It is the least invasive of the
difference in MADRS score when comparing monotherapy and brain stimulation techniques and most frequently targets the
combination therapy, favoring combination therapy (Blier et al., dorsolateral prefrontal cortex (DLPFC). A magnetic field is
2009). A later study confirmed these results. Remission rates were produced around the brain and an electrical current is delivered
higher with combination therapy versus fluoxetine monotherapy through the cortex. Frequencies vary from 1 to 50 Hz with low-
(Blier et al., 2010). A large trial assessed the effect of monotherapy frequency pulses causing less excitability in the cortex than higher
or combination therapy for depression in patients with heart frequency pulsations. This form of treatment has been shown in
disease. Neither depressed patients with heart disease or those several meta-analyses to be as effective as both an augmenting
without heart disease exhibited any differences in response to agent and monotherapy for the treatment of major depression. A
antidepressant monotherapy versus combination therapy (Kerber 2013 meta-analysis included 8 randomized controlled trials
et al., 2012). comprised of 263 patients. After a mean of 12.6  3.9 rTMS
sessions, 38.2% and 15.1% of subjects receiving active LF-rTMS and
sham rTMS were responders, respectively (OR = 3.35) whereas 34.6%
9. Substance P
and 9.7% of subjects were classified as remitters (OR = 4.76) (Berlim
et al., 2013). However, different meta-analyses have produced
Substance P is a CNS neuropeptide neurotransmitter that acts
conflicting results as to whether active rTMS is more effective than
primarily on neurokinin (NK) receptors. Although an early study
sham rTMS. Most randomized controlled trials found active rTMS to
(Kramer et al., 1998) suggested antidepressant efficacy of an NK1
be more effective than sham rTMS. Problematic is the lack of
antagonist, MK869 (aprepitant) later double blind studies failed to
standardization of rTMS parameters with frequencies ranging
confirm those findings (Keller et al., 2006).
between 0.17 Hz and 40 Hz, treatments lasting between 5 and
20 days (magnetic stimulation requires daily treatment), and
10. Electroconvulsive therapy between 30 and 4800 pulses applied per treatment session. For this
reason, clear guidelines are necessary.
Electroconvulsive therapy, usually used in cases of medication Deep brain TMS (DTMS) over the prefrontal cortex has been
refractory depression, induces a seizure in the patient and is widely shown to help apathy and depression. Higher intensity stimulation
believed to be the most effective antidepressant treatment is likely more effective than lower intensity stimulation and
(Matthews et al., 2013; Kellner et al., 2006). In fact, it is considered baseline apathy may help predict whether or not DTMS can lead to
first-line for major depressive disorder with severe suicidal remission (Levkovitz et al., 2009, 2011). Studies have also shown
ideation or severe depression with psychotic features because of that continuation of H-coil rTMS may help maintain an antide-
its repeatedly proven efficacy and fast onset of action. Although in pressant effect following the acute treatment phase (Harel et al.,
almost all cases it is utilized after failure of one or more 2014). Side effects are relatively mild and include headaches,
antidepressant medications, a meta-analysis showed that the dizziness, and scalp discomfort (Hovington et al., 2013).
overall remission rate was 48.0% (281/585) for patients with and
64.9% (242/373) for patients without prior pharmacotherapy 11.3. Theta burst stimulation
failure. The requirement for general anesthesia and a muscle
paralyzing agent prevents an overt tonic–clonic seizure. ECT’s Theta burst stimulation (TBS) is a form of TMS that affects
memory impairment is a major drawback, as well as the stigma synaptic plasticity rapidly and can produce long-lasting effects on
associated with its use. corticospinal excitability. Whereas intermittent TBS causes a

Please cite this article in press as: Block, S.G., Nemeroff, C.B., Emerging antidepressants to treat major depressive disorder. Asian J.
Psychiatry (2014), http://dx.doi.org/10.1016/j.ajp.2014.09.001
G Model
AJP-644; No. of Pages 10

8 S.G. Block, C.B. Nemeroff / Asian Journal of Psychiatry xxx (2014) xxx–xxx

potentiation-like effect on cortical excitability, continuous TBS g = 0.743, 95% CI 0.21–1.27). However, results between studies
causes a reduction in excitability. A 2014 study compared differed more than explained by chance (Kalu et al., 2012). A
continuous TBS, intermittent TBS, a combination of continuous 2013 study assessed MADRS scores between sham and active
and intermittent TBS and sham TBS. Intermittent TBS combined TDCS groups and found significant benefit in dysphoria for active
with continuous TBS was most effective for treating depression (Li TDCS. Both active and sham tDCS improved vegetative and
et al., 2014). Another study provided intermittent enhancing TBS to retardation symptoms (Alonzo et al., 2013). Further research is
the left dorsolateral prefrontal cortex (dlPFC) and inhibiting needed.
constant TBS (cTBS) to the right dlPFC, or bilateral sham
stimulation. Bilateral TBS was a safe, more effective form of 12. Conclusion
treating MDD than bilateral sham stimulation (OR = 3.86) (Plewnia
et al., 2014). While there are many effective treatments for depression, many
patients continue to suffer with unrelenting severe depressive
11.4. Magnetic seizure therapy symptoms. Unfortunately, a trial and error approach to therapy has
become the norm, rather than the exception. Some guidelines exist
Magnetic seizure therapy (MST), not FDA approved, uses a to help physicians optimize therapy, but no significant recent
device that causes a series of magnetic pulses to be delivered to the breakthroughs have been reported. However, as personalized
patient while the subject is under general anesthesia. It can be medicine is realized with its focus on genomics and imaging, a
thought of as a hybrid of TMS and ECT in that it combines patient-centered protocol tailored to each individual is an
the antidepressant effects of both therapies without causing the attainable goal.
cognitive impairment characteristic of ECT. Compared to ECT, MST
is more focal, stimulating primarily only cortical gray matter, as 13. Expert opinion
opposed to both gray and white matter. Additionally, MST differs
from ECT in how it propagates the seizure. However, there is no For years, the treatment of depression has focused on drugs that
dosing algorithm or suggested coil configuration at this time. In a act on monoamine neurotransmitter systems. Recent develop-
2013 pilot study of 13 patients who received 18 treatments with ments in the field have addressed different neural pathways
100-Hz MST, five patients met clinical response criteria, without involved in the etiology of depression. Despite these advances, few
cognitive side effects (Fitzgerald et al., 2013). Trials are still limited breakthroughs have been realized. As in other branches of
and relatively small; therefore, further research is required in this medicine, an evidence-based, patient tailored approach must be
area (McClintock et al., 2011). utilized. A better understanding of the pathophysiology of
depression is a prerequisite for further advances in the field.
11.5. Deep brain stimulation (DBS) Currently available antidepressants and augmenting agents are
clearly effective in the treatment of the majority of depressed
Deep brain stimulation is not FDA approved for the treatment of patients.
refractory depression. The procedure involves implanting electro-
References
des into the brain bilaterally and delivering a pulse that alters the
activity of targeted circuits. It is approved for Parkinson’s disease Addy, C., Assaid, C., Hreniuk, D., et al., 2009. Single-dose administration of MK-0657,
and has replaced ablative therapy because of its reversible nature an NR2B-selective NMDA antagonist, does not result in clinically meaningful
and therapeutic efficacy (Mayberg, 2007). Lozano et al. (2012) improvement in motor function in patients with moderate Parkinson’s disease.
J. Clin. Pharmacol. 49 (7) 856–864.
studied 21 patients with subcallosal cingulate gyrus (SCG) Alonzo, A., Chan, G., Martin, D., Mitchell, P.B., Loo, C., 2013. Transcranial direct
Brodman’s area 25 deep brain stimulation and found a 57% current stimulation (tDCS) for depression: analysis of response using a
response rate at 1 month, 48% at 6 months, and 29% at 12 three-factor structure of the Montgomery-Åsberg depression rating scale.
J. Affect. Disord. 150 (Aug. (1)) 91–95.
months. Following completion of a one-year study of DBSadmi- Appelberg, B.G., Syvalahti, E.K., Koskinen, T.E., Mehtonen, O.P., Muhonen, T.T.,
nistered bilaterally to the SCG, 20 patients were assessed at Naukkarinen, H.H., 2001. Patients with severe depression may benefit from
baseline, year 1, year 2 and year 3; 62.5%, 46.2%, and 75.0% buspirone augmentation of selective serotonin reuptake inhibitors: results
from a placebo-controlled, randomized, double-blind, placebo wash-in study.
responded after years 1, 2, and 3, respectively. HAMD scores
J. Clin. Psychiatry 62 (6) 448–452.
decreased significantly at all time points (Schlaepfer et al., 2013). A Baldomero, E.B., Ubago, J.G., Cercos, C.L., Ruiloba, J.V., Calvo, C.G., Lopez, R.P., 2005.
recent industry sponsored trial of SCG directed DBS did not Venlafaxine extended release versus conventional antidepressants in the re-
demonstrate efficacy versus sham controls, a finding similar to an mission of depressive disorders after previous antidepressant failure: ARGOS
study. Depress. Anxiety 22 (2) 68–76.
internal capsule directed DBS study conducted earlier. Seven Baldwin, D.S., Hawley, C.J., Mellors, K., CN104-070 Study Group, 2001. A random-
patients with refractory depression had DBS delivered to the ized, double-blind controlled comparison of nefazodone and paroxetine in the
supero-lateral branch of the medial forebrain bundle; 6/7 achieved treatment of depression: safety, tolerability and efficacy in continuation phase
treatment. J. Psychopharmacol. 15 (3) 161–165.
response criteria 7 days after stimulation with 6/7 continuing to Bauer, M., Dopfmer, S., 1999. Lithium augmentation in treatment-resistant de-
meet response at 6 months and 4/7 met remission (Bewernick pression: meta-analysis of placebo-controlled studies. J. Clin. Psychophar-
et al., 2010). DBS of the nucleus accumbens in 10 patients resulted macol. 19 (5) 427–434.
Bauer, M., Adli, M., Bschor, T., et al., 2010. Lithium’s emerging role in the treatment
in 5 who were responders at 12 months. Compared to those of refractory major depressive episodes: augmentation of antidepressants.
patients who were nonresponders, responders had deceased Neuropsychobiology 62 (1) 36–42.
metabolism in the amygdala. DBS also produced anxiolytic effects Bayramgurler, D., Karson, A., Ozer, C., Utkan, T., 2013. Effects of long-term etaner-
cept treatment on anxiety- and depression-like neurobehaviors in rats.
(Kennedy et al., 2011).
Physiol. Behav. 119, 145–148.
Currier, M.B., Nemeroff, C.B., 2010. Inflammation and mood disorders: proinflam-
11.6. Transcranial direct current stimulation matory cytokines and the pathogenesis of depression. Anti-Inflammatory &
Anti-Allergy Agents Med. Chem. 9 (September (3)) 212–220.
Berlim, M.T., Van den Eynde, F., Jeff Daskalakis, Z., 2013. Clinically meaningful
Transcranial direct current stimulation (tDCS) is a noninva- efficacy and acceptability of low-frequency repetitive transcranial magnetic
sive treatment that modulates cortical excitability and neural stimulation (rTMS) for treating primary major depression: a meta-analysis of
activity. Studies have disputed whether it is an effective randomized, double-blind and sham-controlled trials. Neuropsychopharmacol-
ogy 38 (4) 543–551.
treatment for major depression. A 2012 meta analysis of six Berton, O., Nestler, E.J., 2006. New approaches to antidepressant drug discovery:
RCTs found active tDCS more beneficial than sham tDCS (Hedges’ beyond monoamines. Nat. Rev. Neurosci. 7 (2) 137–151.

Please cite this article in press as: Block, S.G., Nemeroff, C.B., Emerging antidepressants to treat major depressive disorder. Asian J.
Psychiatry (2014), http://dx.doi.org/10.1016/j.ajp.2014.09.001
G Model
AJP-644; No. of Pages 10

S.G. Block, C.B. Nemeroff / Asian Journal of Psychiatry xxx (2014) xxx–xxx 9

Bewernick, B.H., Hurlemann, R., Matusch, A., et al., 2010. Nucleus accumbens deep Hovington, C.L., McGirr, A., Lepage, M., Berlim, M.T., 2013. Repetitive transcranial
brain stimulation decreases ratings of depression and anxiety in treatment- magnetic stimulation (rTMS) for treating major depression and schizophrenia:
resistant depression. Biol. Psychiatry 67 (2) 110–116. a systematic review of recent meta-analyses. Ann. Med. 45 (4) 308–321.
Binder, E.B., Salyakina, D., Lichtner, P., et al., 2004. Polymorphisms in FKBP5 are Howland, R.H., 2013. Ketamine for the treatment of depression. J. Psychosoc. Nurs.
associated with increased recurrence of depressive episodes and rapid response Ment. Health Serv. 51 (1) 11–14.
to antidepressant treatment. Nat. Genet. 36 (12) 1319–1325. Jahn, H., Schick, M., Kiefer, F., Kellner, M., Yassouridis, A., Wiedemann, K., 2004.
Blasey, C.M., Debattista, C., Roe, R., Block, T., Belanoff, J.K., 2009. A multisite trial of Metyrapone as additive treatment in major depression: a double-blind and
mifepristone for the treatment of psychotic depression: a site-by-treatment placebo-controlled trial. Arch. Gen. Psychiatry 61 (12) 1235–1244.
interaction. Contemp. Clin. Trials 30 (4) 284–288. Kalu, U.G., Sexton, C.E., Loo, C.K., Ebmeier, K.P., 2012. Transcranial direct current
Blier, P., Gobbi, G., Turcotte, J.E., et al., 2009. Mirtazapine and paroxetine in major stimulation in the treatment of major depression: a meta-analysis.
depression: a comparison of monotherapy versus their combination from Psychol. Med. 42 (Sep. (9)) 1791–1800.
treatment initiation. Eur. Neuropsychopharmacol. 19 (7) 457–465. Kasper, S., McEwen, B.S., 2008. Neurobiological and clinical effects of the antide-
Blier, P., Ward, H.E., Tremblay, P., Laberge, L., Hebert, C., Bergeron, R., 2010. pressant tianeptine. CNS Drugs 22 (1) 15–26.
Combination of antidepressant medications from treatment initiation for major Kasper, S., Hajak, G., Wulff, K., et al., 2010. Efficacy of the novel antidepressant
depressive disorder: a double-blind randomized study. Am. J. Psychiatry 167 agomelatine on the circadian rest-activity cycle and depressive and anxiety
(3) 281–288. symptoms in patients with major depressive disorder: a randomized, double-
Bloch, M.H., Hannestad, J., 2012. Omega-3 fatty acids for the treatment of depres- blind comparison with sertraline. J. Clin. Psychiatry 71 (2) 109–120.
sion: systematic review and meta-analysis. Mol. Psychiatry 17 (12) 1272–1282. Kasper, S., Corruble, E., Hale, A., Lemoine, P., Montgomery, S.A., Quera-Salva, M.A.,
Boulenger, J.P., Loft, H., Olsen, C.K., 2014. Efficacy and safety of vortioxetine (lu 2013. Antidepressant efficacy of agomelatine versus SSRI/SNRI: results from a
AA21004), 15 and 20 mg/day: a randomized, double-blind, placebo-controlled, pooled analysis of head-to-head studies without a placebo control. Int. Clin.
duloxetine-referenced study in the acute treatment of adult patients with major Psychopharmacol. 28 (1) 12–19.
depressive disorder. Int. Clin. Psychopharmacol. 29 (3) 138–149. Kato, M., Chang, C.M., 2013a. Augmentation treatments with second-generation
Carpenter, L.L., Yasmin, S., Price, L.H., 2002. A double-blind, placebo-controlled antipsychotics to antidepressants in treatment-resistant depression. CNS Drugs
study of antidepressant augmentation with mirtazapine. Biol. Psychiatry 51 (2) 27 (Suppl. 1) S11–S19.
183–188. Kato, M., Chang, C.M., 2013b. Augmentation treatments with second-generation
Chang, C.M., Sato, S., Han, C., 2013. Evidence for the benefits of nonantipsychotic antipsychotics to antidepressants in treatment-resistant depression. CNS Drugs
pharmacological augmentation in the treatment of depression. CNS Drugs 27 27 (Suppl 1) S11–S19.
(Suppl. 1) S21–S27. Katona, C., Hansen, T., Olsen, C.K., 2012. A randomized, double-blind, placebo-
Diazgranados, N., Ibrahim, L., Brutsche, N.E., et al., 2010. A randomized add-on trial controlled, duloxetine-referenced, fixed-dose study comparing the efficacy and
of an N-methyl-D-aspartate antagonist in treatment-resistant bipolar depres- safety of lu AA21004 in elderly patients with major depressive disorder. Int.
sion. Arch. Gen. Psychiatry 67 (8) 793–802. Clin. Psychopharmacol. 27 (4) 215–223.
Dunlop, B.W., Crits-Christoph, P., Evans, D.L., et al., 2007. Coadministration of Keller, M.B., McCullough, J.P., Klein, D.N., et al., 2000. A comparison of nefazodone,
modafinil and a selective serotonin reuptake inhibitor from the initiation of the cognitive behavioral-analysis system of psychotherapy, and their combi-
treatment of major depressive disorder with fatigue and sleepiness: a double- nation for the treatment of chronic depression. N. Engl. J. Med. 342 (20) 1462–
blind, placebo-controlled study. J. Clin. Psychopharmacol. 27 (6) 614–619. 1470.
Duval, F., Lebowitz, B.D., Macher, J.P., 2006. Treatments in depression. Dialogues Keller, M., Montgomery, S., Ball, W., et al., 2006. Lack of efficacy of the substance p
Clin. Neurosci. 8 (2) 191–206. (neurokinin1 receptor) antagonist aprepitant in the treatment of major depres-
Eyding, D., Lelgemann, M., Grouven, U., et al., 2010. Reboxetine for acute treatment sive disorder. Biol. Psychiatry 59 (3) 216–223.
of major depression: systematic review and meta-analysis of published and Kellner, C.H., Knapp, R.G., Petrides, G., et al., 2006. Continuation electroconvulsive
unpublished placebo and selective serotonin reuptake inhibitor controlled therapy vs pharmacotherapy for relapse prevention in major depression: a
trials. BMJ 341, c4737. multisite study from the consortium for research in electroconvulsive therapy
Feiger, A.D., Bielski, R.J., Bremner, J., et al., 1999. Double-blind, placebo-substitution (CORE). Arch. Gen. Psychiatry 63 (12) 1337–1344.
study of nefazodone in the prevention of relapse during continuation treatment Kennedy, S.H., Giacobbe, P., Rizvi, S.J., et al., 2011. Deep brain stimulation for
of outpatients with major depression. Int. Clin. Psychopharmacol. 14 (1) 19–28. treatment-resistant depression: follow-up after 3 to 6 years. Am. J. Psychiatry
Ferreri, M., Lavergne, F., Berlin, I., Payan, C., Puech, A.J., 2001. Benefits from 168 (5) 502–510.
mianserin augmentation of fluoxetine in patients with major depression Kerber, K.B., Wisniewski, S.R., Luther, J.F., et al., 2012. Effects of heart disease on
non-responders to fluoxetine alone. Acta Psychiatr. Scand. 103 (1) 66–72. depression treatment: results from the COMED study. Gen. Hosp. Psychiatry 34
Fitzgerald, P.B., Hoy, K.E., Herring, S.E., Clinton, A.M., Downey, G., Daskalakis, Z.J., (1) 24–34.
2013. Pilot study of the clinical and cognitive effects of high-frequency mag- Kishi, T., Meltzer, H.Y., Iwata, N., 2013. Augmentation of antipsychotic drug action
netic seizure therapy in major depressive disorder. Depress. Anxiety 30 (2) by azapirone 5-HT1A receptor partial agonists: a meta-analysis. Int. J. Neurop-
129–136. sychopharmacol. 16 (6) 1259–1266.
Freeman, M.P., Rapaport, M.H., 2011. Omega-3 fatty acids and depression: from Kling, M.A., Coleman, V.H., Schulkin, J., 2009. Glucocorticoid inhibition in the
cellular mechanisms to clinical care. J. Clin. Psychiatry 72 (2) 258–259. treatment of depression: can we think outside the endocrine hypothalamus?
Freeman, M.P., 2009. Omega-3 fatty acids in major depressive disorder. J. Clin. Depress. Anxiety 26 (7) 641–649.
Psychiatry 70 (Suppl. 5) 7–11. Koesters, M., Guaiana, G., Cipriani, A., Becker, T., Barbui, C., 2013. Agomelatine
Garlow, S.J., Dunlop, B.W., Ninan, P.T., Nemeroff, C.B., 2012. The combination of efficacy and acceptability revisited: systematic review and meta-analysis
triiodothyronine (T3) and sertraline is not superior to sertraline monotherapy of published and unpublished randomised trials. Br. J. Psychiatry 203 (3)
in the treatment of major depressive disorder. J. Psychiatr. Res. 46 (11) 1406– 179–187.
1413. Komossa, K., Depping, A.M., Gaudchau, A., Kissling, W., Leucht, S., 2010. Second-
Geretsegger, C., Bitterlich, W., Stelzig, R., Stuppaeck, C., Bondy, B., Aichhorn, W., generation antipsychotics for major depressive disorder and dysthymia.
2008. Paroxetine with pindolol augmentation: a double-blind, randomized, Cochrane Database Syst. Rev. 12, CD008121 (12):CD008121.
placebo-controlled study in depressed in-patients. Eur. Neuropsychopharma- Kramer, M.S., Cutler, N., Feighner, J., et al., 1998. Distinct mechanism for antide-
col. 18 (2) 141–146. pressant activity by blockade of central substance P receptors. Science 281
Goodwin, G.M., Emsley, R., Rembry, S., Rouillon, F., Agomelatine Study Group, (5383) 1640–1645.
2009. Agomelatine prevents relapse in patients with major depressive dis- Kriston, L., von Wolff, A., Westphal, A., Holzel, L.P., Harter, M., 2014. Efficacy and
order without evidence of a discontinuation syndrome: a 24-week random- acceptability of acute treatments for persistent depressive disorder: a network
ized, double-blind, placebo-controlled trial. J. Clin. Psychiatry 70 (8) 1128– meta-analysis. Depress. Anxiety.
1137. Levkovitz, Y., Harel, E.V., Roth, Y., Braw, Y., Most, D., Katz, L.N., Sheer, A., Gersner, R.,
Gregoire, A.J., Kumar, R., Everitt, B., Henderson, A.F., Studd, J.W., 1996. Transdermal Zangen, A., 2009. Deep transcranial magnetic stimulation over the prefrontal
oestrogen for treatment of severe postnatal depression. Lancet 347 (9006) 930– cortex: evaluation of antidepressant and cognitive effects in depressive
933. patients. Brain Stimul. 2 (Oct (4)) 188–200.
Guaiana, G., Gupta, S., Chiodo, D., Davies, S.J., Haederle, K., Koesters, M., 2013. Levkovitz, Y., Sheer, A., Harel, E.V., Katz, L.N., Most, D., Zangen, A., Isserles, M., 2011.
Agomelatine versus other antidepressive agents for major depression. Cochrane Differential effects of deep TMS of the prefrontal cortex on apathy and depres-
Database Syst. Rev. 12, CD008851. sion. Brain Stimul. 4 (Oct (4)) 266–274.
Harel, E.V., Rabany, L., Deutsch, L., Bloch, Y., Zangen, A., Levkovitz, Y., 2014. H-coil Li, C.T., Chen, M.H., Juan, C.H., Huang, H.H., Chen, L.F., Hsieh, J.C., Tu, P.C., Bai, Y.M.,
repetitive transcranial magnetic stimulation for treatment resistant major Tsai, S.J., Lee, Y.C., Su, T.P., 2014. Efficacy of prefrontal theta-burst stimulation in
depressive disorder: an 18-week continuation safety and feasibility study. refractory depression: a randomized sham-controlled study. Brain 137 (Jul (Pt.
World J. Biol. Psychiatry 15 (May (4)) 298–306. 7)) 2088–2098.
Henigsberg, N., Mahableshwarkar, A.R., Jacobsen, P., Chen, Y., Thase, M.E., 2012. A Lopes Rocha, F., Fuzikawa, C., Riera, R., Ramos, M.G., Hara, C., 2013. Antidepressant
randomized, double-blind, placebo-controlled 8-week trial of the efficacy and combination for major depression in incomplete responders—a systematic
tolerability of multiple doses of lu AA21004 in adults with major depressive review. J. Affect. Disord. 144 (1–2) 1–6.
disorder. J. Clin. Psychiatry 73 (7) 953–959. Lozano, A.M., Giacobbe, P., Hamani, C., et al., 2012. A multicenter pilot study of
Holtzheimer 3rd, P.E., Nemeroff, C.B., 2006. Emerging treatments for depression. subcallosal cingulate area deep brain stimulation for treatment-resistant de-
Expert Opin. Pharmacother. 7 (17) 2323–2339. pression. J. Neurosurg. 116 (2) 315–322.
Holtzheimer, P.E., Nemeroff, C.B., 2008. Novel targets for antidepressant therapies. Magni, L.R., Purgato, M., Gastaldon, C., et al., 2013. Fluoxetine versus other types of
Curr. Psychiatry Rep. 10 (6) 465–473. pharmacotherapy for depression. Cochrane Database Syst. Rev. 7, CD004185.

Please cite this article in press as: Block, S.G., Nemeroff, C.B., Emerging antidepressants to treat major depressive disorder. Asian J.
Psychiatry (2014), http://dx.doi.org/10.1016/j.ajp.2014.09.001
G Model
AJP-644; No. of Pages 10

10 S.G. Block, C.B. Nemeroff / Asian Journal of Psychiatry xxx (2014) xxx–xxx

Mahableshwarkar, A.R., Jacobsen, P.L., Chen, Y., 2013. A randomized, double-blind Raison, C.L., Rutherford, R.E., Woolwine, B.J., et al., 2013. A randomized controlled
trial of 2.5 mg and 5 mg vortioxetine (lu AA21004) versus placebo for 8 weeks in trial of the tumor necrosis factor antagonist infliximab for treatment-resistant
adults with major depressive disorder. Curr. Med. Res. Opin. 29 (3) 217–226. depression: the role of baseline inflammatory biomarkers. JAMA Psychiatry 70
Martin, J.L., Martin-Sanchez, E., 2012. Systematic review and meta-analysis of vagus (1) 31–41.
nerve stimulation in the treatment of depression: variable results based on Rakofsky, J.J., Holtzheimer, P.E., Nemeroff, C.B., 2009. Emerging targets for antide-
study designs. Eur. Psychiatry 27 (3) 147–155. pressant therapies. Curr. Opin. Chem. Biol. 13 (3) 291–302.
Mathew, S.J., Manji, H.K., Charney, D.S., 2008. Novel drugs and therapeutic targets Ravindran, A.V., Kennedy, S.H., O’Donovan, M.C., Fallu, A., Camacho, F., Binder,
for severe mood disorders. Neuropsychopharmacology 33 (9) 2080–2092. C.E., 2008. Osmotic-release oral system methylphenidate augmentation of
Mathews, D.C., Henter, I.D., Zarate, C.A., 2012. Targeting the glutamatergic system to treat antidepressant monotherapy in major depressive disorder: results of a dou-
major depressive disorder: rationale and progress to date. Drugs 72 (10) 1313–1333. ble-blind, randomized, placebo-controlled trial. J. Clin. Psychiatry 69 (1)
Matthews, J.D., Siefert, C.J., Blais, M.A., et al., 2013. A double-blind, placebo- 87–94.
controlled study of the impact of galantamine on anterograde memory im- Rush, A.J., Marangell, L.B., Sackeim, H.A., et al., 2005. Vagus nerve stimulation for
pairment during electroconvulsive therapy. J. ECT 29 (3) 170–178. treatment-resistant depression: a randomized, controlled acute phase trial.
Mayberg, H.S., 2007. Defining the neural circuitry of depression: toward a new Biol. Psychiatry 58 (5) 347–354.
nosology with therapeutic implications. Biol. Psychiatry 61 (6) 729–730. Rush, A.J., Trivedi, M.H., Wisniewski, S.R., et al., 2006. Bupropion-SR, sertraline, or
McClintock, S.M., Tirmizi, O., Chansard, M., Husain, M.M., 2011. A systematic review venlafaxine-XR after failure of SSRIs for depression. N. Engl. J. Med. 354 (12)
of the neurocognitive effects of magnetic seizure therapy. Int. Rev. Psychiatry 23 1231–1242.
(5) 413–423. Sanacora, G., Berman, R.M., Cappiello, A., et al., 2004. Addition of the alpha2-
Montejo, A.L., Prieto, N., Terleira, A., et al., 2010. Better sexual acceptability of antagonist yohimbine to fluoxetine: effects on rate of antidepressant response.
agomelatine (25 and 50 mg) compared with paroxetine (20 mg) in healthy male Neuropsychopharmacology 29 (6) 1166–1171.
volunteers. an 8-week, placebo-controlled study using the PRSEXDQ-SALSEX Sanacora, G., Treccani, G., Popoli, M., 2012. Towards a glutamate hypothesis of
scale. J. Psychopharmacol. 24 (1) 111–120. depression: an emerging frontier of neuropsychopharmacology for mood dis-
Murrough, J.W., Iosifescu, D.V., Chang, L.C., et al., 2013. Antidepressant efficacy of orders. Neuropharmacology 62 (1) 63–77.
ketamine in treatment-resistant major depression: a two-site randomized Schatzberg, A.F., Nemeroff, C.B., 2009. The American Psychiatric Publishing Text-
controlled trial. Am. J. Psychiatry 170 (10) 1134–1142. book of Psychopharmacology. American Psychiatric Publishing, Inc., Arlington,
Musselman, D.L., Miller, A.H., Porter, M.R., et al., 2001. Higher than normal plasma VA, USA.
interleukin-6 concentrations in cancer patients with depression: preliminary Schatzberg, A.F., 2014. A word to the wise about ketamine. Am. J. Psychiatry 171 (3)
findings. Am. J. Psychiatry 158 (8) 1252–1257. 262–264.
Naughton, M., Clarke, G., O’Leary, O.F., Cryan, J.F., Dinan, T.G., 2014. A review of Schlaepfer, T.E., Bewernick, B.H., Kayser, S., Madler, B., Coenen, V.A., 2013. Rapid
ketamine in affective disorders: current evidence of clinical efficacy, limitations effects of deep brain stimulation for treatment-resistant major depression. Biol.
of use and pre-clinical evidence on proposed mechanisms of action. J. Affect. Psychiatry 73 (12) 1204–1212.
Disord. 156, 24–35. Schneider, L.S., Small, G.W., Hamilton, S.H., Bystritsky, A., Nemeroff, C.B., Meyers,
Nelson, J.C., Papakostas, G.I., 2009. Atypical antipsychotic augmentation in major B.S., 1997. Estrogen replacement and response to fluoxetine in a multicenter
depressive disorder: a meta-analysis of placebo-controlled randomized trials. geriatric depression trial. Fluoxetine Collaborative Study Group. Am. J. Geriatr.
Am. J. Psychiatry 166 (9) 980–991. Psychiatry 5 (2) 97–106.
Nelson, J.C., Mazure, C.M., Jatlow, P.I., Bowers Jr., M.B., Price, L.H., 2004. Combining Shiroma, P.R., Johns, B., Kuskowski, M., et al., 2014. Augmentation of response and
norepinephrine and serotonin reuptake inhibition mechanisms for treatment of remission to serial intravenous subanesthetic ketamine in treatment resistant
depression: a double-blind, randomized study. Biol. Psychiatry 55 (3) 296–300. depression. J. Affect. Disord. 155, 123–129.
Nemeroff, C.B., Widerlov, E., Bissette, G., et al., 1984. Elevated concentrations of CSF Shire Reports Top-line Results from Two Phase 3 Studies for Vyvanse1 (lisdex-
corticotropin-releasing factor-like immunoreactivity in depressed patients. amfetamine Dimesylate) Capsules (CII) as an Adjunctive Treatment for Adults
Science 226 (4680) 1342–1344. with Major Depressive Disorder. N.p., 06 Feb. 2014. Web. Apr. 2014. hhttp://
Nierenberg, A.A., Fava, M., Trivedi, M.H., et al., 2006. A comparison of lithium and www.shire.com/shireplc/en/investors/investorsnews/irshirenews?id=921i.
T(3) augmentation following two failed medication treatments for depression: Silverstone, T., 2001. Moclobemide vs. imipramine in bipolar depression: a multi-
a STAR*D report. Am. J. Psychiatry 163 (9) 1519–1530 (Quiz 1665). centre double-blind clinical trial. Acta Psychiatr. Scand. 104 (2) 104–109.
Ozomaro, U., Wahlestedt, C., Nemeroff, C.B., 2013. Personalized medicine in psy- Soares, C.N., Frey, B.N., 2010. Challenges and opportunities to manage depression
chiatry: problems and promises. BMC Med. 11, 132, http://dx.doi.org/10.1186/ during the menopausal transition and beyond. Psychiatr. Clin. North Am. 33 (2)
1741-7015-11-132. 295–308.
Papakostas, G.I., Nelson, J.C., Kasper, S., Moller, H.J., 2008. A meta-analysis of clinical Soares, C.N., 2008. Practical strategies for diagnosing and treating depression in
trials comparing reboxetine, a norepinephrine reuptake inhibitor, with selec- women: menopausal transition. J. Clin. Psychiatry 69 (10) e30.
tive serotonin reuptake inhibitors for the treatment of major depressive disor- Sogaard, J., Lane, R., Latimer, P., et al., 1999. A 12-week study comparing moclo-
der. Eur. Neuropsychopharmacol. 18 (2) 122–127. bemide and sertraline in the treatment of outpatients with atypical depression.
Papakostas, G.I., Cooper-Kazaz, R., Appelhof, B.C., et al., 2009. Simultaneous J. Psychopharmacol. 13 (4) 406–414.
initiation (coinitiation) of pharmacotherapy with triiodothyronine and a selec- Tanghe, A., Geerts, S., Van Dorpe, J., Brichard, B., Bruhwyler, J., Geczy, J., 1997.
tive serotonin reuptake inhibitor for major depressive disorder: a quantitative Double-blind randomized controlled study of the efficacy and tolerability
synthesis of double-blind studies. Int. Clin. Psychopharmacol. 24 (1) 19–25. of two reversible monoamine oxidase A inhibitors, pirlindole and moclobemide,
Park, S.H., Ishino, R., 2013. Liver injury associated with antidepressants. Curr. Drug in the treatment of depression. Acta Psychiatr. Scand. 96 (2) 134–141.
Saf. 8 (3) 207–223. Trivedi, M.H., Fava, M., Wisniewski, S.R., et al., 2006. Medication augmentation after
Plewnia, C., Pasqualetti, P., Große, S., Schlipf, S., Wasserka, B., Zwissler, B., Fallgatter, the failure of SSRIs for depression. N. Engl. J. Med. 354 (12) 1243–1252.
A., 2014. Treatment of major depression with bilateral theta burst stimulation: Tyring, S., Gottlieb, A., Papp, K., et al., 2006. Etanercept and clinical outcomes,
a randomized controlled pilot trial. J. Affect. Disord. 156 (Mar.) 219–223. fatigue, and depression in psoriasis: double-blind placebo-controlled random-
Pope Jr., H.G., Amiaz, R., Brennan, B.P., et al., 2010. Parallel-group placebo-con- ised phase III trial. Lancet 367 (9504) 29–35.
trolled trial of testosterone gel in men with major depressive disorder display- Versiani, M., Amrein, R., Stabl, M., 1997. Moclobemide and imipramine in chronic
ing an incomplete response to standard antidepressant treatment. J. Clin. depression (dysthymia): an international double-blind, placebo-controlled
Psychopharmacol. 30 (2) 126–134. trial. International Collaborative Study Group. Int. Clin. Psychopharmacol. 12
Portella, M.J., de Diego-Adelino, J., Ballesteros, J., et al., 2011. Can we really (4) 183–193.
accelerate and enhance the selective serotonin reuptake inhibitor antidepres- Wagstaff, A.J., Ormrod, D., Spencer, C.M., 2001. Tianeptine: a review of its use in
sant effect? A randomized clinical trial and a meta-analysis of pindolol in depressive disorders. CNS Drugs 15 (3) 231–259.
nonresistant depression. J. Clin. Psychiatry 72 (7) 962–969. Zarate Jr., C.A., Singh, J.B., Carlson, P.J., et al., 2006. A randomized trial of an N-
Preskorn, S.H., Baker, B., Kolluri, S., Menniti, F.S., Krams, M., Landen, J.W., 2008. An methyl-D-aspartate antagonist in treatment-resistant major depression. Arch.
innovative design to establish proof of concept of the antidepressant effects of Gen. Psychiatry 63 (8) 856–864.
the NR2B subunit selective N-methyl-D-aspartate antagonist, CP-101,606, in Zarrouf, F.A., Artz, S., Griffith, J., Sirbu, C., Kommor, M., 2009. Testosterone and
patients with treatment-refractory major depressive disorder. J. Clin. Psycho- depression systematic review and meta-analysis. J. Psychiatr. Pract. 15 (4)
pharmacol. 28 (6) 631–637. 289–305.

Please cite this article in press as: Block, S.G., Nemeroff, C.B., Emerging antidepressants to treat major depressive disorder. Asian J.
Psychiatry (2014), http://dx.doi.org/10.1016/j.ajp.2014.09.001

You might also like