Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Journal of Affective Disorders 156 (2014) 24–35

Contents lists available at ScienceDirect

Journal of Affective Disorders


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

Review

A review of ketamine in affective disorders: Current evidence


of clinical efficacy, limitations of use and pre-clinical evidence
on proposed mechanisms of action
Marie Naughton a,n, Gerard Clarke a,b, Olivia F. O′Leary b,c, John F. Cryan b,c,
Timothy G. Dinan a,b
a
Department of Psychiatry, University College Cork, Western Road, Cork City, Cork, Ireland
b
Alimentary Pharmabiotic Centre, University College Cork, Cork, Ireland
c
Department of Anatomy and Neuroscience, University College Cork, Cork, Ireland

art ic l e i nf o a b s t r a c t

Article history: Introduction: Recent research has seen low-dose ketamine emerge as a novel, rapid-acting antidepres-
Received 6 October 2013 sant. Ketamine, an N-methy-D-aspartate (NMDA) receptor antagonist, leads to effects on the glutama-
Received in revised form tergic system and abnormalities in this neurotransmittor system are present in depression. This article
19 November 2013
aims to (1) review the clinical literature on low-dose ketamine as a rapid-acting antidepressant in
Accepted 21 November 2013
Available online 10 December 2013
affective disorders, (2) provide a critical overview of the limitations of ketamine and research attempts to
overcome these (3) discuss the proposed mechanisms of action of ketamine and (4) point towards future
Keywords: research directions.
Ketamine Method: The electronic database Pubmed, Web of Science and sciencedirect were searched using the
N-Methyl-D-aspartate receptor antagonist
keywords: ketamine, N-methyl-D-aspartate receptor antagonist, rapid-acting antidepressant, depression,
Rapid-acting antidepressant
treatment-resistant depression, bipolar depression, suicidal ideation, electroconvulsive therapy, mechan-
Treatment-resistant depression
Bipolar depression ism of action.
Suicidal ideation Result: The literature demonstrates evidence supporting a rapid-acting antidepressant effect of low-dose
intravenous ketamine in major depressive disorder, in bipolar depression and in depression with suicidal
ideation. There are mixed results as to whether ketamine leads to a reduction in time to remission in
patients undergoing electroconvulsive therapy (ECT). Efforts to unravel ketamine′s therapeutic mechan-
ism of action have implicated the mammalian target of rapamycin (mTOR)-dependent synapse formation
in the rat prefrontal cortex, eukaryotic elongation factor 2 phosphorylation (p-eEF2) and glycogen
synthase kinase (GSK-3). Ketamine′s limiting factors are the transient nature of its antidepressant effect
and concerns regarding abuse, and research efforts to overcome these are reviewed.
Conclusion: Current and future research studies are using ketamine as a promising tool to evaluate the
glutamatergic neurotransmittor system to learn more about the pathophysiology of depression and
develop more specific rapid-acting antidepressant treatments.
& 2013 Elsevier B.V. All rights reserved.

Contents

1. Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2. Aims . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3. Method. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
4. Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
4.1. Clinical trials of ketamine in depression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
4.2. Clinical trials of ketamine in bipolar affective disorder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
4.3. Clinical trials of ketamine on suicidal ideation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
4.4. Preclinical studies and clinical trials to sustain the antidepressant effect of ketamine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

n
Corresponding author. Tel.: þ 353 214 901 224; fax: þ353 214 922 584.
E-mail addresses: marienaughton@beaumont.ie, m.naughton@ucc.ie (M. Naughton), g.clarke@ucc.ie (G. Clarke), o.oleary@ucc.ie (O.F. O′Leary), j.cryan@ucc.ie (J.F. Cryan),
t.dinan@ucc.ie (T.G. Dinan).

0165-0327/$ - see front matter & 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jad.2013.11.014
M. Naughton et al. / Journal of Affective Disorders 156 (2014) 24–35 25

4.5. Psychotomimetic side-effects and safety of ketamine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27


4.6. Ketamine combined with electroconvulsive therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
4.7. Postulated mechanisms of action underlying the rapid antidepressant effect of ketamine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
4.8. Future direction of research on ketamine as an antidepressant . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
4.8.1. Route of administration of ketamine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
4.8.2. Stereoisomers of ketamine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
4.8.3. Caution using ketamine in cancer patients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
4.8.4. Biomarkers of ketamine's antidepressant effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
5. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
Role of funding source . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
Conflict of interest. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

1. Background Ketamine is an approved anaesthetic agent for diagnostic and


surgical procedures in adults, obstetric patients and children
Major depressive disorder (MDD) is a debilitating mental (Lanning and Harmel, 1975). Its use had been associated with
illness that affects millions of people worldwide leading to severe dissociative reactions on emerging from anaesthesia; however, it
health and socioeconomic consequences (Kessler et al., 2003). remains a desirable anaesthetic because of its short half-life
Despite antidepressant treatment patients continue to experience (180 min) and the lack of respiratory depression (Clemens et al.
low remission rates, residual subsyndromal symptoms, relapses 1982). Over the past 40 years, ketamine has been administered as an
and persistent functional impairment. It is widely accepted that all anaesthetic to several million people and has a good safety profile
(Lahti et al., 2001; Corrsen et al., 1988; Reich and Silvay, 1989; White
current antidepressants require a lag period of several weeks
et al., 1982). Thirteen years ago the first clinical report of ketamine
before improvements in mood and wellbeing are felt. Another
being an effective, rapid-acting antidepressant emerged. The exciting
limitation of current monoamine-based antidepressants was high-
prospect of a rapid-acting antidepressant could lead to significant
lighted in the large well-known trial; sequenced treatment alter-
benefits for patients and would completely change prescribing and
native to relieve depression (STARnD Trial) (Rush et al., 2006a,
how mental health services are delivered.
2006b; Trivedi et al., 2006; Nierenberg et al., (2006); Rush, 2006c;
Fava et al., 2006). This study showed modest and diminished
returns from sequential trials of existing antidepressants in
patients who had not benefited from the selective serotonin 2. Aims
reuptake inhibitor, citalopram. Thus, there is a critical, unmet
The overall objective of this article is to review the evidence to
need to both identify and test novel drug targets for mood
date on the use of ketamine as a rapid-acting antidepressant. The
disorders in order to develop more effective treatments.
specific aims are to:
For years researchers have explored antidepressant options that
side-stepped the lag period for improvement in symptoms. Simulta-
 Review the clinical literature to date on low-dose intravenous
neously, there has been growing appreciation that investigation into
ketamine as a rapid-acting antidepressant in affective disorders,
the pathophysiology of mood disorders has been focused on mono-
 Provide a critical overview of the limitations of ketamine use in
aminergic systems and recently there has been more extensive
affective disorders and research attempts to overcome these,
research into other neurotransmitter signalling cascades such as
 Discuss the proposed mechanisms of action of ketamine,
the glutamatergic systems (Manji et al., 2003; Skolnick et al.,
derived mainly from preclinical studies and
2001). Leading on from this research, the glutamatergic system
 Point towards future research directions with ketamine as a
may offer a rational, rapid-acting target for drug development in
promising tool to develop novel, more effective, rapid-acting
mood disorders (Sanacora et al., 2008). Ketamine is a high affinity
antidepressants.
non-competitive antagonist at the N-methyl-D-aspartate (NMDA)
receptor, an ionotropic receptor in the glutamatergic system.
Glutamate is the major mediator of excitatory synaptic transmis-
sion in the mammalian brain (Orrego and Villanueva, 1993) and has
a prominent role in synaptic plasticity, learning and memory. A 3. Method
growing body of preclinical research implicates the glutamatergic
system in the pathophysiology of major depression and the in The search was performed using Pubmed, Web of science and
mechanism of action of antidepressant treatments (Skolnick et al., sciencedirect for papers published up to June 2013 using the
1996; Cryan and Dev, 2007; O′Connor and Cryan, 2010; O′Connor following search terms (MeSH/All fields): ketamine, N-methyl-
et al., 2010). In particular NMDA receptors appear to be specifically D-aspartate receptor antagonist, rapid-acting antidepressant,
involved (Trullas and Stolnick, 1990; Skolnick et al., 1996) although depression, treatment-resistant depression, bipolar depression,
metabotropic (mGlu) receptors are also known to be linked to suicidal ideation, electroconvulsive therapy, mechanism of action.
depression and the effects of antidepressants (O′Connor et al., 2013). All relevant clinical reports involving the assessment of ketamine′s
Conversely, chronic treatment with traditional antidepressants has antidepressant potential were considered. In addition, we consid-
been shown to affect NMDA receptor function (Mjellem et al., 1993) ered preclinical studies in animal models of depression or those
and receptor binding profiles (Paul et al., 1994). These findings carried out to provide insight into the mechanism of action of
suggest that NMDA receptor antagonism could form a viable ketamine. The bibliographies of relevant studies were also con-
treatment option for depression. However, clinical validation of this sidered. All articles without a focus on ketamine or NMDA receptor
hypothesis has only surfaced over the last decade. antagonists as antidepressants were excluded. Studies were
26 M. Naughton et al. / Journal of Affective Disorders 156 (2014) 24–35

excluded if the study was only published in abstract form or in a this in mind, Murrough et al. (2013) compared intravenous
language other than English. ketamine with midazalom, a benzodiazepine, considered a better
comparator than placebo. Patients with treatment-resistant major
depression experiencing a major depressive episode were ran-
4. Results domly assigned under double-blind conditions to receive a single
intravenous infusion of ketamine or midazolam in a 2:1 ratio
Ketamine and treatment-resistant depression as search terms (N ¼73). They demonstrated that the likelihood of response at 24 h
yielded a total of 70 articles of which 60 were relevant. Ketamine was greater with ketamine than with midazolam (odds ratio 2.18,
and rapid-acting antidepressant yielded 17 papers (15 relevant) 95% confidence interval 1.21–4.14) with response rates of 64% and
and N-methyl-D-aspartate receptor antagonist and rapid-acting 28%, respectively. Depression scores were not significantly differ-
antidepressant yielded 10 papers (7 relevant). Ketamine and ent between treatment groups at 7 days. Similar to previous
bipolar depression revealed 43 articles of which 29 were relevant. studies they demonstrated that side-effects associated with keta-
Ketamine and suicidal ideation revealed 10 articles of which 8 were mine included dissociative symptoms immediately following
relevant. Ketamine, Electroconvulsive therapy, and depression administration in 17% of patients. However, these side-effects
revealed 21 articles (14 relevant). Ketamine, mechanism of action resolved 2 h post-infusion. The authors commented that more
and rapid-acting antidepressant revealed 12 articles of which 10 information on response durability and safety is required before
were relevant. Overall the search yielded 143 relevant initial implementation in clinical practice.
papers. Relevance was determined by all authors based on
whether a significant proportion of the paper was dedicated to 4.2. Clinical trials of ketamine in bipolar affective disorder
ketamine and added to the body of research on ketamine as an
antidepressant. Following on from the initial trials of ketamine in MDD others
have looked at ketamine in bipolar depression (Diazgranados et al.,
4.1. Clinical trials of ketamine in depression 2010a; Zarate et al., 2012) and have shown it to be rapidly effective
but, similar to MDD studies, not lasting for beyond 1–2 weeks.
Berman et al. (2000) was the first to show that a subanaesthetic Patients were maintained on lithium or valproate during the study
(0.5 mg/kg) infusion of intravenous ketamine rapidly reduced timeframe. In contrast to traditional antidepressants, a subanaes-
depressive symptoms in depressed patients in a randomised, thetic dose of ketamine does not induce an affective switch in
placebo crossover design study. Four of the eight patients demon- treatment-resistant major depression (Niciu et al., 2013).
strated 50% or greater decrease in Hamilton depression rating
scores (HDRS) during the 3-day follow up period and ketamine- 4.3. Clinical trials of ketamine on suicidal ideation
induced mood improvement returned to baseline levels (i.e.
clinical impression and HDRS within 5 points of baseline) 1–2 Interest has also been generated regarding ketamine′s effect on
weeks after the infusion. Zarate et al. (2006), in a similar suicidal ideation. Although such patients were excluded from
randomised, placebo-controlled double-blind crossover study of initial RCTs, ketamine rapidly reduced suicidal ideation scores on
18 patients with treatment-resistant depression confirmed keta- the depression rating scales used (Berman et al., 2000; Zarate
mine′s rapid acting antidepressant effects. In this study, subjects et al., 2006). This has since been replicated in 2 open-label
had on average failed six prior antidepressant trials and were investigations of MDD patients and in one RCT of bipolar patients
medication free at least 2 weeks prior to the infusion. Notably, the (Zarate et al. 2012; DiazGranados et al., 2010b; Price et al., 2009).
response rates obtained with ketamine after 24 h (71%) were They demonstrated that ketamine was associated with robust and
similar to those described after 6–8 weeks of treatment with rapid antisuicidal effects. DiazGranados et al. (2010b) gave 33
traditional monoaminergic-based antidepressants (65%) (Entsuah subjects with treatment-resistant MDD a single open-label infu-
et al., 2001; Thase et al., 2005). sion of ketamine (0.5 mg/kg). Patients were rated at baseline, 40,
Since these 2 landmark studies, there have been numerous 80, 120 and 230 min post-infusion using the scale for suicide
other small studies evaluating ketamine′s rapid acting antidepres- ideation (SSI) (Beck et al., 1979). Scores significantly decreased
sant response and the results have been consistent with rapid within 40 min of ketamine infusion and remained improved for up
reductions in depressive symptomatology maintained for approxi- to four hours post infusion. In another study, early antisuicidal
mately 1–2 weeks following a single infusion. Response rates in effects (within one day) were found with ketamine and these
the open-label investigations and controlled trials have ranged effects remained significant for several weeks (Aan het Rot et al.,
from 25% to 85% at 24 h post infusion and from 14% to 70% at 72 h 2010). Repeated ketamine infusions can also reduce suicidality
post infusion (Aan het Rot et al., 2012). Although adverse effects ratings (MADRS-SI) but only as long as the duration of the 12-day
have generally been mild, some patients have experienced brief repeated infusion trial (Price et al., 2009). A study investigating IV
changes in blood pressure, heart rate or respiratory rate. However, ketamine in the emergency room showed beneficial results but
given the paucity of randomised controlled trials, lack of an active this study was not without its flaws, most notably lack of a control
placebo, limited data on long-term outcomes, ketamine adminis- group (Larkin and Beautrais, 2011). However these studies have
tration is not routinely recommended (Aan het Rot et al. 2012). led the way for further studies to explore this in more detail.
The largest clinical trial of ketamine to date (Murrough et al., Studies to date have provided us with clinical proof-of-concept
2013) was a two-site, parallel arm, randomized controlled trial of a that ketamine, an antagonist of the NMDA receptor in the gluta-
single infusion of ketamine compared to an active placebo control matergic system, has rapid antidepressant effects in affective
condition, the anaesthetic midazolam. Although previous clinical disorders and appears to reduce suicidal ideation. Questions remain
studies provide compelling evidence of the antidepressant effects in relation to how these significant findings can be moulded into a
of ketamine, their major weakness is the lack of a psychoactive coherent treatment strategy. The studies to date have shown
placebo-controlled, double-blind design. Ketamine produces psy- that ketamine′s effect peaks at 24 h post infusion and, in general,
choactive effects that are easily recognised by patients who report last 1–2 weeks. Clearly, maintaining the transient antidepressant
brief, transient feeling of being “high”. Hence, limitations in response of ketamine is required. Additionally the psychotomimetic
preserving study blinding may bias patient reporting by diminish- side-effects of ketamine and the concern regarding abuse potential
ing placebo effects, thereby potentially confounding results. With of ketamine make clinical use of ketamine itself unfeasible.
M. Naughton et al. / Journal of Affective Disorders 156 (2014) 24–35 27

Mechanistic insights are thus urgently required in order to produce dose ketamine infusion. Riluzole is a glutamate-modulating agent
a similarly acting agent without the associated safety issues. Thirdly, with neuroprotective and synaptic plasticity-enhancing proper-
one wonders could ECT antidepressant effects be expedited if ties, used in the treatment of amyotrophic lateral sclerosis (Doble,
ketamine is used as the induction agent? How the literature has 1996; Du et al., 2007; Mizuta et al., 2001). Riluzole was adminis-
attempted to address these challenges will be discussed below. tered orally for 28 days to sustained ketamine responders in one
study (Mathew et al., 2010) and to all participating MDD patients
4.4. Preclinical studies and clinical trials to sustain the 4–6 h after receiving an IV ketamine infusion in another study
antidepressant effect of ketamine (Ibrahim et al., 2012). Unfortunately, both studies failed to provide
any benefit over placebo in maintaining response to a single
Strategies to sustain the rapid antidepressant response to keta- ketamine dose. One case study found positive results with oral
mine have largely centred on repeated infusions, usually adminis- memantine, a low-to-moderate affinity, non-competitive NMDA
tered on alternate days over an extended period of time. Repeated antagonist (Kollmar et al., 2008), however, a placebo-controlled
infusions of ketamine were assessed preclinically by Parise et al. study of memantine found no antidepressant effects (Zarate et al.,
(2013) in adolescent male rats (postnatal day 35) who received two 2006b). Thus, to date, efforts to maintain ketamine′s antidepres-
ketamine (0, 5, 10, or 20 mg/kg) injections, 4 h apart, after exposure sant effect have been disappointing. What is promising is that
to day 1 of the forced swim test (FST). The next day, rats were re- another low-affinity NMDA receptor antagonist, AZD6765, demon-
exposed to the FST to assess ketamine-induced antidepressant-like strated antidepressant effects in humans at doses that do not
responses. Separate groups were exposed to chronic unpredictable produce psychosis (Zarate et al., 2013).
stress to confirm findings from the FST. After these initial experi-
ments, adolescent naive rats were exposed to either 1 or 15 4.5. Psychotomimetic side-effects and safety of ketamine
consecutive days (PD35-49) of ketamine (20 mg/kg) twice daily.
Ketamine′s influence on behavioural reactivity to rewarding (i.e., A vital question that should always be asked when a potentially
sucrose preference) and aversive (i.e., elevated plus-maze, FST) new medication (or an old medication used for a new purpose) is
circumstances was then assessed 2 months after treatment. To being considered for use is safety. There is significant evidence to
control for age-dependent effects, adult rats (PD75-89) were exposed demonstrate ketamine′s safety profile when given on a single
to identical experimental conditions. What was found was that occasion as an anaesthetic agent but a lot less is known about its
ketamine (20 mg/kg) reversed the chronic unpredictable stress- safety when given repeatedly at subanaesthetic doses. No such
induced depression-like behaviours in the FST. Repeated ketamine safety studies have been done with depressed patients. However
exposure resulted in anxiolytic- and antidepressant-like responses Zarate et al. (2006) demonstrated that adverse effects occurred
2 months after drug exposure. None of the ketamine doses used was more commonly in participants taking ketamine than those taking
capable of inducing drug-seeking behaviours as measured by place placebo and these were perceptual disturbances, confusion, eleva-
preference conditioning (Parise et al., 2013). The authors concluded tions in blood pressure, euphoria, dizziness and increased libido.
by indicating that repeated ketamine exposure induces enduring The majority of adverse effects ceased within 80 min after the
resilient-like responses regardless of age of exposure. These findings, infusion. In no case did euphoria or derealisation/depersonalisa-
although preliminary, point to ketamine, and its repeated exposure, tion persist beyond 110 min. Murrough et al. (2013), in their two-
as a potentially useful antidepressant during adolescence. Further site randomized controlled clinical trial of ketamine in patients
similar studies are awaited. with treatment-resistant depression reported that eight of the
Results with repeated infusions clinically have been less posi- 47 patients who received ketamine (17%) had significant dissocia-
tive. Aan het Rot et al. (2010) looked at response to repeated IV tive symptoms. Blood pressure in the ketamine group rose from
ketamine infusions in 10 patients in an in-patient setting in 122/72 mm Hg (pre-treatment) to 141/81 (40 min after infusion),
treatment resistant depression (TRD). The primary efficacy mea- and two subjects required their infusions to be stopped for
sure was change from baseline in the Montgromery–Asberg hemodynamic reasons. Berman et al. (2000) found transient
depression rating scale (MADRS) score. If patients showed a cognitive deficits and euphoria induced by ketamine infusion as
greater than 50% reduction in MDRAS scores on day 2 (9 of 10 evidenced by increases in the brief psychiatric rating scale (Overall
patients) they received five additional infusions on an outpatient and Gorham, 1962). However this increase returned to baseline
basis (days 3, 5,8,10 and 12). Of these nine patients, eight relapsed within 2 h following the infusion. Despite these side-effects one
within an estimated mean 30 days after the first infusion or 19 must consider the adverse side effect profile of many currently
(SD ¼13) after the sixth infusion. However the small sample size used antidepressants and if these findings with ketamine could be
and lack of a control group limit the extent to which these findings directly compared with monoaminergic antidepressants, it may
can be generalised. Murrough et al. (2013b) in a similar study represent some advantage for ketamine over existing therapies.
design with 24 participants with treatment-resistant depression A significant limitation of ketamine as a therapeutic agent is that
had a very similar result with antidepressant effects lasting on it is one of several “club drugs” that is abused. ketamine or “Special
average 18 days after the last infusion. Aan het Rot et al. (2010) K” is widely available to induce a dissociative state of relaxed
concluded with the suggestion that the most practical use of wellbeing. Zarate et al. (2010), however, pointed out that misuse of
repeated ketamine infusions might be on a short-term basis to therapeutically relevant agents is not a new phenomenon in psy-
elicit rapid relief until a more sustainable treatment and relapse chiatry (i.e. anticholinergic drugs, stimulants and benzodiazepines)
prevention strategy could be demonstrated. A more recent study and should not preclude their study as putative treatments. This view
(Segmiller et al., 2013) administered 6 infusions of S-ketamine point would be reiterated by Nutt et al. (2013) who indicate that
over a 4 week period to 6 patients with treatment resistant important and unfortunate outcomes of controls placed on certain
depression and demonstrated an improvement in the 21-item psychoactive drugs makes research into mechanism of actions and
Hamilton depression rating scale scores over the study timeframe potential therapeutic uses difficult. However despite the fact that the
but depressive symptoms were only evaluated pre and 120 min studies above demonstrate that the psychotomimetic effects of
post the 6 ketamine infusions and how the improvement was ketamine appear transient, concerns regarding abuse could lead to
maintained in the longer term was not commented on. the therapeutic benefits being overlooked.
Other clinical studies have attempted to maintain ketamine′s Mathew et al. (2010) in addition to administering riluzole in the
antidepressant effects by administering riluzole following a low- study outlined above also looked to see if pre-treatment with
28 M. Naughton et al. / Journal of Affective Disorders 156 (2014) 24–35

lamotrigine might attenuate ketamine′s psychotomimetic effects ketamine predicted lower response rate at 24 h). This suggests a
and enhance antidepressant activity. Unfortunately, lamotrigine potential baseline neurocognitive predictor of ketamine response
failed to reduce the transient psychotomimetic or dissociative side and an inverse relationship between the cognitive effects of
effects associated with ketamine use and did not enhance its ketamine and antidepressant efficacy.
antidepressant effects. In humans ketamine abuse has been associated with cystitis
Further study is required to determine whether the therapeutic (Chen et al., 2009) and bilary dilatation (Wong et al., 2009).
benefits of NMDA receptor inhibition can be realised using alter- However, Krystal et al. (2013) commented that such findings in
native agents that target a specific subunit of the NMDA receptors chronic ketamine abusers may over estimate the clinical risks of
without causing psychotomimetic side-effects. Preskorn et al. long-term treatment. Ketamine abusers often take multiple sub-
(2008) set out to investigate the antidepressant efficacy of the stances, administer higher ketamine doses than needed to treat
NR2B subunit selective antagonist, CP-101,606, which inhibits depression, and administer these doses more frequently than
NMDA receptors by an allosteric mechanism (Mott et al., 1998) would be needed for treatment (Morgan and Curran, 2012).
that is distinct from that of ketamine. NMDA receptors are Nonetheless, the development of long-term ketamine treatment
tetrameric proteins composed of 2 NR1 subunits and 2 NR2 for depression would need to be accompanied by careful studies of
subunits (Dingledine et al., 1999). There are 4 different NR2 its safety.
subunits (NR2A-D) that are differently localised throughout the
CNS. NMDA receptors containing the NR2B subunit are localised 4.6. Ketamine combined with electroconvulsive therapy
primarily in the forebrain including the hippocampus, a region
implicated in the pathophysiology of major depression (Campbell Electroconvulsive therapy (ECT) is recognised as a highly
and MacQueen, 2004). In this study 30 SSRI-resistant patients effective treatment of unipolar and bipolar depression (Weiner,
were randomised to CP-101,606 or a matched placebo infusion (15 2001) and it is generally considered to have a more rapid onset of
in each group). Of the CP-101,606-treated subjects, 60% met action than standard antidepressant agents. The data on the onset
criteria for response and 33% met criteria for remission on day 5. of ECTs antidepressant action suggest that a range of 5–7 treat-
This is comparable to previous studies (Berman et al., 2000; Zarate ments (approx 3 weeks) are required to obtain a significant
et al., 2006). About 6 of the 15 patients who received CP 101,606 reduction in symptom severity (Nobler et al., 1997) which is still
had a dissociative response compared to 2 on placebo and all considerably longer than the time of antidepressant effect
resolved within 6 hours of discontinuation of the infusion. The observed with ketamine. One of the most promising aspects of
authors indicated that 5 of the 9 CP-101,606 treated patients who ketamine is that even patients resistant to electroconvulsive
met HDRS criteria for antidepressant response did not have a therapy (ECT) may benefit from it (Zarate et al., 2006; Ibrahim
dissociative response from the infusion indicating that it is capable et al., 2011). The reverse, whether patients who do not respond to
of producing an antidepressant response without also producing a subanaesthetic IV ketamine doses might still benefit from ECT, is
dissociative reaction. While this is a small study, it points the way still not known. Given the fact that ketamine is a licensed
forward for further research to develop a new class of antidepres- anaesthetic agent, it is reasonable to question whether co admin-
sant with a reduced potential for psychotomimetic side-effects. istration of ketamine could expedite the antidepressant response
This might appeal to clinicians and patients alike, particularly of ECT. Interestingly ketamine has been used in ECT anaesthesia
those with a history of addictions. for decades (Green, 1973; Brewer et al., 1972) with preliminary
Linked with the concept above is the interesting effect of evidence suggesting that ketamine anaesthesia in ECT may
ketamine on subjects with alcohol dependence and those with improve seizure duration relative to other anaesthetic agents that
family histories of alcohol dependence. Compared to healthy are commonly used (Krystal et al., 2003b; McDaniel et al., 2006).
controls, subjects with alcohol dependence had fewer perceptual However, there is no evidence that the direct antidepressant effect
alterations and decreased dysphoric mood during ketamine infu- of the ketamine was considered as a potential benefit to the use of
sion (Krystal et al., 2003). This attenuation of ketamine-induced this drug in these early studies.
perceptual alterations was also observed in healthy individuals There have been some case reports in the ECT setting demon-
with a positive family history of alcohol dependence (Petrakis strating a more rapid antidepressant effect with ketamine use
et al., 2004). Indeed, Phelps et al. (2009) found that patients with (Goforth and Holsinger, 2007). Two uncontrolled studies have since
MDD who had a family history of alcohol dependence had a better shown that depression scores may decrease faster during ECT in
short-term outcome (greater and faster improvement in depres- patients given ketamine than in patients given propofol or thiopen-
sive symptoms) after ketamine infusion than subjects with no tone (Okamoto et al., 2010; Kranaster et al., 2011). However, an RCT of
family history of alcohol dependence. Genetic variation in the thiopental alone or thiopental plus ketamine (0.5 mg/kg) for anaes-
NMDA receptor 2A gene has been associated with alcohol depen- thesia before each ECT did not appear to enhance or expedite the
dence (Schumann et al., 2008), pointing towards a potential antidepressant effect of ECT (Abdallah et al., 2012). Thus, further
genetic mechanism underlying this variation in the antidepressant studies investigating ketamine′s augmentation to ECT need to be
response to ketamine in individuals with alcohol dependence. undertaken which should consider the effects of dose, timing and
Despite extensive research in healthy volunteers (Krystal et al., concomitant medications in the study design.
1994, 1999, 2005; Morgan et al., 2004; Parwani et al., 2005; Perry
et al., 2007) there is a paucity of studies examining the neuro- 4.7. Postulated mechanisms of action underlying the rapid
cognitive effects of ketamine in depressed patients. Murrough antidepressant effect of ketamine
et al. (2013c) assessed neuro-cognitive functioning in 25 patients
with TRD using a comprehensive battery following a 40-min The neurobiological mechanisms underlying the antidepressant
intravenous infusion of ketamine (0.5 mg/kg). They demonstrated actions of ketamine are more complex than simple blockade of
that patients who responded to ketamine 24 h following treat- NMDA receptors given its short half-life relative to the antidepressant
ments had poorer baseline neurocognitive performance relative to effect seen in the afore mentioned clinical studies. The very low dose
non-responders and, in particular, slower processing speed. Keta- of ketamine used for these studies first produces mild psychotomi-
mine was associated with selective impairments in memory recall, metic and dissociative effects 30–40 min after administration, effects
and the degree of cognitive change carried negative prognostic that are transient and completely dissipate by 80 min (Zarate et al.,
significance (e.g., negative cognitive effects immediately after 2006). This is presumably because of the rapid metabolism of
M. Naughton et al. / Journal of Affective Disorders 156 (2014) 24–35 29

ketamine (half-life is 180 min in humans; Clemens et al., (1982)). neurons, as expected, and decreased 5-HT and hypocretin-induced
After this initial psychotomimetic phase, the antidepressant effects EPSCs in the same neurons (Li et al., 2011). A single dose of ketamine
are observed at 110 min and are sustained for approximately 7 days completely reversed the deficit in spine density, and this was
after a single dose of ketamine (Zarate et al., 2006). These findings accompanied by a reversal of the deficit in 5-HT and hypocretin
indicate that ketamine initiates a cascade of events that results in a induced EPSC. Together these studies demonstrate that ketamine
rapid response that is sustained even after the drug has been reverses the spine loss caused by chronic stress exposure and
metabolised (see Fig. 1). Preclinical studies have been investigating normalizes PFC connectivity. To examine the possibility that reversal
the intriguing question of ketamine's mechanism of action and in of the synapse loss caused by CUS influences behavior, Li et al. (2011)
this context recent reviews on preclinical ketamine studies are also examined sucrose preference, which provides a measure of
relevant (Duman et al., 2012; Zunszain et al., 2013). anhedonia in rodents. CUS exposure for 3 weeks significantly
The potential role of dendrites and spines in stress-related decreased the preference for a sweetened solution, and this effect
illnesses such as depression is supported by basic studies demon- was completely reversed by a single dose of ketamine. In addition, CUS
strating that exposure to stress causes atrophy of neurons in limbic increased the latency to feed in the NSFT, and this effect was also
brain regions implicated in depression, including the prefrontal reversed by a single dose of ketamine. These effects of ketamine were
cortex (PFC) and hippocampus (McEwen, 2008; Shansky and also sustained for approximately 7 days, similar to the time course for
Morrison, 2009). This includes a decrease in the density of spines, the antidepressant actions of ketamine in treatment resistant
as well as a decrease in the number and length of dendrite branches. depressed patients (Zarate et al., 2006).
These effects could contribute to the reduction in volume of PFC and These synaptogenic and behavioural effects of ketamine outlined
hippocampus determined by imaging the brains of depressed above were shown to be dependent on stimulation of the mamma-
patients (Drevets and Furey, 2010; Macqueen et al., 2008). In contrast lian target of rapamycin (mTOR) (Li et al., 2010). mTOR is a large
to the effects of stress, Li et al. (2010) reported that administration of serine/threonine kinase that regulates the initiation of protein
a low dose of ketamine results in the rapid induction of spine translation. It is ubiquitously expressed and can control new protein
number in layer V pyramidal neurons of the PFC. Spine analysis was synthesis when required for synaptogenesis (Duman et al., 2012).
conducted 24 h after ketamine administration, and it is possible that Cryan and O'Leary (2010) commenting on the significance of these
the induction of spine formation occurs sooner, given the rapid findings indicating that m-TOR mediated pathways may be an
induction of synaptic proteins (i.e., as early as 2 h after ketamine important and novel strategy for the rational design of fast-acting
administration). At the 24 h time point, there is an increase in the antidepressants.
number of mushroom or mature spines, indicating that ketamine Ketamine can enhance AMPA receptor throughput (Zarate and
increases spine stability and function. This possibility was directly Manji, 2008). Maeng et al. (2008) showed, in rodent models of
tested by analysis of neurotransmitter-induced excitatory postsynap- depression, that the α-amino-3-hydroxy-5-methyl-4-isoxazole-
tic currents (EPSCs) in the same PFC layer V pyramidal neurons that propionic acid (AMPA) receptor is required for its antidepressant
were analyzed for spine density. The results demonstrate that effect and postulated that ketamine may result in a rapid anti-
ketamine administration significantly increases the frequency and depressant effect by enhancing AMPA relative to NMDA through-
amplitude of both 5-HT- and hypocretin-induced EPSCs. The increase out critical neuronal circuits. They demonstrated at a cellular level
in EPSP amplitude is consistent with the increase in the density of that ketamine's antidepressant effects were found to be selectively
mushroom spines by ketamine. The increase in 5-HT- and abolished using an AMPA antagonist (NBQX) prior to ketamine
hypocretin-induced EPSCs indicates that there is an increase in infusion. They also demonstrated that the antidepressant effects of
corticocortical and thalamocortical connections, respectively. MK-801 (a non-selective NMDA antagonist) and Ro 25-6981 (an
Consistent with the rapid induction of synaptogenesis, the NR2B selective antagonist) also require AMPA receptor through-
authors above (Li et al., 2010) also found that ketamine produced put, again by pre-treating with NBQX. Interestingly, the antide-
rapid antidepressant effects in several different behavioral models. pressant effect of MK-801 or RO25-6981 were not as long-lasting
This included a significant decrease in immobility in the forced as those of ketamine. An increase in AMPA/NMDA receptor density
swim test (FST), decreased escape failure and latency to escape in ratio has been observed in the hippocampus of rats after ketamine
the learned helplessness (LH) paradigm, and decreased latency to treatment (Tizabi et al., 2012). Thus there is an emerging com-
feed in the novelty suppressed feeding test (NSFT) (Li et al., 2010). plementary pharmacological approach consisting of targeting
Although the FST is responsive to acute administration of typical AMPA receptors (AMPAR) to modulate glutamatergic transmission.
antidepressant agents, LH is only responsive to subchronic (7 days) It may appear counterintuitive that AMPA potentiators have
treatment, and the NSFT, although a model of anxiety, is respon- antidepressant effects, knowing that antidepressants have been
sive to chronic (21 days) administration of a typical antidepres- shown to dampen glutamate release (Martinez-Turrillas et al.,
sant. Together, these findings provide evidence of the fast 2005; Barbon et al., 2006; Tan et al., 2006; Svenningsson et al.,
antidepressant behavioral actions of ketamine in these rodent 2002). However AMPA potentiators work by increasing the
models. Interestingly, Harkin and colleagues demonstrated that response of AMPAR, which is independent on the actual level of
ketamine's antidepressant-like effect in the FST is via a serotonin- glutamate release. Several classes of AMPA potentiators, including
dependant mechanism (Gigliucci et al., 2013). nootropic agents and ampakines, have shown antidepressant
In addition to studies in normal animals, Li et al. (2011) also efficacy in preclinical studies (O'Neill et al., 2004; Li et al., 2001)
examined the influence of ketamine in animals exposed to chronic and a number of compounds are currently under clinical devel-
unpredictable stress (CUS). In the CUS model, repeated exposure opment. Information on the long-lasting clinical effects of a single
to stress over the course of several weeks results in the development dose of AMPA is not yet available however.
of anhedonia, a core symptom of depression, and this effect is reve- Following on from studies above, increasing the ratio of AMPA-
rsed by chronic administration of a typical antidepressant (Wilner, to-NMDA receptor medicated neurotransmission is being assessed
2005; Banasr et al., 2007). In addition, chronic stress exposure as a mechanism of antidepressant effect in preclinical behavioral
decreases the density of spines in the PFC (Liu and Aghajanian, tests of depression with positive results (Tizabi et al., 2012;
2008; Shansky and Morrison, 2009), providing a morphological end- Andreasen et al., 2013). Akinfiresoye and Tizabi (2013) tested
point that can be measured and that is relevant to the atrophy of PFC whether AMPA alone has an antidepressant effect and if the
in depression (Drevets and Furey, 2010). They found that exposure to combination of AMPA and ketamine provides added benefit
CUS for 3 weeks decreased the density of spines in layer V pyramidal in Wistar-Kyoto rats. They demonstrated that chronic AMPA
30 M. Naughton et al. / Journal of Affective Disorders 156 (2014) 24–35

Gln

Glu Pre synaptic


glutamatergic
neuron

Ketamine

AMPA-R
AMPA-R
NMDA-R
NMDA-R
Post synaptic
glutamatergic
eEF2 kinase neuron

eEF2 eEF2 P

BDNF translation

TrK B

GSK-3 Akt ERK

mTor activation

Preclinical behavioural effects Synaptic effects Clinical effects on mood

Increased
spine density

Pre-synaptic membrane
Synapsin

PSD95 GluR1

Post-synaptic membrane
Increased
synaptic proteins

Increased
synaptic plasticity

Fig. 1. Diagrammatic representation of the postulated mechanisms of action of ketamine on the glutamatergic neurotransmitter system. Glutamate (Glu) is synthesized in
presynaptic neurons from glutamine by glutaminase. It is packaged into synaptic vesicles and released into the synaptic cleft. Glutamate binds to either ionotropic N-methyl-
D-aspartate (NMDA) or alpha-amino-3-hydroxy-5-methyl-4-isoxazole-propionic (AMPA) receptors localised in the postsynaptic density (Other receptors such as the metabotrophic
receptors are also involved in glutamate's action but not discussed here). Under resting conditions NMDA-receptor activation leads to activation of eukaryotic elongation factor-2
(eEF2) kinase triggering eEF2 phosphorylation and silencing brain-derived neurotrophic factor (BDNF) translation. Ketamine is a high affinity antagonist at the NMDA receptor and
blockade of this receptor leads to less activation of eEF2 kinase which gradually results in loss of eEF2 phosphorylation and de-suppression of BDNF translation (Autry et al., 2011).
BDNF translation ultimately triggers TrkB (tyrosine-related kinase B) signalling, leading to transphosphorylation and downstream activation of extracellular signalling-related
kinases (ERK) and protein kinase B (PKB/Akt) and suppression of glycogen synthase kinase-3 (GSK-3). These signalling events activate mTOR (mammalian target of rapamycin),
leading to increased synaptic proteins and spine densities (in preclinical studies of rat prefrontal cortices) resulting in synaptogenesis (Li et al., 2010). The ultimate result of this
pathway and synaptogenesis is positive behavioural effects in animal models of depression and thus is considered to be the possible mechanism whereby ketamine induces its
antidepressant effect clinically. Increased AMPA receptor throughput has a central role in the antidepressant effect of ketamine possibly by increasing BDNF release, TrkB receptor
activation, stimulation of mTOR signalling and local protein synthesis (Akinfiresoye and Tizabi, (2013); Jourdi et al., 2009).

treatment resulted in a dose-dependent antidepressant effect in ketamine, at doses that were ineffective on their own, resulted in a
both the forced swim test and sucrose preference test. Moreover, significant antidepressant effect. The behavioral effects were
chronic administration (10–11 days) of combinations of AMPA and associated with increases in hippocampal brain-derived
M. Naughton et al. / Journal of Affective Disorders 156 (2014) 24–35 31

neurotrophic factor, synapsin, and mammalian target of rapamy- whether ketamine affected the inhibitory serine phosphorylation of
cin. These findings provided evidence for an antidepressant effect the abundant GSK-3b isoform in hippocampal synaptosomes. They
of AMPA and suggested the usefulness of AMPA-ketamine combi- did not detect regulation of phosphorylated GSK-3b at 2 h after
nation in treatment of depression. Furthermore, these effects acute ketamine administration. Several important technical differ-
appear to be associated with increases in markers of hippocampal ences exist between (Muller et al. 2013) and the study by Beurel
neurogenesis and synaptogenesis, suggesting a mechanism of et al. to include the locations and timing of GSK-3 evaluation and
their action. the animal model. Ma et al. (2013) examined the effects of ketamine
Expanding on the clinical study (Preskorn et al., 2008) above, and GSK-3 inhibitor SB216763 in the unpredictable, chronic mild
investigating CP-101,606, an NR2B receptor antagonist in the stress (CMS) mouse model of mice and found that a single
treatment of depression, other selective NR2B antagonists have administration of ketamine, but not GSK-3 inhibitor SB216763,
been assessed preclinically and what is becoming apparent is that produces a long-lasting antidepressant action. Thus the jury is still
they have similar behavioural and molecular effects to ketamine. out regarding the role of GSK-3 inhibition in the antidepressant
The rationale for this avenue of research is to determine if effect of ketamine and more work needs to be done in this area.
ketamine's antidepressant effect can be observed without the Using microdialysis, ketamine has been reported to increase
psychomimetic side effects. Li et al. (2010) demonstrated that Ro the extracellular levels of glutamate in the prefrontal cortex of
25-6981 stimulated mTOR signalling and had antidepressant rodents (Lorrain et al., 2003; Moghaddam et al., 1997), leading to
effects in the FST and the novelty suppressed feeding test (NSFT) the hypothesis that hyperglutamatergic mechanisms are involved
and that these effects were blocked by infusion with rapamycin. In in the antidepressant effect of ketamine. This alongside the idea
the CUS paradigm, the NR2B selective antagonist completely that structurally non-related antidepressants may have converging
blocked the deficits in sucrose consumption and novelty sup- mechanisms of actions with relevance for the time of their onset of
pressed feeding resulting from CUS exposure (Li et al., 2011). These action prompted Muller et al. (2013) to investigate whether acute
findings suggest that the actions of ketamine are mediated by administration of ketamine targets the presynaptic molecular
blockage of NR2B receptors, and provide further evidence of a machinery. They demonstrated that ketamine rapidly induces
functional connection between induction of synaptogenesis and changes in the hippocampal presynaptic machinery similar to
antidepressant behavioural responses (Duman et al., 2012). It has those that are obtained only with chronic treatments with tradi-
yet to be shown at a preclinical level if NR2B antagonists can have tional antidepressants. They demonstrated a large reduction in the
antidepressant-like effects without ketamine's psychomimetic accumulation of SNARE (soluble N-ethylmaleimide-sensitive factor
effects. attachment protein receptor) complexes was observed in hippo-
Recent studies by Monteggia et al. (2013) show that ketamine campal synaptic membranes after 1, 2 and 4 h of ketamine
effects are dependent on brain-derived neurotrophic factor (BDNF) administration. Neurotransmitter release requires the assembly
and subsequent activation of the high-affinity BDNF receptor TrkB. of the core SNARE complex, consisting of the plasma membrane
Their findings point towards eukaryotic elongation factor 2 kinase associated proteins syntaxin 1A and SNAP25 (synaptosomal-asso-
(eEF2K), which phosphorylates eEF2 and regulates the elongation ciated protein 25 kDa) and the synaptic vesicle-associated protein
step of protein translation, as a major molecular substrate mediat- VAMP2 (vesicle-associated membrane protein 2) (Südhof, 2004).
ing the rapid antidepressant effect of ketamine (Autry et al., 2011) Post-mortem studies indicate abnormalities in the expression of
Thus eEF2K is emerging as a potential antidepressant target (see individual SNARE proteins and regulatory proteins in the hippo-
Monteggia et al., 2013 for review). campus and frontal cortex of subjects with depression and bipolar
Administration of ketamine to mice has been shown to inhibit disorder (Fatemi et al., 2001; Scarr et al., 2006). In parallel, Muller
brain glycogen synthase kinase-3 (GSK-3), a kinase that, interest- et al. found a selective reduction in the expression of the synaptic
ingly, is also a target of mood-stabilising agents (Klein and Melton vesicle protein synaptotagmin I and an increase in the levels of
1996). Beurel et al. (2011) demonstrated that ketamine increases synapsin I in hippocampal synaptosomes suggesting a mechanism
the phosphorylation of GSK-3, and that mice with a knock out by which ketamine reduces SNARE complex formation, in part, by
mutation that blocks the phosphorylation of GSK-3 do not respond regulating the number of synaptic vesicles in the nerve terminals.
to ketamine in a behavioral model of depression. A more recent Moreover, ketamine reduced Thr (286)-phosphorylated αCaMKII
study (Liu et al., 2013) showed that the synaptogenic and (calcium/calmodulin dependent protein kinase II) and its interac-
antidepressant-like effects of a single otherwise subthreshold dose tion with syntaxin 1A, which identifies CaMKII as a potential target
of ketamine were potentiated when given together with a single for second messenger-mediated actions of ketamine. This suggests
dose of lithium chloride (a non-selective GSK-3 inhibitor) or a that reduction of neurotransmitter release in the hippocampus has
preferential GSK-3β inhibitor. They also demonstrated that these possible relevance for the rapid antidepressant effect of ketamine
effects involved rapid activation of the mTOR signaling pathway, and may represent important molecular targets for antidepressant
increased synaptic spine density/diameter and increased EPSCs in treatment (Muller et al., 2013).
the medial prefrontal cortex (mPFC) layer 5 pyramidal neurons According to Murch (2013) an overlooked connection in the
and antidepressant responses that persist for up to 1 week in the mechanism of action of ketamine in depression is the role of
FST model of depression. What is exciting is that these results magnesium, which acts as physiological NMDA-receptor antago-
demonstrate that low, subthreshold doses of ketamine combined nist. Murch effectively argues that there is overlap between the
with lithium or a selective GSK-3 inhibitor are equivalent to higher actions of ketamine with that of high doses of magnesium in
dose of ketamine indicating the pivotal role of the GSK-3 pathway animal models, ultimately leading to synaptic sprouting. Magne-
in modulating the synaptogenic and antidepressant responses to sium and ketamine lead to synaptic strengthening, as measured by
ketamine. The possible mitigation by GSK-3 inhibitors of the an increase in slow wave sleep in humans. Elaboration of his
eventual fading of ketamine's antidepressant effect remains to be theory is outlined in a recent article (Murch, 2013).
explored. What this research also questions is whether ketamine Overall, what is emerging from the exciting research discussed
combined with lithium in a clinical setting in bipolar depression above is that ketamine seems to compensate for reduced gluta-
has a more significant effect on depressive symptoms than when matergic signaling by disinhibiting the release of glutamate,
combined with valproate or other mood stabilisor. enhancing the stimulation of AMPA receptors and promoting
However the effects of GSK-3 in ketamine's mechanism of action downstream signaling via the Akt/Mtor pathway. Ketamine also
is far from clear and Muller et al. (2013) recently investigated blocks extrasynaptic NMDA glutamate receptors reducing the
32 M. Naughton et al. / Journal of Affective Disorders 156 (2014) 24–35

phosphorylation of EEF2, reducing the inhibition of BDNF, and ketamine and S-ketamine infusion therapy, respectively, in two
alleviating the suppression of dendritic spine growth. The transla- patients with treatment-resistant major depression Both patients
tion of the clinical observations into explanatory neuroscience has experienced psychotomimetic side effects during ketamine infu-
had a transformative impact on antidepressant research that will sion which were absent during treatment with S-ketamine. They
hopefully help in clarifying the future research directions of concluded that S-ketamine might exert similar antidepressant
ketamine and similar compounds. effects as ketamine in drug-resistant depression but may be better
tolerated by the patients. Further studies of S-ketamine are
required.
4.8. Future direction of research on ketamine as an antidepressant

4.8.1. Route of administration of ketamine


There has been limited study of ketamine in depression given 4.8.3. Caution using ketamine in cancer patients
orally which would be a preferable route of administration for Concerns have been raised with regard to up-regulation of
patients. It is known, however, that oral ketamine is used off-label mammalian target of rapamycin (mTOR) by ketamine (Yang et al.,
in palliative care and treatment of therapy-resistant pain. A small 2011) in patients with cancer. Studies have shown that up-
case report study with 2 patients given a single, oral ketamine regulated mTOR may cause the acceleration of tumor growth
(0.5 mg/kg) in a hospice setting for depression demonstrated rapid (Shor et al., 2009). Therefore, further, specific studies addressing
and modestly sustained symptom relief of depressive and anxiety patient prognosis are needed to investigate whether ketamine is
(Irwin and Iglewicz 2010). No adverse effects were noted. This was suitable for the treatment of refractory depression in patients with
followed up by a larger open-labeled study more recently (Irwin cancer.
et al., 2013) where 14 subjects with symptoms of depression or
depression mixed with anxiety warranting psychopharmacological
intervention received daily oral doses of ketamine hydrochloride 4.8.4. Biomarkers of ketamine's antidepressant effects
(0.5 mg/kg) over a 28-day period. The investigators found that over The general consensus in the literature appears to be that
the 28-day trial there was significant improvement in both depres- ketamine itself may never reach mainstream use as a treatment
sive symptoms and symptoms of anxiety for the eight subjects that for TRD or depression in bipolar affective disorder. However many
completed the trial. Side effects were rare; the most common being researchers believe that this body of research will be a means to
diarrhea, trouble sleeping, and trouble sitting still. The response rate learn more about the pathophysiology of depression that will
for depression in this study is similar to those found with IV hopefully offer novel therapeutic agents that act in a similar
ketamine; however, the time to response is more protracted. The fashion to ketamine with longer duration of action and less side-
findings of the potential efficacy of oral ketamine for depression and effects or that act on the effectors of ketamine known to be
the response of anxiety symptoms are novel. Further investigation involved in its mechanism of action. Translational research is
with randomized, controlled clinical trials is necessary to firmly currently exploring the identification of biomarkers that might
establish the efficacy and safety of oral ketamine for the treatment of be involved in prevention, diagnosis, treatment response, severity
depression and anxiety in patients receiving hospice care or other and prognosis of depression and Zarate et al., (2013b) reviewed
subject populations. the human biomarkers of antidepressant effects and summarise a
Another recent study of sublingual ketamine in 26 patients range of potential biomarkers identified from genetic, functional
with refractory unipolar and bipolar depression (Lara et al., 2013) neuroimaging, sleep and clinical studies implicated in ketamine's
demonstrated clear and sustained antidepressant effects on mood, antidepressant action. While, overall, while these biomarker
cognition and sleep in 20 (77%) with only mild and transient light- studies have limited practical implications at present they offer
headedness as a common side-effects (no euphoria, psychotic or huge hope that eventually we will be able to successfully apply
dissociative symptoms). Indeed remission remained in some these techniques to clinical populations. Moving research in this
patients after stopping ketamine, demonstrating again the rapid direction offers opportunities that would ultimately facilitate the
onset of action, high-efficacy and good tolerability of ketamine development of personalised treatment, increasing the probability
that may allow extended treatment as needed. There are other of success.
studies on-going assessing the oral route of administration. The
intranasal route of administration of ketamine is also been
evaluated as it has previously been shown to benefit analgesic-
5. Conclusion
refractory chronic pain patients at a dose comparable to that used
in most IV ketamine studies (Carr et al., 2004).
Subanaesthetic doses of intravenous ketamine offers a unique
opportunity to learn more about the psychopharmacological
4.8.2. Stereoisomers of ketamine dysregulation of the glutamatergic system in depression and there
Ketamine is a chiral compound and its R- and S-stereoisomers have been significant developments in this field in the last decade.
have different binding affinities. S-ketamine has a greater affinity Perhaps the greatest opportunity that emerges from this exciting
for the phencyclidine site of the NMDA receptor than R-ketamine research, however, is to conduct further studies that elucidate the
(Kohrs and Durieux, 1998). Higher incidence of psychedelic side mechanism of action of ketamine's rapid response to determine if
effects as described above has been reported with R-ketamine NMDA receptor antagonists such as ketamine, S-ketamine or
when given in similar analgesic doses as S-ketamine (Liu et al. synthetic analogues of ketamine could play a role in the future
2006; Raeder and Stenseth, 2000). A case series of 4 patients clinical treatment of depression. Also if by directly targeting AMPA
Paslakis et al., (2010) were administered oral S-ketamine at a dose receptors, eEF2K, mTOR activation or GSK-3 suppression, there
of 1.25 mg/kg for 14 days alongside traditional antidepressants in may be a distinct possibility that one would bring about the
an effort to shorten time to onset of action and in 2 of these possibility of fundamentally changing the treatment of depression
patients there was a rapid and sustained improvement in mood by creating an opportunity for rapid relief. If these rapid-acting
within the first week of treatment however the limited number of antidepressants could be safely integrated into treatment, one
patients do not allow conclusions about efficacy. Oral S-ketamine might shorten or mitigate hospitalisation, prevent lost work or
was well tolerated. Paul et al., (2009) described the effect of school days, reduce suicide and reduce healthcare costs.
M. Naughton et al. / Journal of Affective Disorders 156 (2014) 24–35 33

Role of funding source Drevets, W., Furey, M.L., 2010. Replication of scopolamine's antidepressant efficacy
The Alimentary Pharmabiotic Centre is a research centre funded by Science in major depressive disorder: a randomized, placebo-controlled clinical trial.
Foundation Ireland (SFI), through the Irish Government's National Development Biol. Psychiatry 67, 432–438.
Plan. This publication is supported in part by a research grant from SFI under Grant Du, J., Suzuki, K., Wei, Y., et al., 2007. The anticonvulsants lamotrigine, riluzole, and
Number SFI/12/RC/2273 and by the Health Research Board (HRB) through a Health valproate differentially regulate AMPA receptor membrane localization: rela-
Research Award (grant no. HRA_POR/2011/23; T.G.D., J.F.C. and G.C.). The Centre tionship to clinical effects in mood disorders. Neuropsychopharmacology 32
was previously funded by GlaxoSmithKline. J.F.C. is also funded by the European (4), 793–802.
Duman, R.S., Li, N., Liu, R.J., et al., 2012. Signalling pathways underlying the rapid
Community's Seventh Framework Programme (grant no.: FP7/2007–2013, grant
antidepressant actions of ketamine. Neuropharmacology 62, 35–41.
agreement 201 714). G.C. is supported by a NARSAD Young Investigator Grant from
Entsuah, A.R., Huang, H., Thase, M.E., 2001. Response and remission rates in
the Brain and Behavior Research Foundation (Grant Number 20771).
different subpopulations with major depressive disorder administered venla-
faxine, selective serotonin reuptake inhibitors, or placebo. J. Clin. Psychiatry 62,
869–877.
Conflict of interest Fatemi, S.H., Earle, J.A., Stary, J.M., et al., 2001. Altered levels of the synaptosomal
None. associated protein SNAP-25 in hippocampus of subjects with mood disorders
and schizophrenia. Neuroreport 12, 3257e62.
Fava, M., Rush, A.J, Wisniewski, S.R., et al., 2006. A comparison of mirtazepine and
nortriptyline following two consecutive failed medication treatments for
Acknowledgements depressed outpatients: a StarnD Report. Am. J. Psychiatry 163, 1161–1172.
The authors would like to acknowledge the significant contribution of Berne- Gigliucci, V., O'Dowd, G., Casey, S., et al., 2013. Ketamine elicits sustained
niece Riedewald in the drawing of Fig. 1. antidepressant activity via a serotonin dependant mechanism. Psychopharma-
cology 228, 157–166.
Goforth, H.W., Holsinger, T., 2007. Rapid relief of severe major depressive disorder
References by use of preoperative ketamine and electroconvulsive therapy. J. ECT 23,
23–25.
Green, C.D., 1973. Ketamine as an anaesthetic for ECT. Br. J. Psychiatry 122, 123–124.
Aan het Rot, M., Collins, K.A, Murrough, J.W, et al., 2010. Safety and efficacy of Ibrahim, L., Diazgranados, N., Luckenbaugh, D.A., et al., 2011. Rapid decrease in
repeated dose intravenous ketamine for treatment resistant depression. Biol. depressive symptoms with an N-methyl-D-aspartate antagonist in ECT-resistant
Psychiatry 67, 139–145. major depression. Prog. Neuropsychopharmacol. Biol. Psychiatry 35, 1155–1159.
Aan het Rot, M., Zarate, C.A, Charney, D.S, et al., 2012. Ketamine and depression: Ibrahim, L., DiazGranados, N., Franco-Chaves, J., et al., 2012. Course of improvement
where do we go from here? Biol. Psychiatry 72, 537–547. in depressive symptoms to a single intravenous infusion of ketamine vs. add-on
Abdallah, C.G, Fasula, M., Kelmendi, B., et al., 2012. Rapid antidepressant effect of riluzole: results from a 4-week, double blind, placebo-controlled study.
ketamine in the electroconvulsive therapy setting. J. ECT 28 (3), 157–161. Neuropsychopharmaclogy 37, 1526–1533.
Akinfiresoye, L, Tizabi, Y., 2013. Antidepressant effects of AMPA and ketamine Irwin, S.A., Iglewicz, A., 2010. Oral ketamine for the rapid treatment of depression
combination: role of hippocampal BDNF, synaptic and mTOR. Psychopharma- and anxiety in patients receiving hospice care. J. Palliat. Med. 13 (7), 903–908.
cology (Berl) 230 (2), 291–298. Irwin, S.A., Iglewicz, A., Nelesen, R.A., et al., 2013. Daily oral ketamine for treatment
Andreasen, JT, Gynther, M, Rygaard, A., et al., 2013. Does increasing the ratio of of depression and anxiety in patients receiving hospice care: a 28 day open-
AMPA-to-NMDA receptor medicated neurotransmission engender antidepres- label proof-of-concept trial. J. Palliat. Med. 8, 958–965.
sant action? Studies in the mouse forced swim and tail suspension tests. Jourdi, H., Yu-Tien, H., Zhou, M., et al., 2009. Positive AMPA receptor modulation
Neurosci. Lett. 546, 6–10. rapidly stimulates BDNF release and increases dendritic mRNA translation. J.
Autry, A.E, Adachi, M., Nosyreva, E., et al., 2011. NMDA Receptor blockade at rest Neurosci. 29 (27), 8688–8697.
triggers rapid behavioural antidepressant responses. Nature 475 (91), e95. Kessler, R., Berlund, P., Demler, O., et al., 2003. The epidemiology of major
Banasr, M., Valentine, G.W, Li, X.Y, et al., 2007. Chronic stress decreases cell depressive disorder: results for the National Comorbidity Survey Replication
proliferation in adult cerebral cortex of rat: reversal by antidepressant treat- (NCS-R). JAMA 289, 3095–3105.
ment. Biol. Psychiatry 62, 496–504. Klein, P.S., Melton, D.A., 1996. A molecular mechanism for the effect of lithium on
Barbon, A., Popoli, M., La Via, L., et al., 2006. Regulation of editing and expression of development. Proc. Nat. Acad Sci. U.S.A. 93, 8455–8459.
glutamate alpha-amino-propionic-acid (AMPA)/Kainate receptors by antide-
Kohrs, R., Durieux, M.E., 1998. Ketamine: teaching an old drug new tricks.
pressant drugs. Biol. Psychiatry 59, 713–720.
Anaesthesiol. Analgesia 87, 1186–1193.
Beck, A.T, Kovacs, M., Weissman, A., 1979. Assessment of suicidal intention: the
Kollmar, R., Markovic, K., Thurauf, N., et al., 2008. Ketamine followed by memantine
scale for suicide ideation. J. Consult. Clin. Psychol. 47, 343–352.
for the treatment of major depression. Aust. N.Z. J. Psychiatry 42, 170.
Berman, R.M, Cappiello, A., Anand, A., et al., 2000. Antidepressant effects of
Kranaster, L., Kammerer-Ciernioch, J., Hoyer, C., et al., 2011. Clinically favourable
ketamine in depressed patients. Biol. Psychiatry 47, 351–354.
effects of ketamine as an anaesthetic for electroconvulsive therapy: a retro-
Beurel, E., Song, L., Jope, R.S., 2011. Inhibition of glycogen synthase kinase-3 is
spective study. Eur. Arch. Psychiatry Clin. Neurosci. 261, 575–582.
necessary for the rapid antidepressant effects of ketamine in mice. Biol.
Krystal, J.H., Karper, L.P., Seibyl, J.P., et al., 1994. Subanesthetic effects of the non-
Psychiatry 16 (11), 1068–1070.
competitive NMDA antagonist, ketamine, in humans. Psychotomimetic, per-
Brewer, C.L., Davidson, J.R., Hereward, S., 1972. Ketamine (“Ketalar”): a safer
ceptual, cognitive, and neuroendocrine responses. Arch. Gen. Psychiatry 51,
anaesthetic for ECT. Br. J. Psychiatry 120, 679–680.
199–214.
Campbell, S., MacQueen, G., 2004. The role of the hippocampus in the pathophy-
Krystal, J.H., D'Souza, D.C., Karper, L., et al., 1999. Interactive effects of subanaes-
siology of major depression. J. Psychiatr. Neurosci. 29, 417–426.
thetic ketamine and haloperidol in healthy humans. Psychopharmacology
Carr, D.B., Goudas, L.C., Denman, W.T., et al., 2004. Safety and efficacy of intranasal
ketamine for the treatment of breakthrough pain in patients with chronic pain; (Berl) 145, 193–204.
a randomised, double-blind, placebo-controlled, cross-over study. Pain 108, Krystal, J.H., Petrakis, I.L., Krupitsky, E., et al., 2003. NMDA receptor antagonism and
17–27. the ethanol intoxication signal: form alcoholism risk to pharmacotherapy. Ann.
Chen, L.Y., Chen, K.P., Huang, M.C., 2009. Cystitis associated with chronic ketamine N.Y. Acad. Sci. 1003, 176–184.
abuse. Psychiatr. Clin. Neurosci. 63, 591. Krystal, A.D., Weiner, R.D., Dean, M.D., et al., 2003b. Comparison of seizure
Clemens, J., Nimmo, W.S., Grant, I.S., 1982. Bioavailability, pharmacokinetics, and duration, ictal EEG and cognitive effects of ketamine and methohexital
analgesic activity of ketamine in humans. J. Pharm. Sci. 71, 539–542. anaesthesia with ECT. J. Neuropsychiatry Clin. Neurosci. 15, 27–34.
Corrsen, G., Reves, J.G., Stanley, T.H., 1988. Dissociative anaesthesia. In: Corrsen, G., Krystal, J.H., Perry Jr, E.B., Gueorguieva, R., et al., 2005. Comparative and interactive
Reves, J.G, Stanley, T.H (Eds.), Intravenous Anaesthesia and Analgesia. Philadel- human psychopharmacologic effects of ketamine and amphetamine: implica-
phia: Lea and Febiger, pp. 99–173. tions for glutamatergic and dopaminergic model psychoses and cognitive
Cryan, J.F., Dev, K.K., 2007. The glutamatergic system as a potential therapeutic function. Arch. Gen. Psychiatry 62, 985–994.
target for the treatment of anxiety disorders. In: Blanchard, R.J., Blanchard, D.C., Krystal, J.H., Sanacora, G., Duman, R.S., 2013. Rapid-acting glutamatergic antide-
Griebel, G., Nutt, D.J. (Eds.), Handbook of Anxiety and Fear. Elsevier. pressants: the path to ketamine and beyond. Biol. Psychiatry 73, 1133–1141.
Cryan, J.F., O'Leary, 2010. A glutamate pathway to faster acting antidepressants. Lahti, A.C., Warfel, D., Michaelidis, T., et al., 2001. Long-term outcome of patients
Science 329, 913–914. who receive ketamine during research. Biol. Psychiatry 49, 869–875.
Diazgranados, N., Ibrahim, L., Brutsche, N.E., et al., 2010a. A randomised add-on trial Lanning, C.F., Harmel, M.H., 1975. Ketamine anaesthesia. Ann. Rev. Med. 26,
of an N-methyl-D-aspartate antagonist in treatment-resistant bipolar depres- 137–141.
sion. Arch. Gen. Psychiatry 67, 793–803. Lara, D.R., Bisol, L.W., Munari, L.R., 2013. Antidepressant, mood stabilising and
DiazGranados, N., Ibrahim, L.A., Brutsche, N.E., et al., 2010b. Rapid resolution of precognitive effects of very lose dose sublingual ketamine in refractory
suicidal ideation after a single infusion of an N-methyl-D-aspartate antagonist unipolar and bipolar depression. Int. J. Neuropsychopharmacol. 16 (9),
in patients with treatment-resistant major depressive disorder. J. Clin. Psychia- 2111–2117.
try 71 (12), 1605–1611. Larkin, G.L., Beautrais, A.L., 2011. A preliminary naturalistic study of low-dose
Dingledine, R., Borges, K., Bowie, D., et al., 1999. The glutamate receptor ion ketamine for depression and suicide ideation in the emergency department.
channels. Pharmacol. Rev. 51, 7–61. Int. J. Neuropsychopharmacol. 14 (8), 1127–1131.
Doble, A., 1996. The pharmacology and mechanism of action of riluzole. Neurology Li, X., Tizzano, J.P., Griffey, K., et al., 2001. Anti-depressant-like actions of an AMPA
47 (6 Suppl. 4), S233–S241. receptor potentiator (LY392098). Neuropharmacology 40, 1028–1033.
34 M. Naughton et al. / Journal of Affective Disorders 156 (2014) 24–35

Li, N., Lee, B., Liu, R.-J., et al., 2010. mTOR dependant synapse formation underlies Nutt, D.J., King, L.A., Nichols, D.E., 2013. Effects of schedule 1 drug laws on
the rapid antidepressant effects of NMDA antagonists. Science 329, 959–964. neuroscience research and treatment innovation. Nat. Rev. Neurosci. 14 (8),
Li, N., R0-J, Liu, Dwyer, J., et al., 2011. Glutamate N-methyl-D-aspartate receptor 577–585.
antagonists rapidly reverse behavioral and synaptic deficits caused by chronic O'Connor, R.M., Cryan, J.F, 2010. Role of metabotropic glutamate receptors in CNS
stress exposure. Biol. Psychiatry 69, 754–761. disorders. In: Siehler, S., Milligan, G. (Eds.), G Protein-Coupled Receptors.
Liu, R.J., Aghajanian, G.K., 2008. Stress blunts serotonin- and hypocretin-evoked Structure, Signaling, and Physiology. Cambridge University Press, pp. 321–379.
EPSCs in prefrontal cortex: role of corticosterone-mediated apical dendritic O'Connor, R.M., Finger, B.C., Flor, P.J., et al., 2010. Metabotropic glutamate receptor
atrophy. PNAS 105, 359–364. 7: at the interface of cognition and emotion. Eur. J. Pharmacol. 639, 123–131.
Liu, R.J., Fuchikami, M., Dwyer, J.M., et al., 2013. GSK-3 inhibition potentiates the O'Connor, R.M., Pusceddu, M.M., Dinan, T.G., et al., 2013. Impact of early-life stress,
synaptogenic and antidepressant-like effects of subthreshold doses of keta- on group III mGlu receptor levels in the rat hippocampus: effects of ketamine,
mine. Neuropsychopharmacology 38 (11), 2268–2277. electroconvulsive shock therapy and fluoxetine treatment. Neuropharmacology
Lorrain, D.S., Baccei, C.S., Bristow, L.J., et al., 2003. Effects of ketamine and 66, 236–241.
N-methyl-D-aspartate on glutamate and dopamine release in the rat prefrontal Okamoto, N., Nakai, T., Sakamoto, K., et al., 2010. Rapid antidepressant effect of
cortex: modulation by a group II selective metabotropic glutamate receptor ketamine anaesthesia during electroconvulsive therapy of treatment resistant
agonist LY379268. Neuroscience 117, 697e706. depression: comparing ketamine and propofol anaesthesia. J. ECT 26, 223–227.
Ma, X.-C., Dang, Y.-H., Jia, M., et al., 2013. Long-lasting antidepressant action of O'Neill, M.J., Bleakman, D., Zimmerman, D.M., et al., 2004. AMPA receptor
ketamine, but not glycogen synthase kinase-3 inhibitor SB216763, in the potentiators for the treatment of CNS disorders. Curr. Drug Targets CNS Neurol.
chronic mild stress model of mice. PLoS One 8 (2), e56053. Disord. 3, 181–193.
Macqueen, G., Yucel, K., Taylor, V.H., et al., 2008. Posterior hippocampal volumes Orrego, F., Villanueva, S., 1993. The chemical nature of the main central excitatory
are associated with remission rates in patients with major depressive disorder. transmitter: a critical appraisal based upon release studies and synaptic vesicle
Biol. Psychiatry 64, 880–883. localization. Neuroscience 56, 539–555.
Maeng, S., Zarate, C.A. Jr, Du, J., Schloesser, R.J., McCammon, J., Chen, G., Manji, H.K., Overall, J.E., Gorham, D.R., 1962. The brief psychiatric rating scale. Psychol. Rep. 10,
2008. Cellular mechanisms underlying the antidepressant effects of ketamine: 799–812.
role of alpha-amino-3-hydroxy-5-methylisoxazole-4-propionic acid receptors. Parise, E.M., Alcantara, L.F., Warren, B.L., 2013. Repeated ketamine exposure induces
Biol. Psychiatry 63, 349–352. an enduring resilient phenotype in adolescent and adult rats. Biol. Psychiatry 74
Manji, H.K., Quiroz, J.A., Sporn, J., et al., 2003. Enhancing neuronal plasticity and (10), 750–759.
cellular resilience to develop novel improved therapeutics for difficult to treat Parwani, A., Weiler, M.A., Blaxton, T.A., et al., 2005. The effects of a subanaesthetic
depression. Biol. Psychiatry 53, 707–742. dose of ketamine on verbal memory in normal volunteers. Psychopharmacol-
Martinez-Turrillas, R., Del Rio, J., et al., 2005. Sequential changes in BDNF mRNA ogy (Berl) 183, 265–274.
expression and synaptic levels of AMPA receptor subunits in rat hippocampus Paslakis, G., Gilles, M., Meyer-Lindenberg, A., et al., 2010. Oral administration of the
after chronic antidepressant treatment. Neuropharmacology 49, 1178–1188. NMDA receptor antagonist S-ketamine as add-on therapy of depression: a case
Mathew, S.J., Murrough, J.W., Aan het Rot, M., et al., 2010. Riluzole for relapse series. Psychopharmacology 43, 33–35.
Paul, R., Schaaff, N., Padberg, F., et al., 2009. Comparison of racemic ketamine and S-
prevention following intravenous ketamine in treatment-resistant depression:
ketamine in treatment–resistant major depression: report of two cases. World
a pilot randomised, placebo-controlled continuation trial. Int. J. Neuropsycho-
J. Biol. Psychiatry 10 (3), 241–244.
pharmacol. 13, 71–82.
Paul, I.A., Nowak, G., Layer, R.T., et al., 1994. Adaptation of the N-methyl-D-aspartate
McEwen, B., 2008. Central effects of stress hormones in health and disease:
receptor complex following chronic antidepressant treatments. J. Pharmacol.
understanding the protective and damaging effects of stress and stress
Exp. Ther. 269, 95–102.
mediators. Eur. J. Pharmacol. 583, 174–185.
Perry Jr., E.B., Cramer, J.A., Cho, H.S., et al., 2007. Yale ketamine study group.
McDaniel, W.W., Sahota, A.K., Vyas, B.V, et al., 2006. Ketamine appears associated
Psychiatric safety of ketamine in psychopharmacology research. Psychophar-
with better word recall than etomidate after a course of 6 electroconvulsive
macology (Berl) 192, 253–260.
therapies. J. ECT 22, 103–106.
Petrakis, I.L., Limoncelli, D., Gueorguieva, R., et al., 2004. Altered NMDA glutamate
Mizuta, I., Ohta, M., Ohta, K., et al., 2001. Riluzole stimulates nerve growth factor,
receptor antagonist response in individuals with a family vulnerability to
brain-derived neurotrophic factor and glial cell line-derived neurotrophic factor
alcoholism. Am. J. Psychiatry 161, 1776–1782.
synthesis in cultured mouse astrocytes. Neurosci. Lett. 310 (2–3), 117–120.
Phelps, L.E., Brutsche, N.E., Moral, J.R., et al., 2009. Family history of alcohol
Mjellem, N., Lund, A., Hole, K., 1993. Reduction of NMDA induced behaviour after
dependence and initial antidepressant response to an N-methyl-D-aspartate
acute and chronic administration of desipramine in mice. Neuropharmacology
antagonist. Biol. Psychiatry 65, 181–184.
32, 591–595.
Preskorn, S.H., Baker, B., Kolluri, S., et al., 2008. An innovative design to establish
Moghaddam, B., Adams, B., Verma, A., et al., 1997. Activation of glutamatergic
proof of concept of the antidepressant effects of the NR2B subunit selective N-
neurotransmission by ketamine: a novel step in the pathway from NMDA
methyl-D-aspartate antagonist, CP-101,606, in patients with treatment-
receptor blockade to dopaminergic and cognitive disruptions associated with
refractory major depressive disorder. J. Clin. Psychopharmacol. 28 (6), 631–637.
the prefrontal cortex. J. Neurosci. 17, 2921–2927. Price, R.B., Nock, M.K., Charney, D.S., et al., 2009. Effects of intravenous ketamine on
Monteggia, L.M., Gideons, E., Kavalali, E.T., 2013. The role of eukaryotic elongation explicit and implicit measures of suicidality in treatment-resistant depression.
factor 2 kinase in rapid antidepressant action of ketamine. Biol. Psychiatry 73, Biol. Psychiatry 66, 522–526.
1199–1203. Raeder, J.C., Stenseth, L.B., 2000. Ketamine: a new look at an old drug. Curr. Opin.
Morgan, C.J., Mofeez, A., Brandner, B., Bromley, L., Curran, H.V., 2004. Acute effects Anaesthesiol. 13, 463–468.
of ketamine on memory systems and psychotic symptoms in healthy volun- Reich, D.L., Silvay, G., 1989. Ketamine: an update on the first twenty-five years of
teers. Neuropsychopharmacology 29, 208–218. clinical experience. Can. J. Anaesth 36, 186–197.
Morgan, C.J., Curran, H.V., 2012. Ketamine use: a review. Addiction 107, 27–38. Rush, A.J., Trivedi, M.H., Wisniewski, S.R., et al., 2006a. Acute and longer term
Mott, D.D., Doherty, J.J., Zhang, S., et al., 1998. Phenylethanolamines inhibit NMDA outcomes in depressed outpatients requiring one or several treatment steps: a
receptors by enhancing proton inhibition. Nat. Neurosci. 1, 659–667. STARnD report. Am. J. Psychiatry 163, 1905–1917.
Muller, H.K., Wegener, G., Liebenberg, N., et al., 2013. Ketamine regulates the Rush, A.J., Trivedi, M.H., Wisiewski, S.R., et al., 2006b. Bupropion-SR, sertraline, or
presynaptic release machinery in the hippocampus. J. Psychiatr Res. 47 (7), venlafaxine-XR after failure of SSRIs for depression. N. Engl. J. Med. 354,
892–899. 1231–1242.
Murch, H., 2013. Ketamine, magnesium and major depression—from pharmacology Rush, A.J., 2006c. StarnD: what have we learned? Am. J. Psychol. 164, 201–204.
to pathophysiology and back. J. Psychiatr Res. 47 (7), 955–965. Sanacora, G., Zarate, C.A., Krystal, J.H., et al., 2008. Targeting the glutamatergic
Murrough, J.W., Iosifescu, D.V., Chang, L.C., et al., 2013. Antidepressant efficacy of system to develop novel, improved therapeutics for mood disorders. Nat. Rev.
ketamine in treatment-resistant major depression: a two site randomized Drug Discovery 7, 426–437.
controlled trial. Am. J. Psychiatry 2013 (170), 1134–1142. Scarr, E., Gray, L., Keriakous, D., et al., 2006. Increased levels of SNAP-25 and
Murrough, J.W., Perez, A.M., Stern, J., et al., 2013b. Rapid and longer-term synaptophysin in the dorsolateral prefrontal cortex in bipolar I disorder. Bipolar
antidepressant effects of repeated ketamine infusions in treatment-resistant Dis. 8, 133e43.
depression. Biol. Psychiatry 74, 250–256. Schumann, G., Johann, M., Frank, J., et al., 2008. Systematic analysis of glutamatergic
Murrough, J.W., Wan, L., Iacoviello, B., et al., 2013c. Neurocognitive effects of neurotransmission genes in alcohol dependence and adolescent risky drinking
ketamine in treatment-resistant major depression: association with antide- behaviour. Arch. J. Psychiatry 65, 826–838.
pressant response. Psychopharmacology , http://dx.doi.org/10.1007/s00213- Segmiller, F., Rüther, T., Linhardt, A., et al., 2013. Repeated S-ketamine infusions in
013-3255-x. therapy resistant depression: a case series. J. Clin. Pharmacol. 53 (9), 996–998.
Niciu, M.J., Luckenbaugh, D.A., Ionescu, D.F., et al., 2013. Subanesthetic dose Shansky, R., Morrison, J.H., 2009. Stress-induced dendritic remodeling in the medial
ketamine does not induce an affective switch in three independent samples prefrontal cortex: effects of circuit, hormones and rest. Brain Res. 2009,
of treatment-resistant major depression. Biol. Psychiatry 74 (10), e23–e24. 108–113.
Nierenberg, A., Fava, M., Trivedi, M.H., et al., 2006. A comparison of lithium and T3 Shor, B., Gibbons, J.J., Abraham, R.T., 2009. Targeting mTOR globally in cancer:
augmentation following two failed medication treatments for depression: a thinking beyond rapamycin. Cell Cycle 8, 3831–3837.
StarnD report. Am. J. Psychiatry 163, 1519–1530. Skolnick, P., Layer, R.T., Popik, P., Nowak, G., Pauli, A., Trullas, R., 1996. Adaptation of
Nobler, M.S., Sacheim, H.A., Moeller, J.R., et al., 1997. Quantifying the speed of N-methyl-D-aspartate (NMDA) receptors following antidepressant treatment:
symptomatic improvement with electroconvulsive therapy: comparison of implications for the pharmacotherapy of depression. Pharmacopsychiatry 29,
alternative statistical methods. Convuls. Ther. 13, 208–221. 23–26.
M. Naughton et al. / Journal of Affective Disorders 156 (2014) 24–35 35

Skolnick, P., Legutko, B., Li, X., et al., 2001. Current perspectives on the development Wilner, P., 2005. Chronic mild stress (CMS) revisited: consistency and behavioural-
of non-biogenic amine-based antidepressants. Pharmacol. Res. 43, 311–423. neurobiological concordance in the effects of CMS. Neuropsychobiology 52,
Südhof, T.C., 2004. The synaptic vesicle cycle. Ann. Rev. Neurosci. 27, 509e47. 90–110.
Svenningsson, P., Tzavara, E.T., Witking, J.M., et al., 2002. Involvement of striatal and Wong, S.W., Lee, K.F., Wong, J., et al., 2009. Dilated common bile ducts mimicking
extrastriatal DARPP-32 in biochemical effects of fluoxetine (Prozac). Proc. Nat. choledochal cysts in ketamine abusers. Hong Kong Med. J. 15, 53–56.
Acad. Sci. U.S.A. 99, 3182–3187. Yang, C., Zhou, Z., Yang, J., 2011. Be prudent of ketamine in treating resistant
Tan, C.H, He, X., Yang, J., 2006. Changes in AMPA subunit expression in the mouse depression in patients with cancer. J. Palliat. Med. 14 (5), 537.
brain after chronic treatment with the antidepressant maprotiline: a link Zarate Jr., C.A., Singh, J.B., Carlson, P.J., et al., 2006. A randomized trial of an N-
between noradrenergic and glutamatergic function? Exp. Brain Res. 170, methyl-D-aspartate antagonist in treatment resistant major depression. Arch.
448–456. Gen. Psychiatry 63, 856–864.
Thase, M.E., Haight, B.R., Richard, N., et al., 2005. Remission rates following Zarate, C.A., Singh, J., Quiroz, J.A., et al., 2006b. A double-blind, placebo-controlled
antidepressant therapy with bupropion or selective serotonin reuptake inhibi- study of memantine in the treatment of major depression. Am. J. Psychiatry
163, 153–155.
tors: a meta-analysis of original data from 7 randomised controlled trials. J.
Zarate Jr, C.A., Manji, H.K., 2008. The role of AMPA receptor modulation in the
Clin. Psychiatry 66, 974–981.
treatment of neuropsychiatric diseases. Exp. Neurol. 211, 7–10.
Tizabi, Y., Bhatti, B.H., Manaye, K.F., et al., 2012. Antidepressant-like effects of low
Zarate Jr, C.A., Machado-Viera, R., Henter, I., et al., 2010. Glutamatergic modulators:
ketamine dose is associated with increased hippocampal AMPA/NMDA receptor
the future for treating mood disorders? Harv. Rev. Psychiatry 18 (5), 293–303.
density ratio in female Wistar-Kyoto rats. Neuroscience 213, 72–80. Zarate Jr, C.A., Brutsche, N.E., Ibrahim, L., et al., 2012. Replication of ketamine's
Trivedi, M.H., Fava, M., Wisniewski, S.R., et al., 2006. Medication augmentation after antidepressant efficacy in bipolar depression: a randomised controlled add-on
the failure of SSRIs for depression. N. Engl. J. Med. 354, 1243–1252. trial. Biol. Psychiatry 71, 939–946.
Trullas, R., Stolnick, P., 1990. Functional antagonists at the NMDA receptor complex Zarate Jr, C.A., Mathews, D., Ibrahim, L., et al., 2013. A randomised trial of a low-
exhibit antidepressant actions. Eur. J. Pharmacol. 185 (1), 1–10. trapping nonselective N-methyl-D-aspartate channel blocker in major depres-
Weiner, R.D., 2001. The Practice of Electroconvulsive Therapy: Recommendations sion. Biol. Psychiatry 74 (4), 257–264.
for Treatment, Training and Privileging: A Task Force Report of the American Zarate, C., Mathews, D.C., Furey, M.L., 2013b. Human biomarkers of rapid anti-
Psychiatric Association, second ed. American Psychiatric Publishing, Inc, depressant effects. Biol. Psychiatry 73, 1142–1155.
Arlington, VA. (American Psychiatric Association Committee on Electroconvul- Zunszain, P.A., Horowitz, M.A., Cattaneo, A., et al., 2013. Ketamine: synaptogenesis,
sive Therapy). immunomodulation and glycogen synthase kinase-3 as underlying mechanisms of
White, P.F., Way, W.L., Trevor, A.J., 1982. Ketamine—its pharmacology and ther- its antidepressant properties. Mol. Psychiatry, http://dx.doi.org/10.1038/mp.2013.87.
apeutic uses. Anaesthesiology 56, 119–136. ([Epub ahead of print] PubMed PMID: 23877835).

You might also like