The American Psychiatric Association (APA) has updated its Privacy Policy and Terms of Use, including with new information specifically addressed to individuals in the European Economic Area. As described in the Privacy Policy and Terms of Use, this website utilizes cookies, including for the purpose of offering an optimal online experience and services tailored to your preferences.

Please read the entire Privacy Policy and Terms of Use. By closing this message, browsing this website, continuing the navigation, or otherwise continuing to use the APA's websites, you confirm that you understand and accept the terms of the Privacy Policy and Terms of Use, including the utilization of cookies.

×
Clinical SynthesisFull Access

Advances in Psychopharmacology for Anxiety Disorders

Abstract

Anxiety disorders are among the most prevalent psychiatric disorders, but compared with mood and psychotic disorders, are understudied in terms of newer pharmacotherapeutic approaches. Certain anxiety disorders, such as generalized anxiety disorder, and related conditions such as posttraumatic stress disorder (PTSD), may be refractory to the approved first-line treatments, like selective serotonin reuptake inhibitors (SSRIs) and serotonin norepinephrine reuptake inhibitors (SNRIs). Here we will review the most recent pharmacological treatment modalities under investigation for anxiety disorders and related conditions such as PTSD. The review includes a discussion of neurotransmitter and peptide pathways. We will present novel treatment options from medication classes that are widely studied in anxiety, such as the serotonin and gamma-aminobutyric acid (GABA) system, and also more novel mechanisms like glutamate modulators (e.g., ketamine, riluzole and d-cycloserine), corticotropin releasing factor and vasopressin receptor antagonists, neuropeptides such as neuropeptide Y and oxytocin, anticonvulsants such as pregabalin and gabapentin, mifepristone, and adrenergic agents such as propranolol and prazosin. Our review of the literature suggests that while there are some agents under investigation that may appear promising in the future, most of them are relatively early in development and there are very few new medications that carry immediate promise for the treatment of anxiety disorders. We hope this review encourages further investigation of novel therapeutics for anxiety, focusing primarily on new drug development of nonmonoamine and peptide systems.

Introduction

Anxiety disorders are the most common class of psychiatric disorders (1). They are commonly associated with functional impairment (2) and are, along with depression, among the leading causes of disability and work absences (3). The cost burden of anxiety disorders may also be greater than any other psychiatric disorders, even mood disorders, due to their high prevalence and the increased costs of medical and psychiatric treatment (4, 5). In spite of this, compared with mood, schizophrenia spectrum, and autism spectrum disorders, there has been less research for novel therapeutics for the anxiety disorders in the last two decades. The most recent Food and Drug Administration (FDA)-approved treatments for anxiety disorders have been repurposed antidepressant treatments. Currently, the mainstays of pharmacological treatment for anxiety disorders are selective serotonin reuptake inhibitors (SSRIs), serotonin norepinephrine reuptake inhibitors (SNRIs), and benzodiazepines. Anxiety is responsive to several forms of psychotherapy including cognitive behavioral therapy (CBT), but medication treatments will be the focus of this review.

Anxiety disorders, as described in the Diagnostic and Statistical Manual, Fourth Edition (DSM-IV-TR), consisted of the following diagnoses (if one were to exclude Not Otherwise Specified and Anxiety Disorders due to Substances or Medical Conditions): generalized anxiety disorder (GAD), obsessive-compulsive disorder (OCD), panic disorder (PD) with or without agoraphobia, posttraumatic stress disorder (PTSD), social phobia, and specific phobia (6). The DSM-5 includes a total of nine distinct anxiety disorders (see Table 1). It classifies OCD under its own class of disorders, called obsessive-compulsive and related disorders, while PTSD is listed under trauma- and stressor-related disorders (7). For the purpose of this review, we will not discuss OCD, a distinct disorder in terms of pathophysiology and treatment, while PTSD will be included as newer therapeutics for PTSD overlap with other anxiety disorders. Additionally, anxiety due to a substance or medical disorder will not be discussed as the treatment often involves management of the underlying cause.

Table 1. Anxiety Disorders
AgoraphobiaNow a distinct disorder in DSM-5
Generalized Anxiety Disorder

Panic Disorder

Selective Mutism
Newly classified in DSM-5
Separation Anxiety Disorder
Newly classified in DSM-5
Social Anxiety Disorder (Social Phobia)

Specific Phobia

Anxiety Disorder Due to Another Medical Condition

Substance/Medication-Induced Anxiety Disorder
Diagnoses merged in DSM-5
Other Specified Anxiety Disorder
Split from “Not Otherwise Specified” from DSM-IV-TR
Unspecified Anxiety Disorder
Split from “Not Otherwise Specified” from DSM-IV-TR
Table 1. Anxiety Disorders
Enlarge table

The current treatments for anxiety disorders are summarized in Table 2. It is notable that, with the exception of specific phobia, the first line treatment for all anxiety disorders are SSRIs and SNRIs. Previously, tricyclic antidepressants (TCAs) were the favored choice for PD and OCD but have been relegated to second-line treatment due to concerns about side effects, including anticholinergic effects, weight gain, sedation, cardiac effects, and the danger of death in overdose. Monoamine oxidase inhibitors (MAOIs) have proven efficacy in anxiety disorders but their use is limited by dietary restrictions, drug interactions, and side effects. Benzodiazepines, a medication class dating back to the 1950s, were considered a first-line treatment for PD and generalized anxiety as well, but due to concerns about tolerance and dependence, it is recommended that they be used judiciously, especially in patients with a past substance abuse history and in special populations (children/adolescents, elderly, and medically ill persons). Finally, alternative treatments that are currently not approved by the FDA include anticonvulsants, including gabapentin and pregabalin, and antipsychotics including quetiapine. Although there is limited evidence to support their use, they may be appropriate in cases of refractory anxiety, with the caveat that tolerability and safety may be a concern in the case of antipsychotics.

Table 2. Current Treatments for Anxiety Disorders
DisorderFDA-Approved MedicationsOff-Label MedicationsPsychotherapyRecommendations
Generalized Anxiety
SSRI: escitalopram; paroxetine SNRI: duloxetine; venlafaxine (XR)
SSRI: fluoxetine; sertraline; citalopram TCAs MAOIs Mirtazapine Pregabalin Gabapentin Hydroxyzine Antipsychotics: quetiapine
CBT
First-Line: SSRI or SNRI Second-Line: benzodiazepines; buspirone; TCAs; pregabalin CBT
Panic Disorder
SSRI: fluoxetine; paroxetine; sertraline SNRI: venlafaxine (XR) Benzodiazepines: clonazepam; alprazolam TCAs: clomipramine; imipramine
SSRI: escitalopram SNRI: duloxetine; desvenlafaxine TCAs MAOIs Phenelzine Pregabalin Gabapentin Benzodiazepines: lorazepam
CBT Psychodynamic
First-Line: CBT or SSRI, SNRI or Combination
Social Anxiety Ddisorder
SSRI: paroxetine; sertraline; fluvoxamine SNRI: venlafaxine (XR)
SSRI: fluoxetine MAOIs Pregabalin Gabapentin Benzodiazepines D-cycloserine Propranolol
Exposure therapy CBT
First-Line: SSRI or SNRI Exposure therapy Second-Line: MAOIs
Specific Phobias
None
D-cycloserine
Exposure therapy
First-Line: exposure therapy
Posttraumatic Stress Disorder
SSRI: paroxetine; sertraline
SSRI: fluoxetine SNRI: venlafaxine (XR); duloxetine; desvenlafaxine TCAs MAOIs Mirtazapine Prazosin Antipsychotics: risperidone Anticonvulsants: lamotrigine; topiramate; oxcarbazepine
Prolonged exposure therapy
First-Line: SSRI and/or exposure therapy Second-Line: antipsychotics
Anxiety
Benzodiazepines: clonazepam; alprazolam; chlordiazepoxide; oxazepam Antipsychotics: trifluoperazine Buspirone Hydroxyzine
Mirtazapine TCAs Antipsychotics: olanzapine; quetiapine Pregabalin Gabapentin Tiagabine


Table 2. Current Treatments for Anxiety Disorders
Enlarge table

Although anxiety disorder may respond to appropriate first- and second-line pharmacologic treatments with or without psychotherapy, there is reasonable concern about the dearth of options for the large number of patients who are nonresponsive to current treatment options. Most of the current pharmacotherapies for anxiety disorders act via the monoamine system (e.g., SSRIs, SNRIs, TCAs, and MAOIs) or as agonists of gamma-aminobutyric acid (GABA-A) receptors (the benzodiazepines). There is a clear need for more efficacious treatments with novel mechanisms. We will therefore focus on the promising next-generation antianxiety treatments, which are currently in development, discussing them in the context of their putative mechanisms of action.

Novel Treatments

While the previous focus of pharmacological research in anxiety disorders was on the role of the serotonin system, there has been a shift in experimental therapeutics in the last 20 years toward other novel pathways such as cannabinoid, glutamate, and neuropeptide systems (8). These agents have been tested in preclinical studies using animal models of anxiety such as classical fear conditioning and elevated plus-maze; a few of them are undergoing testing in early human studies. We will present these compounds and the current state of their development; for agents currently undergoing testing in the United States we will include the National Clinical Trial (NCT) numbers from clinicaltrials.gov.

Serotonin

The serotonin system has been reported to be a primary mediator in the etiology of anxiety and its disorders. As noted in the Introduction and in Table 2, the majority of treatments for anxiety, including SSRIs, SNRIs, and azapirones like buspirone, are serotonergic. Thus, there is interest in therapeutics to develop agents that target specific serotonin receptors for an anxiolytic effect. For example, buspirone, a 5-HT1A agonist, is being investigated for depressive symptoms and neuroprotection in GAD (NCT01546896). Given the documented efficacy of buspirone for GAD, and preclinical research, there have been other nonazapirone 5-HT1A agonists under development for anxiety treatment (9), but the results so far have been disappointing. Gepirone (BMY-13,805, ORG-13,011), a selective 5-HT1A receptor partial agonist, has undergone clinical development as an anxiolytic and antidepressant agent. Early studies suggested efficacy for anxiety symptoms in MDD subjects (10). However, it failed to receive FDA approval for the treatment of either anxiety or depression. PRX-00023, a selective 5-HT1A partial agonist, did not show superior efficacy on anxiety measures compared with placebo in an RCT, although modest efficacy on some depressive measures was noted (11). Additionally, there is a phase 2 trial of another selective 5-HT1A partial agonist, TGFK08AA, in phase 2 development for GAD (12).

There are also medications that act on multiple serotonin receptors that are being studied for anxiety. The SSRI vilazodone, which is approved by the FDA for MDD, is also a 5-HT1A receptor agonist and is under study in a RCT for social anxiety disorder (NCT01712321). Vilazodone is also in phase 3 trials for GAD (NCT01844115, NCT01766401, NCT01629966) and in a phase 4 trial for treatment of PTSD with comorbid depression (NCT01715519). Vortioxetine (Lu AA21004) is also a 5-HT1A agonist and a 5-HT3 antagonist, and was recently studied for MDD and GAD on the basis of preclinical studies supporting its antidepressant and anxiolytic effects (13). Vortioxetine was recently FDA-approved for MDD, but its efficacy for anxiety is not yet demonstrated. While one RCT reported efficacy of vortioxetine in reducing anxiety in GAD (14) and another reported efficacy in preventing relapse in GAD (15), three subsequent randomized, double-blind reports indicated that vortioxetine did not separate from placebo in GAD on primary or secondary outcome measures (1618).

5-HT6 receptor antagonists have been reported to have anxiolytic properties in animal studies (19). There are no known published trials of these compounds in humans although there are two novel drugs under development, AVN-101 and AVN-397, which are being studied for the treatment of anxiety (8).

Finally, agomelatine is a melatonin-1/melatonin-2 agonist and 5-HT2C receptor antagonist that has been studied in depression but also may have anxiolytic properties (20). Agomelatine was reported to be efficacious for GAD compared with placebo (21), and in a randomized, double-blind discontinuation study was reported, to have a lower rate of relapse compared with placebo in GAD (22). There are no other known ongoing studies of agomelatine in anxiety disorders.

Glutamate Modulators

Glutamate, which is a precursor for GABA, is the primary excitatory neurotransmitter of the central nervous system. Glutamate is comprised of ionotropic and metabotropic receptors (mGluR), with the latter being a target of pharmacological treatments for mood, psychotic, and anxiety disorders. Several animal studies have reported anxiolytic effects of mGluR modulators, in particular mGluR1, mGluR2, mGlu3, and mGluR5 (23). However, results in human studies have been more disappointing. For example, LY354740, an mGluR 2–3 agonist, did not show a difference from placebo in a randomized controlled comparison study with paroxetine (24) in patients with PD. After studies with LY354740 were halted due to concerns about bioavailability, LY544344, a pro-drug of LY354740, was studied for GAD in an 8-week randomized placebo-controlled trial and found to be efficacious, but the study was halted due to concerns about convulsive activity in preclinical trials (25). Currently there are studies underway for two mGluR modulators: ADX-71149, an mGluR2 positive allosteric modulator (NCT01582815), in phase 2, tested for major depressive disorder with anxiety symptoms; and, RGH-618, an mGluR1/5 antagonist with promising preclinical data, currently undergoing phase 1 studies and potential future development for (unspecified) anxiety disorders (8).

Ketamine is a glutamate N-methyl-d-aspartate (NMDA) receptor antagonist used in anesthesia and pain, and also used as a substance of abuse with hallucinogenic and psychotomimetic properties. Over the past 10 years several studies have reported rapid antidepressant effects after a single administration of IV ketamine in treatment-resistant major depressive disorder (TRD) (26, 27). Multiple trials are ongoing using intravenous and intranasal ketamine for MDD and bipolar depression (28, 29), but ketamine is also currently being developed as a treatment for anxiety (including clinical trials in PTSD and OCD). Preclinical studies have reported the anxiolytic effects of ketamine in several animal models of anxiety, including the elevated-plus maze (30). A recently completed study (NCT00749203) in 40 subjects with PTSD involved administration of IV ketamine 0.5 mg/kg and midazolam in a crossover design. Compared with IV midazolam, IV ketamine was associated with superior efficacy in reducing PTSD symptoms at 24 hours (as measured with the self-report impact of event scale) (31). Additionally, there are open-label studies of oral ketamine for the treatment of comorbid depression and anxiety (32, 33).

Riluzole is a glutamate modulator approved for the treatment of amyotrophic lateral sclerosis (ALS), which has also been studied as an adjunctive agent in TRD (3436), with trials currently ongoing for MDD, OCD, and multiple neurological conditions, given riluzole’s neuroprotective properties. Animal studies supported the efficacy of riluzole in models of anxiety (37, 38). We only found a single clinical therapeutic trial with riluzole in the literature: Mathew and coworkers enrolled 18 subjects with GAD in an 8-week study with open-label riluzole 100 mg/day (39). Twelve of the 15 patients (80%) who completed the trial responded [experiencing ≥50% improvement in Hamilton Anxiety Rating Scale (HAM-A) anxiety scores] and eight of 15 patients (53%) experienced remission of anxiety (39). The study was, however, limited by open design and a low number of subjects. Subsequent functional imaging studies reported that patients with GAD who were openly treated with riluzole experienced changes in hippocampal volumes and N-acetylaspartate (NAA) concentrations which correlated with improvement on anxiety scales, when compared with a treatment group of healthy volunteers (40, 41). A phase 2 randomized, double-blind, placebo-controlled study of riluzole in PTSD is currently beginning at Yale University (NCT02019940).

AMPA (α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid) receptors are ionotropic glutamate receptors. There are several animal studies of AMPA modulators, including PEPA, primarily in fear extinction models of anxiety, showing positive anxiolytic effects (42). One AMPA modulator, Org 26576, has been studied for MDD (43). To date, however, there are no known studies in the pipeline for AMPA modulators for anxiety disorders.

Among glutamatergic drugs, one of the best studied in anxiety disorders is d-cycloserine (DCS), an NMDA partial agonist, previously used as a treatment for tuberculosis and studied in schizophrenia. In animal and human studies DCS has been shown to facilitate fear extinction (44). Although it is not known to be efficacious as a monotherapy for anxiety (45), DCS has been successfully used in the augmentation of exposure or cognitive behavioral therapy, as discussed below. DCS has been reported to reduce anxiety in persons performing cognitive tasks (46), facilitate declarative learning (47), and to reduce reactivity to phobic stimuli in persons with a specific phobia on a functional MRI (48).

DCS has been studied for augmentation of psychotherapy in PD, social anxiety disorder, OCD, PTSD, and specific phobias. In PD, Otto and others reported that patients receiving DCS instead of placebo for augmentation of CBT had statistically significant improvement in clinical measures (49). Another study in PD, a randomized controlled trial (RCT), found no difference between DCS or placebo when added to CBT+flooding, but did note that DCS may speed up clinical improvement (50). We found one study actively recruiting for DCS augmentation of CBT in PD (NCT01680107), while another active study is investigating the effects of DCS enhancement, versus placebo, on CBT in PD (NCT00790868).

Several studies of DCS have been published in social anxiety disorder. Two earlier investigations reported a positive response of 50 mg DCS on the augmentation of exposure therapy in social phobia (51, 52). More recently, one study found that DCS did not augment the effect of CBT on sleep quality compared with placebo (53). Smits and coworkers reported the DCS was efficacious in the augmentation to CBT only in subjects with low conscientiousness and high agreeableness (54) or in those who reported successful reduction in fear after treatment (55). In another study, the same group of researchers found no separation between DCS and placebo in the remission or response rates, but did report a more rapid response in the DCS group (56). Given these negative results, it is not surprising that there are no known ongoing studies of DCS augmentation of social anxiety disorder.

d-cycloserine has also been studied extensively for specific phobias. While one study reported the efficacy of DCS in augmenting exposure therapy for acrophobia (57), a subsequent study did not find DCS to separate from placebo in postexposure treatment of height phobia (58). Two studies of spider phobia reported no difference between DCS and placebo but Gutner and colleagues reported a reduction in disgust in the DCS group to subliminal fear cues (59, 60). There is a trial currently recruiting to test DCS for augmentation of one-session exposure therapy for acrophobia (NCT01037101).

In an RCT with 25 subjects, DCS has been reported to be superior to placebo in improving PTSD symptoms when administered as augmentation to virtual reality exposure therapy (61). Previously, another group studying DCS augmentation for exposure therapy for PTSD found that DCS separated from placebo in participants with high conscientiousness and low extraversion, while both groups with lower education had worse outcomes (62). The same group reported DCS did not differentiate from placebo in its overall treatment effect but reported a greater response to DCS compared with placebo in patients with more severe PTSD at baseline and those that required longer treatment (63). Interestingly, one group reported worse outcomes for persons receiving DCS versus placebo for augmentation of exposure therapy (64). Additionally, there are currently several on-going studies of DCS in PTSD: a VA study comparing hydrocortisone, D-cycloserine, and placebo on fear extinction in veterans with PTSD (NCT00674570), a trial investigating DCS and mifepristone for blocking memory consolidation in PTSD (NCT01490697), and two studies of DCS in PTSD in youth and adolescents (NCT01157416, NCT01157429).

On balance, DCS is the most widely studied of the glutamatergic strategies and the literature thus far supports a modest benefit for the augmentation of psychotherapy. However, several studies do not detect this efficacy.

Memantine is an NMDA receptor antagonist approved for the treatment of moderate to severe Alzheimer dementia. It has been tested previously in animal studies as monotherapy and augmentation for depression and OCD (65). To date, it has only been studied openly in small trials. One study reported minimal improvement in seven patients with GAD, while 10 OCD subjects experienced a modest benefit (66). In another case series of 15 patients with GAD with or without comorbid social anxiety disorder, Schwartz et al. reported a reduction in anxiety from baseline (67). Another small open-label study reported an improvement in memory and hyperarousal in four patients with PTSD (68). There are no known active studies of memantine for anxiety disorders other than OCD.

Anticonvulsants

A previous review of anticonvulsants for anxiety disorders (69) reported strong evidence for use of certain medications in anxiety. The strongest evidence is for the use of pregabalin in GAD with positive separation from placebo in multiple placebo-controlled randomized trials (70, 71). Pregabalin is thought to act on the alpha-2 delta subunit of calcium channels to reduce neurotransmitter release, and thus has antiepileptic effects. Pregabalin has FDA approvals for neuropathic pain, postherpetic neuralgia, fibromyalgia, and as an adjunct treatment for partial seizures. It is also approved for use in GAD in Europe, but after the FDA did not approve its use in 2009, Pfizer withdrew its application in 2010. In one study, pregabalin was shown to have efficacy versus placebo in generalized social anxiety disorder, at 450 mg/day (72), while another study found pregabalin at 600 mg/day (but not at 450 mg/day or 300 mg/day) to be superior to placebo in a randomized, double-blind study (73).

Gabapentin has a mechanism of action similar to pregabalin. It is also approved for neuropathic pain, postherpetic neuralgia, and partial seizures, but has been widely used off-label for various indications, including anxiety. It was found to be efficacious in a small (N=69), randomized, double-blind, placebo-controlled study in social anxiety disorder (74). Another RCT found a treatment difference between gabapentin and placebo only in patients with severe panic symptoms (75). There are currently no known ongoing trials of gabapentin for anxiety disorders.

Tiagabine is an anticonvulsant that is thought to inhibit GABA uptake thereby increasing GABA activity. It is approved for partial seizures and there is preclinical evidence for its anxiolytic effects (76). While open-label studies for PD, GAD, and PTSD are suggestive of efficacy (69), several RCTs, however, do not support the efficacy of tiagabine in GAD (77) or PD (78). A small, randomized, double-blind crossover study with gabapentin and tiagabine in social anxiety disorder reported that both drugs could be effective in reducing the anxiety score (79).

Lamotrigine is an anticonvulsant thought to inhibit voltage-dependent sodium channels resulting in decreased glutamate release. It is FDA-approved for treatment of bipolar disorder and for seizures. In animal studies, lamotrigine has been shown to have anxiolytic properties (80). There are, however, very few studies of lamotrigine in anxiety disorders. One small case series reported improvement in symptoms in PD with agoraphobia (81). One small (N=15) randomized, double-blind, placebo-controlled trial of lamotrigine in PTSD, reported an improvement in symptoms, suggesting the need for further large-scale study (82). To date, however, no studies of lamotrigine for anxiety disorders are underway, with the exception of a phase 3 trial comparing topiramate and lamotrigine in persons with PTSD and alcohol dependence (NCT00571246).

Neuropeptides

Neuropeptides are small proteins playing the role of neuronal signaling molecules; they are involved in a variety of brain functions, including analgesia, reward, social behaviors, learning, and memory. Specific neuropeptides also play significant roles in modulating fear and anxiety.

Oxytocin is a neuropeptide with a significant role in attachment and prosocial behaviors in both animals and humans. It has been studied in autism and schizophrenia, given the severe dysfunction in interpersonal relationships in these disorders. There is evidence in healthy adults that oxytocin has positive effects on emotional modulation (83); in high-stress individuals it can reduce negative cognitive appraisals elicited by high-stress tasks (84). In subjects with generalized social anxiety disorder, oxytocin reportedly reduced amygdalar reactivity to fearful faces as measured with functional MRI (85).

Oxytocin has to be administered either sublingually or intranasally due to its poor absorption in the digestive tract. It is well-tolerated with no known serious side effects. A randomized, double-blind study comparing intranasal oxytocin to placebo as an augmentation to exposure therapy in social anxiety disorder did not report a difference in treatment outcomes between groups, but subjects receiving oxytocin did report more positive evaluations and opinions of their own performance compared with the control group (86). Currently, there is one known ongoing clinical trial of oxytocin nasal spray and its effects on social behavior in social anxiety disorder (NCT 01856530).

Substance P, also known as neurokinin-1 (NK-1), a neuropeptide widely found in the central nervous system, has also been associated with anxiety in preclinical studies. NK-1 receptor antagonists therefore have been studied for anxiety disorders. In a randomized, double blind trial, comparing it to paroxetine and placebo for social anxiety disorder, the NK-1 antagonist LY686017 did not separate from placebo on outcome measures (87). The selective NK-1 receptor antagonist GR205171 showed promise as a potential anxiolytic in an investigation of cerebral blood flow in patients with social anxiety disorder, with efficacy superior to placebo and comparable to citalopram (88). This same compound, however, did not separate from placebo in an 8-week RCT with 39 subjects with chronic PTSD (89). Another NK-1 receptor antagonist, L-759274, was reported to not be efficacious in the treatment of GAD in a proof-of-concept RCT (90). Currently, GW823296 (Orvepitant) is undergoing a phase 2 randomized, double-blind, placebo controlled fixed-dose study of for PTSD (NCT01000493).

Neuropeptide Y (NPY) is one of the most abundant neuropeptides in the brain. It has been shown in preclinical trials to be related to stress and anxiety responses, and treatment in animal studies have been shown to have anxiolytic effects (91). In humans, NPY has been closely linked to trauma, with multiple studies showing reduced plasma and cerebrospinal fluid levels of NPY in persons with PTSD as compared with healthy adults (92). There are two published studies of intranasal NPY, one as a tolerability study in healthy volunteers (93) and another in overweight subjects to test the effects on appetite and body weight regulation (94). Currently, our group at Icahn School of Medicine at Mount Sinai is conducting a phase 1, randomized, double-blind, dose escalation study of intranasal NPY in PTSD (NCT01533519).

Arginine vasopressin (AVP) has been shown in animal models to be related to anxiety responses, and vasopressin V1A and V1B receptor antagonists may have anxiolytic properties (95, 96). Only one published human study of V1 antagonists has been published. Griebel and coworkers compared a vasopressin V1B antagonist, SSR149415, in a randomized, double-blind study to escitalopram, paroxetine, and placebo in MDD and GAD, and found that SSR149415 did not separate from placebo in outcome measures for GAD (97). Currently under study is a novel V1A receptor antagonist, SRX246, which is being studied in healthy volunteers. In a functional MRI study it attenuated the effects of vasopressin on amygdala reactivity to angry faces (98). SRX246 is currently being investigated in a phase 1 trial for PTSD (8).

Cholecystokinin (CCK) is a peptide which helps regulate gastric secretions and motility and biliary function in the gastrointestinal system; in the nervous system it is involved in memory and in anxiety and fear processing. Preclinical research has supported the findings of the role of CCK-2 in anxiety and the potential use of CCK-2 antagonists as anxiolytics. Subsequent studies of these agents, however, failed to yield positive results (8). There are currently no CCK antagonists being studied for anxiety disorders.

In summary, neuropeptides represent a promising avenue in the treatment of anxiety disorders but no clear winner has yet emerged.

Corticotrophin-releasing factor (CRF) receptor antagonists have been reported to have anxiolytic effects in animal models (99). Their role is related to modulating the chronic hypercortisolemia associated with chronic stress; CRF dysfunction has been described across the entire spectrum of anxiety disorders (100). To date, however, there are few published investigations of CRF antagonists in anxiety. One randomized, double-blinded study of the CRF-1 antagonist Pexacerfont (BMS-562086), which compares it to escitalopram and placebo in GAD, reported that Pexacerfont did not separate from placebo on primary outcome measures (101). Currently ongoing, a multicenter phase 2 RCT of another CRF-1 antagonist, Verucerfont (GSK561679) is being studied in women with PTSD (NCT01018992). There are also completed but not published studies for social anxiety disorder comparing GSK561679 and the CRF-1 antagonist Emicerfont (GW876008) to alprazolam and placebo (NCT00555139), and a phase 2 trial comparing GW876008 to paroxetine and placebo (NCT00397722).

GABA Agonists

The efficacy of benzodiazepines, GABA-A agonists, in anxiety disorder is well documented but, as noted above, long-term risks make it important to develop newer GABAergic medications. To date, however, several GABA-A receptor subtype agonists have either failed to reach the market due to lack of efficacy or poor tolerability (8). AZD7325, a GABA-A alpha-2-3 modulator, failed to separate from placebo in a phase 2 comparative trial with placebo and alprazolam for GAD (NCT00808249). The novel GABA modulator IW-2143 (formerly BNC-210) is in phase 1 studies for anxiety (Bionomics).

Mifepristone

Mifepristone, also known as RU-486, is a progesterone inhibitor used for early pregnancy termination (hence the nickname “the morning-after pill”). Mifepristone has been studied for psychotic depression, and for improving cognition in bipolar disorder and schizophrenia. There is extensive literature on hypothalamic-pituitary-adrenal axis dysfunction in PTSD. Preclinical studies support the use of mifepristone for preventing fear reconsolidation (102), and the same group is studying this effect in humans with PTSD in a phase 4 trial (NCT01490697). There is one small (N=8) randomized, double-blind, placebo-controlled study of mifepristone in combat PTSD, which reported a reduction in scores in rating scales (103). Currently, there is a multicenter randomized, double-blind, placebo-controlled trial of mifepristone in combat-related PTSD (NCT01946685). One group is studying mifepristone for cognitive impairment in late-life anxiety disorders (NCT01333098).

Alpha- and Beta-Adrenergic Agents

Guanfacine is an alpha−2 receptor agonist approved for the treatment of hypertension, which was later found to have potential effects on attention and was approved for ADHD. However, it did not separate from placebo in two RCTs for PTSD (104, 105). An open-label trial of guanfacine in children and adolescents reported potential improvement in symptoms (106). To date, there are no known active studies of guanfacine in anxiety disorders.

Propranolol is a beta-blocker approved for multiple indications including hypertension, arrhythmias, migraine prophylaxis, and tremor. It has been widely used in the treatment of performance anxiety and social anxiety disorder but with limited evidence for its efficacy. There is also limited evidence for its effects in PD or other anxiety disorders. Propranolol gained interest as a treatment in PTSD after Pitman and colleagues published a study on the potential effect of preventing the development of PTSD when administering propranolol shortly after a traumatic event (107). Subsequent larger-scale studies did not support the initial findings of efficacy (108) and propranolol became a controversial topic due to concerns that it would “block” memories (109). Animal studies, however, have suggested that propranolol, due to its central activity, can prevent reconsolidation of traumatic memories through its effects on protein synthesis. Several studies have reported the potential efficacy of propranolol enhancement in PTSD (110). There is an ongoing phase 2 study of propranolol for the reduction of traumatic memories in PTSD (NCT01713556) and a study of the effects of propranolol on fear responses in anxiety and PTSD (NCT01631682).

Natural Remedies

Although an extensive literature exists on potential anxiolytic efficacy of more than 20 herbal compounds, for most of them the evidence base is still inconclusive. To date, the only compound with efficacy data supported by multiple quality RCT studies is kava, a plant from the South Pacific whose active ingredients are lipophilic resinous compounds named kavalactones. A Cochrane meta-analysis of six RCTs using kava in anxiety disorders revealed a statistically significant mean reduction of five points on the HAM-A over placebo (111). Another recent pooled analysis supports the use of kava in the treatment of anxiety, with a significant result occurring in four out of six RCTs reviewed (mean Cohen’s coefficient d=1.1) (112). However, kava can be associated with liver toxicity, including some cases of severe liver toxicity. The data for other compounds (e.g., galphimia, echinacea, chamomile, ginkgo, passionflower, roseroot, ashwagandha, Iranian borage, lemon balm, and milk thistle) are still insufficient to support their efficacy.

Discussion

Based on our review of novel therapeutics for anxiety disorders across multiple neurotransmitter and peptide pathways, it is evident that there still remains a paucity of efficacious treatment options for anxiety disorders. Despite the potential promise of several newer compounds, such as certain metabotropic glutamate receptor modulators, CFR antagonists, and vasopressin antagonists, in phase 2 studies these drugs have not yet yielded positive results in larger-scale trials. Interestingly, the compounds that show greater promise are those that are already in use for other indications and have been repurposed for anxiety, such as mifepristone, pregabalin, and DCS. There is also evidence for using prazosin in PTSD nightmares and possibly even for treating daytime symptoms. In contrast to the well-replicated data for the efficacy of ketamine in depression, the efficacy for PTSD is very preliminary; it remains to be seen whether the effects of ketamine can be replicated in a larger PTSD population.

Overall, the tendency to “look backward” at drug compounds that are already in use is only going to yield so much success. While prazosin, ketamine, D-cycloserine, and pregabalin may find a wide use for disorders like PTSD and GAD, there needs to be a stronger impetus to continue exploring other pathways, such as GABA, glutamate, neuropeptide, and serotonin systems, for newer pharmacotherapeutics. There are several promising agents under development, but recent history has not been encouraging. Given the high prevalence and morbidity of anxiety disorders among the general population and the high unmet need for effective treatments for these conditions, it is crucial that our field place a renewed emphasis on the discovery of novel, more effective pharmacological treatments.

Address correspondence to Dan V. Iosifescu, M.D., M.Sc., Director, Mood and Anxiety Disorders Program, Associate Professor of Psychiatry and Neuroscience, Icahn School of Medicine at Mount Sinai, One Gustave L. Levy Place, Box 1230, New York, NY 10029; e-mail:

Author Information and CME Disclosure

Amir Garakani, M.D., Department of Psychiatry, Icahn School of Medicine at Mount Sinai, New York, NY; Silver Hill Hospital, New Canaan, CT; Department of Psychiatry, Yale University School of Medicine, New Haven, CT

James W. Murrough, M.D., Department of Psychiatry and Department of Neuroscience, Icahn School of Medicine at Mount Sinai, New York, NY

Dan V. Iosifescu, M.D., M.Sc., Department of Psychiatry and Department of Neuroscience, Icahn School of Medicine at Mount Sinai, New York, NY

Dr. Garakani reports no financial relationships with commercial interests. In the past 2 years, Dr. Murrough has received research support from NIMH, the American Foundation for Suicide Prevention, the Doris Duke Charitable Foundation, Janssen Pharmaceuticals, and Avanir Pharmaceuticals; he has served on advisory boards for Genentech and Janssen Pharmaceuticals and has received consulting fees from ProPhase. Dr. Iosifescu has received research funding through Mount Sinai School of Medicine from NIMH, AstraZeneca, Brainsway, Euthymics, Neosync, and Roche; he has received consulting fees from Avanir, CNS Response, Otsuka, Lundbeck, Servier, and Sunovion.

References

1 Kessler RC, Chiu WT, Demler O, Merikangas KR, Walters EE: Prevalence, severity, and comorbidity of 12-month DSM-IV disorders in the National Comorbidity Survey Replication. Arch Gen Psychiatry 2005; 62:617–627CrossrefGoogle Scholar

2 Iancu SC, Batelaan NM, Zweekhorst MB, Bunders JF, Veltman DJ, Penninx BW, van Balkom AJ: Trajectories of functioning after remission from anxiety disorders: 2-year course and outcome predictors. Psychol Med 2014; 44:593–605CrossrefGoogle Scholar

3 Ahola K, Virtanen M, Honkonen T, Isometsä E, Aromaa A, Lönnqvist J: Common mental disorders and subsequent work disability: a population-based Health 2000 Study. J Affect Disord 2011; 134:365–372CrossrefGoogle Scholar

4 Smit F, Cuijpers P, Oostenbrink J, Batelaan N, de Graaf R, Beekman A: Costs of nine common mental disorders: implications for curative and preventive psychiatry. J Ment Health Policy Econ 2006; 9:193–200Google Scholar

5 Alonso J, Angermeyer MC, Bernert S, Bruffaerts R, Brugha TS, Bryson H, de Girolamo G, Graaf R, Demyttenaere K, Gasquet I, Haro JM, Katz SJ, Kessler RC, Kovess V, Lépine JP, Ormel J, Polidori G, Russo LJ, Vilagut G, Almansa J, Arbabzadeh-Bouchez S, Autonell J, Bernal M, Buist-Bouwman MA, Codony M, Domingo-Salvany A, Ferrer M, Joo SS, Martínez-Alonso M, Matschinger H, Mazzi F, Morgan Z, Morosini P, Palacín C, Romera B, Taub N, Vollebergh WA; ESEMeD/MHEDEA 2000 Investigators, European Study of the Epidemiology of Mental Disorders (ESEMeD) Project: Disability and quality of life impact of mental disorders in Europe: results from the European Study of the Epidemiology of Mental Disorders (ESEMeD) project. Acta Psychiatr Scand Suppl 2004; 109(420):38–46Google Scholar

6 American Psychiatric Association; Diagnostic and Statistical Manual of Mental Disorders, 4th ed. Washington, DC, American Psychiatric Publishing, 2000Google Scholar

7 American Psychiatric Association; Diagnostic and Statistical Manual of Mental Disorders, 5th ed. Arlington, VA, American Psychiatric Publishing, 2013CrossrefGoogle Scholar

8 Griebel G, Holmes A: 50 years of hurdles and hope in anxiolytic drug discovery. Nat Rev Drug Discov 2013; 12:667–687CrossrefGoogle Scholar

9 Celada P, Bortolozzi A, Artigas F: Serotonin 5-HT1A receptors as targets for agents to treat psychiatric disorders: rationale and current status of research. CNS Drugs 2013; 27:703–716CrossrefGoogle Scholar

10 Alpert JE, Franznick DA, Hollander SB, Fava M: Gepirone extended-release treatment of anxious depression: evidence from a retrospective subgroup analysis in patients with major depressive disorder. J Clin Psychiatry 2004; 65:1069–1075CrossrefGoogle Scholar

11 Rickels K, Mathew S, Banov MD, Zimbroff DL, Oshana S, Parsons EC, Donahue SR, Kauffman M, Iyer GR, Reinhard JF: Effects of PRX-00023, a novel, selective serotonin 1A receptor agonist on measures of anxiety and depression in generalized anxiety disorder: results of a double-blind, placebo-controlled trial. J Clin Psychopharmacol 2008; 28:235–239CrossrefGoogle Scholar

12 Fabre Kramer Pharmaceuticals: http://www.fabrekramer.com/TGFK08AA.html. Last Accessed April 9, 2014Google Scholar

13 Guilloux JP, Mendez-David I, Pehrson A, Guiard BP, Repérant C, Orvoën S, Gardier AM, Hen R, Ebert B, Miller S, Sanchez C, David DJ: Antidepressant and anxiolytic potential of the multimodal antidepressant vortioxetine (Lu AA21004) assessed by behavioural and neurogenesis outcomes in mice. Neuropharmacology 2013; 73:147–159CrossrefGoogle Scholar

14 Bidzan L, Mahableshwarkar AR, Jacobsen P, Yan M, Sheehan DV: Vortioxetine (Lu AA21004) in generalized anxiety disorder: results of an 8-week, multinational, randomized, double-blind, placebo-controlled clinical trial. Eur Neuropsychopharmacol 2012; 22:847–857CrossrefGoogle Scholar

15 Baldwin DS, Loft H, Florea I: Lu AA21004, a multimodal psychotropic agent, in the prevention of relapse in adult patients with generalized anxiety disorder. Int Clin Psychopharmacol 2012; 27:197–207CrossrefGoogle Scholar

16 Rothschild AJ, Mahableshwarkar AR, Jacobsen P, Yan M, Sheehan DV: Vortioxetine (Lu AA21004) 5 mg in generalized anxiety disorder: results of an 8-week randomized, double-blind, placebo-controlled clinical trial in the United States. Eur Neuropsychopharmacol 2012; 22:858–866CrossrefGoogle Scholar

17 Mahableshwarkar AR, Jacobsen PL, Serenko M, Chen Y: A randomized, double-blind, fixed-dose study comparing the efficacy and tolerability of vortioxetine 2.5 and 10 cmg in acute treatment of adults with generalized anxiety disorder. Hum Psychopharmacol 2014; 29:64–72CrossrefGoogle Scholar

18 Mahableshwarkar AR, Jacobsen PL, Chen Y, Simon JS: A randomised, double-blind, placebo-controlled, duloxetine-referenced study of the efficacy and tolerability of vortioxetine in the acute treatment of adults with generalised anxiety disorder. Int J Clin Pract 2014; 68:49–59CrossrefGoogle Scholar

19 Wesołowska A, Nikiforuk A, Stachowicz K: Anxiolytic-like and antidepressant-like effects produced by the selective 5-HT6 receptor antagonist SB-258585 after intrahippocampal administration to rats. Behav Pharmacol 2007; 18:439–446CrossrefGoogle Scholar

20 MacIsaac SE, Carvalho AF, Cha DS, Mansur RB, McIntyre RS: The mechanism, efficacy, and tolerability profile of agomelatine. Expert Opin Pharmacother 2014; 15:259–274CrossrefGoogle Scholar

21 Stein DJ, Ahokas AA, de Bodinat C: Efficacy of agomelatine in generalized anxiety disorder: a randomized, double-blind, placebo-controlled study. J Clin Psychopharmacol 2008; 28:561–566CrossrefGoogle Scholar

22 Stein DJ, Ahokas A, Albarran C, Olivier V, Allgulander C: Agomelatine prevents relapse in generalized anxiety disorder: a 6-month randomized, double-blind, placebo-controlled discontinuation study. J Clin Psychiatry 2012; 73:1002–1008CrossrefGoogle Scholar

23 Krystal JH, Mathew SJ, D’Souza DC, Garakani A, Gunduz-Bruce H, Charney DS: Potential psychiatric applications of metabotropic glutamate receptor glutamate receptor agonists and antagonists. CNS Drugs 2010; 24:669–693CrossrefGoogle Scholar

24 Bergink V, Westenberg HG: Metabotropic glutamate II receptor agonists in panic disorder: a double blind clinical trial with LY354740. Int Clin Psychopharmacol 2005; 20:291–293CrossrefGoogle Scholar

25 Dunayevich E, Erickson J, Levine L, Landbloom R, Schoepp DD, Tollefson GD: Efficacy and tolerability of an mGlu2/3 agonist in the treatment of generalized anxiety disorder. Neuropsychopharmacology 2008; 33:1603–1610CrossrefGoogle Scholar

26 Mathew SJ, Shah A, Lapidus K, Clark C, Jarun N, Ostermeyer B, Murrough JW: Ketamine for treatment-resistant unipolar depression: current evidence. CNS Drugs 2012; 26:189–204CrossrefGoogle Scholar

27 Murrough JW, Iosifescu DV, Chang LC, Al Jurdi RK, Green CE, Perez AM, Iqbal S, Pillemer S, Foulkes A, Shah A, Charney DS, Mathew SJ: Antidepressant efficacy of ketamine in treatment-resistant major depression: a two-site randomized controlled trial. Am J Psychiatry 2013; 170:1134–1142CrossrefGoogle Scholar

28 Zarate CA, Singh JB, Carlson PJ, Brutsche NE, Ameli R, Luckenbaugh DA, Charney DS, Manji HK: A randomized trial of an N-methyl-D-aspartate antagonist in treatment-resistant major depression. Arch Gen Psychiatry 2006; 63:856–864CrossrefGoogle Scholar

29 Zarate CA, Brutsche NE, Ibrahim L, Franco-Chaves J, Diazgranados N, Cravchik A, Selter J, Marquardt CA, Liberty V, Luckenbaugh DA: Replication of ketamine’s antidepressant efficacy in bipolar depression: a randomized controlled add-on trial. Biol Psychiatry 2012; 71:939–946CrossrefGoogle Scholar

30 Engin E, Treit D, Dickson CT: Anxiolytic- and antidepressant-like properties of ketamine in behavioral and neurophysiological animal models. Neuroscience 2009; 161:359–369CrossrefGoogle Scholar

31 Feder A, Parides MK, Murrough JW, Perez AM, Morgan JE, Saxena S, Kirkwood K, Aan Het Rot M, Wan LB, Iosifescu DV, Charney DS: Efficacy of intravenous ketamine in chronic posttraumatic stress disorder: a randomized controlled trial. JAMA Psychiatry (in press)Google Scholar

32 Irwin SA, Iglewicz A: Oral ketamine for the rapid treatment of depression and anxiety in patients receiving hospice care. J Palliat Med 2010; 13:903–908CrossrefGoogle Scholar

33 Irwin SA, Iglewicz A, Nelesen RA, Lo JY, Carr CH, Romero SD, Lloyd LS: Daily oral ketamine for the treatment of depression and anxiety in patients receiving hospice care: a 28-day open-label proof-of-concept trial. J Palliat Med 2013; 16:958–965CrossrefGoogle Scholar

34 Zarate CA, Payne JL, Quiroz J, Sporn J, Denicoff KK, Luckenbaugh D, Charney DS, Manji HK: An open-label trial of riluzole in patients with treatment-resistant major depression. Am J Psychiatry 2004; 161:171–174CrossrefGoogle Scholar

35 Sanacora G, Kendell SF, Levin Y, Simen AA, Fenton LR, Coric V, Krystal JH: Preliminary evidence of riluzole efficacy in antidepressant-treated patients with residual depressive symptoms. Biol Psychiatry 2007; 61:822–825CrossrefGoogle Scholar

36 Mathew SJ, Murrough JW, aan het Rot M, Collins KA, Reich DL, Charney DS: Riluzole for relapse prevention following intravenous ketamine in treatment-resistant depression: a pilot randomized, placebo-controlled continuation trial. Int J Neuropsychopharmacol 2010; 13:71–82CrossrefGoogle Scholar

37 Stutzmann JM, Cintrat P, Laduron PM, Blanchard JC: Riluzole antagonizes the anxiogenic properties of the beta-carboline FG 7142 in rats. Psychopharmacology (Berl) 1989; 99:515–519CrossrefGoogle Scholar

38 Sugiyama A, Saitoh A, Iwai T, Takahashi K, Yamada M, Sasaki-Hamada S, Oka J, Inagaki M, Yamada M: Riluzole produces distinct anxiolytic-like effects in rats without the adverse effects associated with benzodiazepines. Neuropharmacology 2012; 62:2489–2498CrossrefGoogle Scholar

39 Mathew SJ, Amiel JM, Coplan JD, Fitterling HA, Sackeim HA, Gorman JM: Open-label trial of riluzole in generalized anxiety disorder. Am J Psychiatry 2005; 162:2379–2381CrossrefGoogle Scholar

40 Mathew SJ, Price RB, Mao X, Smith EL, Coplan JD, Charney DS, Shungu DC: Hippocampal N-acetylaspartate concentration and response to riluzole in generalized anxiety disorder. Biol Psychiatry 2008; 63:891–898CrossrefGoogle Scholar

41 Abdallah CG, Coplan JD, Jackowski A, Sato JR, Mao X, Shungu DC, Mathew SJ: A pilot study of hippocampal volume and N-acetylaspartate (NAA) as response biomarkers in riluzole-treated patients with GAD. Eur Neuropsychopharmacol 2013; 23:276–284CrossrefGoogle Scholar

42 Myers KM, Carlezon WA, Davis M: Glutamate receptors in extinction and extinction-based therapies for psychiatric illness. Neuropsychopharmacology 2011; 36:274–293CrossrefGoogle Scholar

43 Nations KR, Dogterom P, Bursi R, Schipper J, Greenwald S, Zraket D, Gertsik L, Johnstone J, Lee A, Pande Y, Ruigt G, Ereshefsky L: Examination of Org 26576, an AMPA receptor positive allosteric modulator, in patients diagnosed with major depressive disorder: an exploratory, randomized, double-blind, placebo-controlled trial. J Psychopharmacol 2012; 26:1525–1539CrossrefGoogle Scholar

44 Davis M, Ressler K, Rothbaum BO, Richardson R: Effects of D-cycloserine on extinction: translation from preclinical to clinical work. Biol Psychiatry 2006; 60:369–375CrossrefGoogle Scholar

45 Heresco-Levy U, Kremer I, Javitt DC, Goichman R, Reshef A, Blanaru M, Cohen T: Pilot-controlled trial of D-cycloserine for the treatment of post-traumatic stress disorder. Int J Neuropsychopharmacol 2002; 5:301–307CrossrefGoogle Scholar

46 Bailey JE, Papadopoulos A, Lingford-Hughes A, Nutt DJ: D-Cycloserine and performance under different states of anxiety in healthy volunteers. Psychopharmacology (Berl) 2007; 193:579–585CrossrefGoogle Scholar

47 Onur OA, Schlaepfer TE, Kukolja J, Bauer A, Jeung H, Patin A, Otte DM, Shah NJ, Maier W, Kendrick KM, Fink GR, Hurlemann R: The N-methyl-D-aspartate receptor co-agonist D-cycloserine facilitates declarative learning and hippocampal activity in humans. Biol Psychiatry 2010; 67:1205–1211CrossrefGoogle Scholar

48 Nave AM, Tolin DF, Stevens MC: Exposure therapy, D-cycloserine, and functional magnetic resonance imaging in patients with snake phobia: a randomized pilot study. J Clin Psychiatry 2012; 73:1179–1186CrossrefGoogle Scholar

49 Otto MW, Tolin DF, Simon NM, Pearlson GD, Basden S, Meunier SA, Hofmann SG, Eisenmenger K, Krystal JH, Pollack MH: Efficacy of D-cycloserine for enhancing response to cognitive-behavior therapy for panic disorder. Biol Psychiatry 2010; 67:365–370CrossrefGoogle Scholar

50 Siegmund A, Golfels F, Finck C, Halisch A, Räth D, Plag J, Ströhle A: D-cycloserine does not improve but might slightly speed up the outcome of in-vivo exposure therapy in patients with severe agoraphobia and panic disorder in a randomized double blind clinical trial. J Psychiatr Res 2011; 45:1042–1047CrossrefGoogle Scholar

51 Hofmann SG, Meuret AE, Smits JA, Simon NM, Pollack MH, Eisenmenger K, Shiekh M, Otto MW: Augmentation of exposure therapy with D-cycloserine for social anxiety disorder. Arch Gen Psychiatry 2006; 63:298–304CrossrefGoogle Scholar

52 Guastella AJ, Richardson R, Lovibond PF, Rapee RM, Gaston JE, Mitchell P, Dadds MR: A randomized controlled trial of D-cycloserine enhancement of exposure therapy for social anxiety disorder. Biol Psychiatry 2008; 63:544–549CrossrefGoogle Scholar

53 Zalta AK, Dowd S, Rosenfield D, Smits JA, Otto MW, Simon NM, Meuret AE, Marques L, Hofmann SG, Pollack MH: Sleep quality predicts treatment outcome in CBT for social anxiety disorder. Depress Anxiety 2013; 30:1114–1120CrossrefGoogle Scholar

54 Smits JA, Hofmann SG, Rosenfield D, DeBoer LB, Costa PT, Simon NM, O’Cleirigh C, Meuret AE, Marques L, Otto MW, Pollack MH: D-cycloserine augmentation of cognitive behavioral group therapy of social anxiety disorder: prognostic and prescriptive variables. J Consult Clin Psychol 2013; 81:1100–1112CrossrefGoogle Scholar

55 Smits JA, Rosenfield D, Otto MW, Marques L, Davis ML, Meuret AE, Simon NM, Pollack MH, Hofmann SG: D-cycloserine enhancement of exposure therapy for social anxiety disorder depends on the success of exposure sessions. J Psychiatr Res 2013; 47:1455–1461CrossrefGoogle Scholar

56 Hofmann SG, Smits JA, Rosenfield D, Simon N, Otto MW, Meuret AE, Marques L, Fang A, Tart C, Pollack MH: D-Cycloserine as an augmentation strategy with cognitive-behavioral therapy for social anxiety disorder. Am J Psychiatry 2013; 170:751–758CrossrefGoogle Scholar

57 Ressler KJ, Rothbaum BO, Tannenbaum L, Anderson P, Graap K, Zimand E, Hodges L, Davis M: Cognitive enhancers as adjuncts to psychotherapy: use of D-cycloserine in phobic individuals to facilitate extinction of fear. Arch Gen Psychiatry 2004; 61:1136–1144CrossrefGoogle Scholar

58 Tart CD, Handelsman PR, Deboer LB, Rosenfield D, Pollack MH, Hofmann SG, Powers MB, Otto MW, Smits JA: Augmentation of exposure therapy with post-session administration of D-cycloserine. J Psychiatr Res 2013; 47:168–174CrossrefGoogle Scholar

59 Guastella AJ, Dadds MR, Lovibond PF, Mitchell P, Richardson R: A randomized controlled trial of the effect of D-cycloserine on exposure therapy for spider fear. J Psychiatr Res 2007; 41:466–471CrossrefGoogle Scholar

60 Gutner CA, Weinberger J, Hofmann SG: The effect of D-cycloserine on subliminal cue exposure in spider fearful individuals. Cogn Behav Ther 2012; 41:335–344CrossrefGoogle Scholar

61 Difede J, Cukor J, Wyka K, Olden M, Hoffman H, Lee FS, Altemus M: D-Cycloserine augmentation of exposure therapy for post-traumatic stress disorder: a pilot randomized clinical trial. Neuropsychopharmacology 2014; 39:1052–1058CrossrefGoogle Scholar

62 de Kleine RA, Hendriks GJ, Smits JA, Broekman TG, van Minnen A: Prescriptive variables for D-cycloserine augmentation of exposure therapy for posttraumatic stress disorder. J Psychiatr Res 2014; 48:40–46CrossrefGoogle Scholar

63 de Kleine RA, Hendriks GJ, Kusters WJ, Broekman TG, van Minnen A: A randomized placebo-controlled trial of D-cycloserine to enhance exposure therapy for posttraumatic stress disorder. Biol Psychiatry 2012; 71:962–968CrossrefGoogle Scholar

64 Litz BT, Salters-Pedneault K, Steenkamp MM, Hermos JA, Bryant RA, Otto MW, Hofmann SG: A randomized placebo-controlled trial of D-cycloserine and exposure therapy for posttraumatic stress disorder. J Psychiatr Res 2012; 46:1184–1190CrossrefGoogle Scholar

65 Sani G, Serra G, Kotzalidis GD, Romano S, Tamorri SM, Manfredi G, Caloro M, Telesforo CL, Caltagirone SS, Panaccione I, Simonetti A, Demontis F, Serra G, Girardi P: The role of memantine in the treatment of psychiatric disorders other than the dementias: a review of current preclinical and clinical evidence. CNS Drugs 2012; 26:663–690CrossrefGoogle Scholar

66 Feusner JD, Kerwin L, Saxena S, Bystritsky A: Differential efficacy of memantine for obsessive-compulsive disorder vs. generalized anxiety disorder: an open-label trial. Psychopharmacol Bull 2009; 42:81–93Google Scholar

67 Schwartz TL, Siddiqui UA, Raza S: Memantine as an augmentation therapy for anxiety disorders. Case Rep Psychiatry 2012; 2012:749–796Google Scholar

68 Battista MA, Hierholzer R, Khouzam HR, Barlow A, O’Toole S: Pilot trial of memantine in the treatment of posttraumatic stress disorder. Psychiatry 2007; 70:167–174CrossrefGoogle Scholar

69 Mula M, Pini S, Cassano GB: The role of anticonvulsant drugs in anxiety disorders: a critical review of the evidence. J Clin Psychopharmacol 2007; 27:263–272CrossrefGoogle Scholar

70 Boschen MJ: A meta-analysis of the efficacy of pregabalin in the treatment of generalized anxiety disorder. Can J Psychiatry 2011; 56:558–566CrossrefGoogle Scholar

71 Boschen MJ: Pregabalin: dose-response relationship in generalized anxiety disorder. Pharmacopsychiatry 2012; 45:51–56CrossrefGoogle Scholar

72 Greist JH, Liu-Dumaw M, Schweizer E, Feltner D: Efficacy of pregabalin in preventing relapse in patients with generalized social anxiety disorder: results of a double-blind, placebo-controlled 26-week study. Int Clin Psychopharmacol 2011; 26:243–251CrossrefGoogle Scholar

73 Feltner DE, Liu-Dumaw M, Schweizer E, Bielski R: Efficacy of pregabalin in generalized social anxiety disorder: results of a double-blind, placebo-controlled, fixed-dose study. Int Clin Psychopharmacol 2011; 26:213–220CrossrefGoogle Scholar

74 Pande AC, Davidson JR, Jefferson JW, Janney CA, Katzelnick DJ, Weisler RH, Greist JH, Sutherland SM: Treatment of social phobia with gabapentin: a placebo-controlled study. J Clin Psychopharmacol 1999; 19:341–348CrossrefGoogle Scholar

75 Pande AC, Pollack MH, Crockatt J, Greiner M, Chouinard G, Lydiard RB, Taylor CB, Dager SR, Shiovitz T: Placebo-controlled study of gabapentin treatment of panic disorder. J Clin Psychopharmacol 2000; 20:467–471CrossrefGoogle Scholar

76 Schmitt U, Lüddens H, Hiemke C: Anxiolytic-like effects of acute and chronic GABA transporter inhibition in rats. J Neural Transm 2002; 109:871–880CrossrefGoogle Scholar

77 Pollack MH, Tiller J, Xie F, Trivedi MH: Tiagabine in adult patients with generalized anxiety disorder: results from 3 randomized, double-blind, placebo-controlled, parallel-group studies. J Clin Psychopharmacol 2008; 28:308–316CrossrefGoogle Scholar

78 Zwanzger P, Eser D, Nothdurfter C, Baghai TC, Möller HJ, Padberg F, Rupprecht R: Effects of the GABA-reuptake inhibitor tiagabine on panic and anxiety in patients with panic disorder. Pharmacopsychiatry 2009; 42:266–269CrossrefGoogle Scholar

79 Urbano MR, Spiegel DR, Laguerta N, Shrader CJ, Rowe DF, Hategan LF: Gabapentin and tiagabine for social anxiety: a randomized, double-blind, crossover study of 8 adults. Prim Care Companion J Clin Psychiatry 2009; 11:123CrossrefGoogle Scholar

80 Mirza NR, Bright JL, Stanhope KJ, Wyatt A, Harrington NR: Lamotrigine has an anxiolytic-like profile in the rat conditioned emotional response test of anxiety: a potential role for sodium channels? Psychopharmacology (Berl) 2005; 180:159–168CrossrefGoogle Scholar

81 Masdrakis VG, Papadimitriou GN, Oulis P: Lamotrigine administration in panic disorder with agoraphobia. Clin Neuropharmacol 2010; 33:126–128CrossrefGoogle Scholar

82 Hertzberg MA, Butterfield MI, Feldman ME, Beckham JC, Sutherland SM, Connor KM, Davidson JR: A preliminary study of lamotrigine for the treatment of posttraumatic stress disorder. Biol Psychiatry 1999; 45:1226–1229CrossrefGoogle Scholar

83 Di Simplicio M, Massey-Chase R, Cowen PJ, Harmer CJ: Oxytocin enhances processing of positive versus negative emotional information in healthy male volunteers. J Psychopharmacol 2009; 23:241–248CrossrefGoogle Scholar

84 Alvares GA, Chen NT, Balleine BW, Hickie IB, Guastella AJ: Oxytocin selectively moderates negative cognitive appraisals in high trait anxious males. Psychoneuroendocrinology 2012; 37:2022–2031CrossrefGoogle Scholar

85 Labuschagne I, Phan KL, Wood A, Angstadt M, Chua P, Heinrichs M, Stout JC, Nathan PJ: Oxytocin attenuates amygdala reactivity to fear in generalized social anxiety disorder. Neuropsychopharmacology 2010; 35:2403–2413CrossrefGoogle Scholar

86 Guastella AJ, Howard AL, Dadds MR, Mitchell P, Carson DS: A randomized controlled trial of intranasal oxytocin as an adjunct to exposure therapy for social anxiety disorder. Psychoneuroendocrinology 2009; 34:917–923CrossrefGoogle Scholar

87 Tauscher J, Kielbasa W, Iyengar S, Vandenhende F, Peng X, Mozley D, Gehlert DR, Marek G: Development of the 2nd generation neurokinin-1 receptor antagonist LY686017 for social anxiety disorder. Eur Neuropsychopharmacol 2010; 20:80–87CrossrefGoogle Scholar

88 Furmark T, Appel L, Michelgård A, Wahlstedt K, Ahs F, Zancan S, Jacobsson E, Flyckt K, Grohp M, Bergström M, Pich EM, Nilsson LG, Bani M, Långström B, Fredrikson M: Cerebral blood flow changes after treatment of social phobia with the neurokinin-1 antagonist GR205171, citalopram, or placebo. Biol Psychiatry 2005; 58:132–142CrossrefGoogle Scholar

89 Mathew SJ, Vythilingam M, Murrough JW, Zarate CA, Feder A, Luckenbaugh DA, Kinkead B, Parides MK, Trist DG, Bani MS, Bettica PU, Ratti EM, Charney DS: A selective neurokinin-1 receptor antagonist in chronic PTSD: a randomized, double-blind, placebo-controlled, proof-of-concept trial. Eur Neuropsychopharmacol 2011; 21:221–229CrossrefGoogle Scholar

90 Michelson D, Hargreaves R, Alexander R, Ceesay P, Hietala J, Lines C, Reines S: Lack of efficacy of L-759274, a novel neurokinin 1 (substance P) receptor antagonist, for the treatment of generalized anxiety disorder. Int J Neuropsychopharmacol 2013; 16:1–11CrossrefGoogle Scholar

91 Serova LI, Tillinger A, Alaluf LG, Laukova M, Keegan K, Sabban EL: Single intranasal neuropeptide Y infusion attenuates development of PTSD-like symptoms to traumatic stress in rats. Neuroscience 2013; 236:298–312CrossrefGoogle Scholar

92 Sah R, Geracioti TD: Neuropeptide Y and posttraumatic stress disorder. Mol Psychiatry 2013; 18:646–655CrossrefGoogle Scholar

93 Lacroix JS, Ricchetti AP, Morel D, Mossimann B, Waeber B, Grouzmann E: Intranasal administration of neuropeptide Y in man: systemic absorption and functional effects. Br J Pharmacol 1996; 118:2079–2084CrossrefGoogle Scholar

94 Hallschmid M, Benedict C, Born J, Fehm HL, Kern W: Manipulating central nervous mechanisms of food intake and body weight regulation by intranasal administration of neuropeptides in man. Physiol Behav 2004; 83:55–64CrossrefGoogle Scholar

95 Shimazaki T, Iijima M, Chaki S: The pituitary mediates the anxiolytic-like effects of the vasopressin V1B receptor antagonist, SSR149415, in a social interaction test in rats. Eur J Pharmacol 2006; 543:63–67CrossrefGoogle Scholar

96 Bleickardt CJ, Mullins DE, Macsweeney CP, Werner BJ, Pond AJ, Guzzi MF, Martin FD, Varty GB, Hodgson RA: Characterization of the V1a antagonist, JNJ-17308616, in rodent models of anxiety-like behavior. Psychopharmacology (Berl) 2009; 202:711–718CrossrefGoogle Scholar

97 Griebel G, Beeské S, Stahl SM: The vasopressin V(1b) receptor antagonist SSR149415 in the treatment of major depressive and generalized anxiety disorders: results from 4 randomized, double-blind, placebo-controlled studies. J Clin Psychiatry 2012; 73:1403–1411CrossrefGoogle Scholar

98 Lee RJ, Coccaro EF, Cremers H, McCarron R, Lu SF, Brownstein MJ, Simon NG: A novel V1a receptor antagonist blocks vasopressin-induced changes in the CNS response to emotional stimuli: an fMRI study. Front Syst Neurosci 2013; 7:100CrossrefGoogle Scholar

99 Bailey JE, Papadopoulos A, Diaper A, Phillips S, Schmidt M, van der Ark P, Dourish CT, Dawson GR, Nutt DJ: Preliminary evidence of anxiolytic effects of the CRF(1) receptor antagonist R317573 in the 7.5% CO(2) proof-of-concept experimental model of human anxiety. J Psychopharmacol 2011; 25:1199–1206CrossrefGoogle Scholar

100 Risbrough VB, Stein MB: Role of corticotropin releasing factor in anxiety disorders: a translational research perspective. Horm Behav 2006; 50:550–561CrossrefGoogle Scholar

101 Coric V, Feldman HH, Oren DA, Shekhar A, Pultz J, Dockens RC, Wu X, Gentile KA, Huang SP, Emison E, Delmonte T, D’Souza BB, Zimbroff DL, Grebb JA, Goddard AW, Stock EG: Multicenter, randomized, double-blind, active comparator and placebo-controlled trial of a corticotropin-releasing factor receptor-1 antagonist in generalized anxiety disorder. Depress Anxiety 2010; 27:417–425CrossrefGoogle Scholar

102 Pitman RK, Milad MR, Igoe SA, Vangel MG, Orr SP, Tsareva A, Gamache K, Nader K: Systemic mifepristone blocks reconsolidation of cue-conditioned fear; propranolol prevents this effect. Behav Neurosci 2011; 125:632–638CrossrefGoogle Scholar

103 Golier JA, Caramanica K, Demaria R, Yehuda R: A pilot study of mifepristone in combat-related PTSD. Depress Res Treat 2012; 2012:393251Google Scholar

104 Neylan TC, Lenoci M, Samuelson KW, Metzler TJ, Henn-Haase C, Hierholzer RW, Lindley SE, Otte C, Schoenfeld FB, Yesavage JA, Marmar CR: No improvement of posttraumatic stress disorder symptoms with guanfacine treatment. Am J Psychiatry 2006; 163:2186–2188CrossrefGoogle Scholar

105 Davis LL, Ward C, Rasmusson A, Newell JM, Frazier E, Southwick SM: A placebo-controlled trial of guanfacine for the treatment of posttraumatic stress disorder in veterans. Psychopharmacol Bull 2008; 41:8–18Google Scholar

106 Connor DF, Grasso DJ, Slivinsky MD, Pearson GS, Banga A: An open-label study of guanfacine extended release for traumatic stress related symptoms in children and adolescents. J Child Adolesc Psychopharmacol 2013; 23:244–251CrossrefGoogle Scholar

107 Pitman RK, Sanders KM, Zusman RM, Healy AR, Cheema F, Lasko NB, Cahill L, Orr SP: Pilot study of secondary prevention of posttraumatic stress disorder with propranolol. Biol Psychiatry 2002; 51:189–192CrossrefGoogle Scholar

108 Stein MB, Kerridge C, Dimsdale JE, Hoyt DB: Pharmacotherapy to prevent PTSD: Results from a randomized controlled proof-of-concept trial in physically injured patients. J Trauma Stress 2007; 20:923–932CrossrefGoogle Scholar

109 Henry M, Fishman JR, Youngner SJ: Propranolol and the prevention of post-traumatic stress disorder: is it wrong to erase the “sting” of bad memories? Am J Bioeth 2007; 7:12–20CrossrefGoogle Scholar

110 de Kleine RA, Rothbaum BO, van Minnen A: Pharmacological enhancement of exposure-based treatment in PTSD: a qualitative review Eur J Psychotraumatol, 2013, p 4Google Scholar

111 Pittler MH, Ernst E: Kava extract for treating anxiety. Cochrane Database Syst Rev 2003; (1):CD003383Google Scholar

112 Sarris J, LaPorte E, Schweitzer I: Kava: a comprehensive review of efficacy, safety, and psychopharmacology. Aust N Z J Psychiatry 2011; 45:27–35CrossrefGoogle Scholar