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

Self-Assessment Activity: Understanding the Evidence for Off-Label Use of Atypical Antipsychotic Medications

In 2011, the Agency for Healthcare Quality and Research (AHRQ) published a comprehensive report (“Off-Label Use of Atypical Antipsychotics: An Update,” Comparative Effectiveness Review No. 43 [1]) summarizing the evidence for off-label use of antipsychotics for treatment of mental disorders (available at www.effectivehealthcare.ahrq.gov/reports/final.cfm). As described in the report, prescribing of atypical antipsychotics has moved beyond approved indications; however, the effectiveness, benefits, and adverse effects in off-label uses are not well understood. Through a grant from AHRQ, the American Psychiatric Association (APA) developed a FREE multipart CME program to disseminate the conclusions of the AHRQ report and to update clinicians about new developments regarding the evidence.

Our understanding of the evidence continues to evolve. The developers of this program are aware that off-label uses of antipsychotic medications constitute rapidly changing areas in clinical investigation and that new studies and revised Food and Drug Administration (FDA) approvals appear on a regular basis. The material in this program represents the best information available to the developers based on AHRQ reviews and updated literature as of February 2014.

Overall, a class effect of the atypical antipsychotics for each disorder cannot be assumed, and for most atypicals, adequate supporting evidence for either efficacy or comparative effectiveness is still lacking for many indications. The study of the evidence provided in this program will assist the physician in making informed decisions and provide a foundation for discussions with patients about the risks and benefits of this class of medications.

The AHRQ report addressed the following questions.

•. 

What are the leading off-label uses of atypical antipsychotics in utilization studies? How have trends in utilization changed in recent years, including inpatient versus outpatient use? What new uses are being studied in trials?

•. 

What does the evidence show (and how strong is that evidence) regarding the efficacy and comparative effectiveness of atypical antipsychotics for off-label indications?

•. 

How do atypical antipsychotic medications compare with other drugs, including first-generation antipsychotics, for off-label indications?

•. 

What are the potential adverse effects and/or complications involved with off-label prescribing of atypical antipsychotics? How do they compare within the class and with other drugs used for the conditions?

The task of the AHRQ reviewers was to review off-label use. It is important to note that other established agents as well as psychosocial treatments are available as first line treatments. Table 1 shows how the strength of evidence assessed in the AHRQ report was classified.

Table 1. Strength of Evidence Classificationsa
Strength of EvidenceCriteria
High
High confidence that the evidence reflects the true effect. Further research is very unlikely to change the confidence in the estimate of effect.
Moderate
Moderate confidence that the evidence reflects the true effect. Further research may change the confidence in the estimate of effect and may change the estimate.
Low
Low confidence that the evidence reflects the true effect. Further research is likely to change the confidence in the estimate of effect and is likely to change the estimate.
Insufficient
Evidence either is unavailable or does not permit estimation of an effect.

a From “Off-Label Use of Atypical Antipsychotics: An Update,” Comparative Effectiveness Review No. 43. September 2011, Agency for Healthcare Research and Quality, p 12.

Table 1. Strength of Evidence Classificationsa
Enlarge table

Strength of evidence is not the same as the size of the potential effect (i.e., There may be a moderate to high strength of evidence of a beneficial effect, but the size of the effect may be small).

Self-Assessment Activity

Part one of the evidence dissemination program is the self-assessment activity presented as a special feature in this issue of FOCUS. Participants complete a 25-question, multiple-choice self-assessment and assess their strengths and weaknesses regarding their knowledge of evidence for off-label use of atypical antipsychotics. After completing the quiz online at apaeducation.org, participants receive a score report with peer comparison, review the answers and rationales for the quiz, and access links to the AHRQ report and to other important clinical references that describe the evidence for the risks and benefits of these medications.

Look for part 2 of this activity available both at apaeducation.org and also delivered by e-mail as eFocus clinical vignette exercises. After reading a vignette, answer online questions regarding patient management decisions and read the commentary for the best answer choice. You will be able to compare your treatment decisions to those of your peers and receive additional self-assessment CME credit approved by ABPN for MOC part 2. Through completion of this self-assessment and the 10 vignettes that follow, participants will be able to earn up to 25 self-assessment credits for this multipart program.

Self-Assessment

Go to apaeducation.org to complete the self-assessment.

1. Which of the following atypical antipsychotics has moderate strength evidence of efficacy for off-label use as a monotherapy (single-agent) treatment for major depressive disorder (MDD)?

a. 

Aripirazole.

b. 

Olanzapine.

c. 

Quetiapine.

d. 

Risperidone.

e. 

Ziprasidone.

2. What is the conclusion of the AHRQ report about the efficacy of antipsychotic medications as augmenting agents of antidepressants for MDD?

a. 

All atypical antipsychotics have a strong level of evidence of efficacy as augmentation agents.

b. 

Evidence showed atypical antipsychotics were no better than placebo.

c. 

Aripiprazole, quetiapine, and risperidone have efficacy with moderate or high evidence.

d. 

Head to head trials have shown which agent is most efficacious.

e. 

Ziprasidone and olanzapine have moderate or high evidence of efficacy.

3. When using an atypical antipsychotic agent to augment an antidepressant medication in nonpsychotic unipolar depression, which of the following measures is most likely to be useful for managing side effect(s)?

a. 

Monitoring weight.

b. 

Monitoring serum potassium (K+).

c. 

Monitoring urinary 3-Methoxy-4-hydroxyphenylglycol (MHPG).

d. 

Monitoring finger-to-nose coordination.

4. For treatment of patients with PTSD without psychosis, what does the most recent evidence reflected in the Department of Defense (DoD) guideline currently recommend regarding the use of atypical antipsychotics in the treatment of these patients?

a. 

Recommended as monotherapy.

b. 

Recommended as augmentation.

c. 

Not recommended.

5. A 28-year-old female veteran is receiving cognitive-processing therapy for intrusive memories of dead bodies she saw while serving in Afghanistan. When she has these memories, she feels as if she is there, and feels extremely distressed. In addition, she finds herself avoiding public places, and startling easily when approached by her family. She describes herself as “angry” all the time; however, she does feel her sleep, appetite, and energy are normal. After several months of therapy she is less irritable but continues to have the intrusive visions. Which of the following medications has the best evidence for efficacy in treating her symptoms?

a. 

Quetiapine.

b. 

Sertraline.

c. 

Prazosin.

d. 

Clonazepam.

e. 

Risperidone.

6. According to the 2011 AHRQ review, which atypical has the highest level of confidence for effectiveness to support its use in the management of dementia-related agitation?

a. 

Aripiprazole.

b. 

Quetiapine.

c. 

Risperidone.

d. 

Ziprasidone.

7. A psychiatry consultation is requested for a 72-year-old nursing home resident with major neurocognitive disorder because of Alzheimer’s disease. In the past, the patient had been violent, injuring both himself and others. Behavioral interventions were tried without success. One year prior, the patient was put on risperidone (1 mg) and showed improvement. An attempt to lower his dose 6 months later resulted in worsening behaviors, and the patient was returned to 1 mg. Since then, there have been no incidents. He also receives hydrochlorothiazide and atenolol for hypertension. His gait is somewhat slow but the nurses insist that this was the case before he was started on risperidone.

A recent pharmacy review questioned the use of antipsychotics for a nonpsychotic patient and informed the Director of Nursing that these drugs are associated with increased mortality in patients with dementia. As a result, the Director of Nursing is requesting an alternative drug. The patient’s family, however, expresses concern over the possibility that the patient’s agitation may return. Which of the following would be the most appropriate recommendation for a psychiatric consultant to give?

a. 

Continue the risperidone, titrating to lowest effective dose.

b. 

Continue the risperidone but add adjunctive benztropine.

c. 

Discontinue risperidone and use behavioral interventions instead.

d. 

Discontinue risperidone and administer fluoxetine.

e. 

Switch from risperidone to aripiprazole.

8. Which of the following atypical antipsychotics is most likely to cause weight gain in elderly patients?

a. 

Aripiprazole.

b. 

Olanzapine.

c. 

Quetiapine.

d. 

Risperidone.

e. 

Ziprasidone.

9. In nonelderly adults (18–64) taking atypical antipsychotic medications, what does the evidence show about antipsychotics and weight gain in this patient group?

a. 

All antipsychotics are associated with equal weight gain.

b. 

Weight gain is not associated with off label use.

c. 

Olanzapine is associated with the most weight gain.

d. 

Quetiapine is associated with the most weight gain.

e. 

Risperidone is associated with the most weight gain.

10. A 46-year-old woman sees a psychiatrist for ”chronic anxiety.” She has always been an anxious person but has noticed more symptoms in the past 6 months when she returned to the workforce after being home with her children for 10 years. She has difficulty concentrating, and feels on edge at work. She also has difficulty falling asleep and feels fatigued during the day. She says she is not depressed, just “worried all the time.”

She denies any substance use or medical problems. Her physical last year was unremarkable. Which of the following medications would be a first choice to help this patient?

a. 

Alprazolam.

b. 

Diazepam.

c. 

Paroxetine.

d. 

Quetiapine.

e. 

Risperidone.

11. In the 2011 AHRQ review, which of the following studied atypical medications has the most evidence for effective treatment of moderate generalized anxiety disorder?

a. 

Aripiprazole

b. 

Olanzapine

c. 

Quetiapine

d. 

Risperidone

e. 

Ziprasidone

12. What is the best estimate of risk of tardive dyskinesia with atypical antipsychotics compared with typical antipsychotics?

a. 

No risk.

b. 

The same risk as typical antipsychotics.

c. 

10% of the risk of typical antipsychotics.

d. 

Risk estimates vary between 25% and 55% of the risk of typical antipsychotics

13. A 20-year-old college student is referred to a psychiatrist by the student health service for possible eating disorder. She has lost 30 pounds and now weighs 95 pounds at 5′4”. She says she feels fat and would like to lose more weight. She does not binge or purge, although in high school she once tried using laxatives to lose weight. Her electrolytes and lab tests are all WNL.

What is the next step the psychiatrist should take?

a. 

Arrange a family meeting.

b. 

Observe her eating a meal.

c. 

Begin fluoxetine.

d. 

Begin olanzapine.

e. 

Begin psychosocial interventions.

14. For the treatment of patients with eating disorders, what does the evidence show regarding the use of atypical antipsychotics to increase body mass index (BMI) in patients with anorexia nervosa (AN)?

a. 

Evidence for BMI increase with olanzapine in AN patients.

b. 

Evidence for BMI increase with risperidone in AN patients.

c. 

Evidence for BMI increase with all atypical antipsychotics in AN patients.

d. 

Evidence is inconclusive for increase in BMI in AN patients.

e. 

Evidence for BMI increase with quetiapine in AN patients

15. What do the data show about efficacy of atypical antipsychotics in the treatment of substance use disorders?

a. 

There is low strength evidence of efficacy.

b. 

There is moderate strength evidence of efficacy.

c. 

There is strong evidence of efficacy.

d. 

There have been no studies.

e. 

There is evidence of low to moderate strength for inefficacy.

16. A 35-year-old stockbroker is referred to you for treatment of methamphetamine addiction. Which of the following is the most effective treatment for this addiction?

a. 

Aripiprazole.

b. 

Exposure and response prevention.

c. 

Buprenorphine.

d. 

Cognitive behavior therapy.

e. 

Naloxone.

17. A 45-year-old woman complains that she has difficulty falling asleep almost every night. This is a long standing problem but has worsened over the past several months as her job has become more stressful. She watches TV in bed and has been drinking a glass of wine before bed, which seemed to help, but she then began waking up at 2 a.m. and is not able to get back to sleep. What would be the best treatment at this time?

a. 

Quetiapine.

b. 

CBT and sleep hygiene education.

c. 

Zolpidem.

d. 

Diazepam.

e. 

Trazodone.

18. Treatment of adolescents and children with atypical antipsychotic medications, including olanzapine, risperidone, quetiapine, and aripiprazole, for psychotic and bipolar disorders was associated with:

a. 

An increased risk for obesity.

b. 

An increased risk for metabolic adverse events (cardiovascular, cerebrovascular, or hypertensive adverse event).

c. 

No increased risk for obesity.

d. 

No increased risk for metabolic abnormalities.

e. 

a and b

f. 

c & d

19. The AHRQ report reviewed the evidence for atypical antipsychotics in the treatment of Borderline Personality Disorder (BPD). What does the evidence show regarding treatment of Borderline patients with atypical antipsychotics? For which symptoms of BPD do atypical antipsychotics have the best evidence?

a. 

Affective instability.

b. 

Core disturbances in manifestations of self.

c. 

Impulsivity and aggression.

d. 

Interpersonal dysfunction.

20. Which of the following medications would be the first choice for a 10-year-old boy with ADHD and no serious co-occurring disorders?

a. 

Alprazolam.

b. 

Gabapentin.

c. 

Methylphenidate

d. 

Paroxetine.

e. 

Risperidone.

21. The AHRQ review reported on the effectiveness of atypical antipsychotics as medication augmentation in obsessive compulsive disorder (OCD) as studied in two different experimental designs: one in which specific atypical antipsychotics augment any SRI antidepressant and the other in which they augment citalopram. Some atypical antipsychotics have shown efficacy. When augmenting SRIs, which atypical has the most evidence for effectiveness?

a. 

Aripiprazole.

b. 

Olanzapine.

c. 

Quetiapine.

d. 

Risperidone.

e. 

Ziprasidone.

22. According to the 2011 AHRQ review, for elders (≥65 years old) receiving atypical antipsychotic agents for the management of dementia, stroke (CVA) was significantly higher in medication than placebo subjects for which medication?

a. 

Aripiprazole.

b. 

Olanzapine.

c. 

Quetiapine.

d. 

Risperidone.

e. 

Ziprasidone.

24. In adults treated with atypical antipsychotics, which of the following atypical antipsychotics has shown the lowest risk of extrapyramidal symptoms?

a. 

Aripiprazole.

b. 

Olanzapine.

c. 

Quetiapine.

d. 

Risperidone.

e. 

Ziprasidone.

25. Which of the following is an approved indication for treatment with an atypical for a use other than for psychosis?

a. 

Behavioral symptoms of dementia.

b. 

Monotherapy for major depressive disorder.

c. 

Augmentation of selective serotonin reuptake inhibitors (SSRIs) in OCD.

d. 

Augmentation of SSRIs in major depressive disorder.

CME Information and Program Planner and Consultant Disclosures

Earn up to 5 CME Self-Assessment credits (MOC Part 2)through completion of this special activity online at apaeducation.org

Supported by a grant from the Agency for Healthcare Research and Quality (R18 HS021944)

Deborah J. Hales, M.D., Director, American Psychiatric Association Division of Education, Arlington, VA

Eve K. Mościcki, Sc.D., M.P.H., Director, Practice Research Network American Psychiatric Foundation, Arlington, VA

Robert Boland, M.D., Professor of Psychiatry and Human Behavior, Brown Alpert Medical School

David Fogelson, M.D., Clinical Professor of Psychiatry, UCLA Department of Psychiatry and Bio-behavioral Sciences

Research Support: Genentech.

Ian A. Cook, M.D., Professor of Psychiatry, David Geffen School of Medicine and Semel Institute for Neuroscience and Human Behavior at UCLA

Disclosures past 5 years: Grant Support: Aspect Medical Systems/Covidien, Cyberonics, Eli Lilly, MedAvante, Merck, Neuronetics, Novartis, Pfizer, Seaside Therapeutics, Sepracor, Shire. Adviser/Consultant: Allergan, Ascent Media, Bristol-Myers Squibb, Cyberonics, Eli Lilly, Forest, Janssen, Neuronetics, NeuroSigma, NIH (ITVA), Pfizer, Scale Venture Partners, VA (DSMB) ; Speaker: Bristol-Myers Squibb, CME LLC, Eli Lilly, Medical Education Speakers Network, Neuronetics, NeuroSigma, Pfizer, Wyeth. Biomedical Device Patents assigned to the University of California. Stock Options: NeuroSigma.

Joel Yager, M.D. Professor of Psychiatry, University of Colorado School of Medicine.

Drs. Hales, Moscicki, Boland and Yager report no competing interests.

The American Psychiatric Association (APA) is accredited by the Accreditation Council for Continuing Medical Education (ACCME) to provide continuing medical education for physicians. The APA designates this enduring material for a maximum of five AMA PRA Category 1 Credits. Physicians should claim credit commensurate with the extent of their participation in the activity.

The American Board of Psychiatry and Neurology has approved this program as part of a comprehensive self assessment program, which is mandated by the ABMS as a necessary component of Maintenance of Certification.

The American Board of Psychiatry and Neurology has reviewed the American Psychiatric Association Clinical eFOCUS Program in psychiatry - of which this is a part - and has approved this program as part of a comprehensive self-assessment program, which is mandated by the ABMS as a necessary component of Maintenance of Certification.

Bibliography

Maglione M, Ruelaz Maher A, Hu J, Wang Z, Shanman R, Shekelle PG, Roth B, Hilton L, Suttorp MJ, Ewing BA, Motala A, Perry T. Off-Label Use of Atypical Antipsychotics: An Update. Comparative Effectiveness Review No. 43. (Prepared by the Southern California Evidence-based Practice Center under Contract No. HHSA290-2007-10062-1.) Rockville, MD: Agency for Healthcare Research and Quality. September 2011. Available at: www.effectivehealthcare.ahrq.gov/reports/final.cfm.Google Scholar

AHRQ Comparative Effectiveness Review Number 43. Off-Label Use of Atypical Antipsychotics: An Update. September 2011. Executive SummaryGoogle Scholar

Off-Label Use of Atypical Antipsychotics: An Update. Summary Guide for Clinicians. John M. Eisenberg Center for Clinical Decisions and Communications Science. Comparative Effectiveness Review Summary Guides for Clinicians [Internet]. Rockville (MD): Agency for Healthcare Research and Quality (US); 2012 Aug 01. Available at: http://www.ncbi.nlm.nih.gov/books/NBK142851/Google Scholar

AHRQ Comparative Effectiveness Review Number 43. Off-Label Use of Atypical Antipsychotics: An Update. September 2011. Table A. AHRQ Report: Summary of strength of evidence of efficacy, by drug and condition p ES4. Table B: Summary Update: Efficacy of Atypical Antipsychotics for Off label use), ES 8. Table C. Summary update: safety of atypical antipsychotics for off-label use ES-11-13Google Scholar

Papakostas GI, Vitolo OV, Ishak WW, Rapaport MH, Zajecka JM, Kinrys G, Mischoulon D, Lipkin SH, Hails KA, Abrams J, Ward SG, Meisner A, Schoenfeld DA, Shelton RC, Winokur A, Okasha MS, Bari MA, Fava M: A 12-week, randomized, double-blind, placebo-controlled, sequential parallel comparison trial of ziprasidone as monotherapy for major depressive disorder. J Clin Psychiatry 2012; 73:1541–1547CrossrefGoogle Scholar

Maneeton N, Maneeton B, Srisurapanont M, Martin SD: Quetiapine monotherapy in acute phase for major depressive disorder: a meta-analysis of randomized, placebo-controlled trials. BMC Psychiatry 2012; 12:160CrossrefGoogle Scholar

Nelson JC, Papakostas GI: Atypical antipsychotic augmentation in major depressive disorder: a meta-analysis of placebo-controlled randomized trials. Am J Psychiatry 2009; 166:980–991CrossrefGoogle Scholar

Papakostas GI, Shelton RC, Smith J, Fava M: Augmentation of antidepressants with atypical antipsychotic medications for treatment-resistant major depressive disorder: a meta-analysis. J Clin Psychiatry 2007; 68:826–831CrossrefGoogle Scholar

Kohen I, Voelker S, Manu P: Antipsychotic-induced hyponatremia: case report and literature review. Am J Ther 2008; 15:492–494CrossrefGoogle Scholar

Garvey M, Hollon SD, DeRubeis RJ, Evans MD, Tuason VB: Does 24-h urinary MHPG predict treatment response to antidepressants? I. A review. J Affect Disord 1990; 20:173–179CrossrefGoogle Scholar

Friedman MJ: PTSD: Pharmacotherapeutic Approaches. Focus 2013; 11:315–320LinkGoogle Scholar

Krystal JH, Rosenheck RA, Cramer JA, Vessicchio JC, Jones KM, Vertrees JE, Horney RA, Huang GD, Stock C; Veterans Affairs Cooperative Study No. 504 Group: Adjunctive risperidone treatment for antidepressant-resistant symptoms of chronic military service-related PTSD: a randomized trial. JAMA 2011; 306:493–502CrossrefGoogle Scholar

Department of Veterans Affairs, Department of Defense: Clinical Practice Guideline For Management Of Post-Traumatic Stress, version 2.0. Oct 2010. Available at: http://www.healthquality.va.gov/post_traumatic_stress_disorder_ptsd.aspGoogle Scholar

Hoge CW: Interventions for war-related posttraumatic stress disorder: meeting veterans where they are. JAMA 2011; 306:549–551CrossrefGoogle Scholar

Choosing Wisely http://www.choosingwisely.org/doctor-patient-lists/american-psychiatric-association/accessed March 2014Google Scholar

Herrmann N, Lanctôt KL, Hogan DB: Pharmacological recommendations for the symptomatic treatment of dementia: the Canadian Consensus Conference on the Diagnosis and Treatment of Dementia 2012. Alzheimers Res Ther 2013; 5(Suppl 1):S5CrossrefGoogle Scholar

Lopez OL, Becker JT, Chang Y-F, Sweet RA, Aizenstein H, Snitz B, Saxton J, McDade E, Kamboh MI, DeKosky ST, Reynolds CF, Klunk WE: The long-term effects of conventional and atypical antipsychotics in patients with probable Alzheimer’s disease. Am J Psychiatry 2013; 170:1051–1058CrossrefGoogle Scholar

Keenmon C, Sultzer D: The role of antipsychotic drugs in the treatment of neuropsychiatric symptoms of dementia. Focus 2013; 11:32–38LinkGoogle Scholar

Rummel-Kluge C, Komossa K, Schwarz S, Hunger H, Schmid F, Lobos CA, Kissling W, Davis JM, Leucht S: Head-to-head comparisons of metabolic side effects of second generation antipsychotics in the treatment of schizophrenia: a systematic review and meta-analysis. Schizophr Res 2010; 123:225–233CrossrefGoogle Scholar

Bereza BG, Machado M, Ravindran AV, Einarson TR: Evidence-based review of clinical outcomes of guideline-recommended pharmacotherapies for generalized anxiety disorder. Can J Psychiatry 2012; 57:470–478CrossrefGoogle Scholar

Davidson JR: First-line pharmacotherapy approaches for generalized anxiety disorder. J Clin Psychiatry 2009; 70(Suppl 2):25–31CrossrefGoogle Scholar

Woods SW, Morgenstern H, Saksa JR, Walsh BC, Sullivan MC, Money R, Hawkins KA, Gueorguieva RV, Glazer WM: Incidence of tardive dyskinesia with atypical versus conventional antipsychotic medications: a prospective cohort study. J Clin Psychiatry 2010; 71:463–474CrossrefGoogle Scholar

Yager J, Devlin JM, Halmi KA, Herzog DB, Mitchell III JE, Powers P, Zerbe KJ. Guideline Watch (Aug 2012): Practice Guideline for the Treatment of Patients with Eating Disorders, 3rd Edition. American Psychiatric Association, Arlington VA (link to on POL or better to just provide the pdf and link)Google Scholar

Kishi T, Kafantaris V, Sunday S, Sheridan EM, Correll CU: Are antipsychotics effective for the treatment of anorexia nervosa? Results from a systematic review and meta-analysis. J Clin Psychiatry 2012; 73:e757–e766. Available at doi: 10.4088/JCP.12r07691CrossrefGoogle Scholar

Kafantaris V, Leigh E, Hertz S, Berest A, Schebendach J, Sterling WM, Saito E, Sunday S, Higdon C, Golden NH, Malhotra AK: A placebo-controlled pilot study of adjunctive olanzapine for adolescents with anorexia nervosa. J Child Adolesc Psychopharmacol 2011; 21:207–212CrossrefGoogle Scholar

Kishi T, Sevy S, Chekuri R, Correll CU: Antipsychotics for primary alcohol dependence: a systematic review and meta-analysis of placebo-controlled trials. J Clin Psychiatry 2013; 74:e642–e654. Available at doi: 10.4088/JCP.12r08178CrossrefGoogle Scholar

Kishi T, Matsuda Y, Iwata N, Correll CU: Antipsychotics for cocaine or psychostimulant dependence: systematic review and meta-analysis of randomized, placebo-controlled trials. J Clin Psychiatry 2013; 74:e1169–e1180CrossrefGoogle Scholar

NIDA research report series. Available at: http://www.drugabuse.gov/publications/research-reports/methamphetamine-abuse-addictionGoogle Scholar

Coffin PO, Santos GM, Das M, Santos DM, Huffaker S, Matheson T, Gasper J, Vittinghoff E, Colfax GN: Aripiprazole for the treatment of methamphetamine dependence: a randomized, double-blind, placebo-controlled trial. Addiction 2013; 108:751–761CrossrefGoogle Scholar

Schutte-Rodin S, Broch L, Buysse D, Dorsey C, Sateia M: Clinical guideline for the evaluation and management of chronic insomnia in adults. J Clin Sleep Med 2008; 4:487–504 Available at : http://www.aasmnet.org/Resources/clinicalguidelines/040515.pdfCrossrefGoogle Scholar

Shah C, Sharma TR, Kablinger A: Controversies in the use of second generation antipsychotics as sleep agent. Pharmacol Res 2014; 79:1–8CrossrefGoogle Scholar

Maayan L, Correll CU: Weight gain and metabolic risks associated with antipsychotic medications in children and adolescents. J Child Adolesc Psychopharmacol 2011; 21:517–535. Available at doi: 10.1089/cap.2011.0015CrossrefGoogle Scholar

Leichsenring F, Leibing E, Kruse J, New AS, Leweke F. Borderline personality disorder. Lancet. 2011; 377:74–84. Reprinted in FOCUS: March 2013; 11;249-260Google Scholar

Gunderson JG, Weinberg I, Choi-Kain L: Borderline personality disorder. FOCUS 2013; 11:129–145LinkGoogle Scholar

Ripoll LH: Psychopharmacologic treatment of borderline personality disorder. Dialogues Clin Neurosci 2013; 15:213–224Google Scholar

Birnbaum ML, Saito E, Gerhard T, Winterstein A, Olfson M, Kane JM, Correll CU: Pharmacoepidemiology of antipsychotic use in youth with ADHD: trends and clinical implications. Curr Psychiatry Rep 2013; 15:382CrossrefGoogle Scholar

Muscatello MR, Bruno A, Pandolfo G, Micò U, Scimeca G, Romeo VM, Santoro V, Settineri S, Spina E, Zoccali RA: Effect of aripiprazole augmentation of serotonin reuptake inhibitors or clomipramine in treatment-resistant obsessive-compulsive disorder: a double-blind, placebo-controlled study. J Clin Psychopharmacol 2011; 31:174–179CrossrefGoogle Scholar

Simpson HB, Foa EB, Liebowitz MR, Huppert JD, Cahill S, Maher MJ, McLean CP, Bender J, Marcus SM, Williams MT, Weaver J, Vermes D, Van Meter PE, Rodriguez CI, Powers M, Pinto A, Imms P, Hahn CG, Campeas R: Cognitive-behavioral therapy vs risperidone for augmenting serotonin reuptake inhibitors in obsessive-compulsive disorder: a randomized clinical trial. JAMA Psychiatry 2013; 70:1190–1199CrossrefGoogle Scholar

Dold M, Aigner M, Lanzenberger R, Kasper S: Antipsychotic augmentation of serotonin reuptake inhibitors in treatment-resistant obsessive-compulsive disorder: a meta-analysis of double-blind, randomized, placebo-controlled trials. Int J Neuropsychopharmacol 2013; 16:557–574CrossrefGoogle Scholar

Weiden PJ: EPS profiles: the atypical antipsychotics are not all the same. J Psychiatr Pract 2007; 13:13–24CrossrefGoogle Scholar