The World Journal of Biological Psychiatry - CiteSeerX

82

Transcript of The World Journal of Biological Psychiatry - CiteSeerX

Editorial

Are placebo-controlled studies required in order to prove efficacy of antidepressants?Hans-Jurgen Moller ............................................................................................................................... 130

Reviews/Mini-Reviews

World Federation of Societies of Biological Psychiatry (WFSBP) Guidelines for BiologicalTreatment of Schizophrenia, Part 1: Acute treatment of schizophrenia

Peter Falkai, Thomas Wobrock, Jeffrey Lieberman, Birte Glenthoj, Wagner F. Gattaz,Hans-Jurgen Moller, WFSBP Task Force on Treatment Guidelines for Schizophrenia ...................... 132

Cognitive neuropsychiatry: Conceptual, methodological and philosophical perspectivesJakob Hohwy, Raben Rosenberg ............................................................................................................ 192

Original Investigations/Summaries of Original Research

Agitated depression in bipolar II disorderFranco Benazzi ...................................................................................................................................... 198

Instructions to Authors ........................................................................................................ 206

Author Disclosure Declaration ........................................................................................ 207

The World Journal of Biological PsychiatryVolume 6, No 3, 2005

Contents

EDITORIAL

Are placebo-controlled studies required in order to prove efficacy ofantidepressants?

Imagine you are presented with the following study

results:

Study 1. The results of a randomised, double-blind

control-group study comparing a new antidepressant

with the standard medication amitriptyline show

that the average score on the Hamilton Depression

Scale at the end of the 6-week treatment phase is 4

points higher in the group that received the new

antidepressant than in the amitriptyline group,

although baseline values were comparable. The

difference is not statistically significant. A total of

40 patients were included in each group.

Study 2. In this randomised, double-blind, control-

group study, the same new antidepressant was

compared with the standard medication amitripty-

line. After 6 weeks treatment there is no difference in

the average score on the Hamilton Depression Scale

between the two groups, whereby baseline scores are

comparable. A total of 100 patients were included in

each group. The average improvement was 15% of

the baseline value on the Hamilton Depression

Scale.

One may be tempted to presume that both studies

prove that the new antidepressant has the same

antidepressive efficacy as the standard drug. Only

after more careful analysis do you suspect that due to

the small sample size the first study does not have

enough statistical power to show that the fairly large

difference in efficacy in favour of the new antide-

pressant is statistically significant (b error problem).

Perhaps you brush aside your reservations anyway

because you don’t consider the difference clinically

relevant, perhaps due to an overestimation of the

effect strength of antidepressants, although it is in

the order of magnitude of the difference in efficacy

between an antidepressant and a placebo.

With respect to the second study, which obviously

does not have the problem of insufficient statistical

power �/ after all 100 patients were included in each

treatment arm �/ you don’t find any such methodo-

logical shortcomings, even after more careful analy-

sis. However, you do notice that the average

percentage improvement after 6 weeks treatment is

fairly small in both groups. That still may not give

you a reason to doubt the conclusion that in this

study the efficacy of the new antidepressant is

equivalent to that of the standard antidepressant. If

you were more experienced with assessing the results

of such antidepressant studies, you would probably

conclude that it was a so-called ‘insensitive study’ in

which the standard antidepressant was unable to

demonstrate adequate efficacy, for example, because

the sample was made up of partially or almost fully

refractory patients, and that therefore the standard

antidepressant, which is normally effective, was

unable to show better efficacy than the new anti-

depressant, which is actually non-effective or has

only limited efficacy.

Of course, these examples are fabricated with the

purpose of illustrating the wrong conclusions that

may arise when a new antidepressant is tested

against a standard medication. As a consequence

of the ethical reservations against placebo-controlled

antidepressant studies, potentially millions of people

are treated with antidepressants whose efficacy is

rather on the level of a placebo than of a tricyclic

antidepressant.

Even today in many countries there are still

reservations about placebo-controlled antidepres-

sant studies, even among psychiatrists with experi-

ence in empirical research. On the other hand, false

conclusions that may arise from comparative studies

versus active standard medications are being in-

creasingly better described and are now well known

by experts in clinical psychopathology. Placebo-

controlled studies have been required for registration

by the American licensing authority the FDA for

several decades already. The European licensing

authority, EMEA, has now followed this stipulation

and its relevant guideline, which has just been

reissued, also includes placebo-controlled studies

as a requirement for registration. The various

European countries are required to comply to this

European consensus, i.e. no antidepressant can be

registered in a European country without demon-

strating efficacy in placebo-controlled studies. It is

The World Journal of Biological Psychiatry, 2005; 6(3): 130�/131

ISSN 1562-2975 print/ISSN 1814-1412 online # 2005 Taylor & Francis

DOI: 10.1080/15622970510030108

hard to imagine that two such eminent licensing

authorities as the American FDA and the European

EMEA, which are under political and social control,

did not pay enough attention to the inherent ethical

problems when stipulating their demands. One

therefore has to assume that they see placebo-

controlled studies as irrefutably necessary in order

to test adequately the efficacy of an antidepressant.

Although ethical counterarguments have to be

considered, in the view of these licensing authorities

they can be largely reduced or even excluded by a

relevant differentiation in detailed aspects of the

study design.

The fifth revision of the Declaration of Helsinki

(World Medical Association 2000) determined that

‘The benefits, risks, burdens and effectiveness of a

new method should be tested against those of the

best current prophylactic, diagnostic and therapeutic

methods’. This statement appeared to leave the

randomised, control-group comparison between

the new antidepressant and the standard medication

as the only possibility and to exclude the placebo-

controlled comparison. The cited statement there-

fore seemed to contradict the demand for placebo-

controlled studies which was derived from scientific

arguments, and was therefore so strongly criticised

that the World Medical Association (WMA) saw

itself constrained to re-discuss this problem in a

working group. In a note of clarification of the

revised declaration of Helsinki the WMA notes

that the use of placebo in drug trials is justified if

there are compelling and scientifically methodologi-

cal reasons, and provided that the patients who

receive placebo will not be subject to any additional

risk of serious or irreversible harm.

The EMEA (EMEA 2001) picked up on this

controversy and made the following remarks: the

placebo response, which is also included as a sub-

factor in the overall response to the verum, can differ

considerably; placebo-controlled comparisons are

therefore necessary in order to make an adequate

assessment of efficacy. The comparison with placebo

also helps to differentiate between symptoms of the

illness and adverse events. Placebo-controlled stu-

dies may be ethically acceptable if the use of placebo

is necessary for important methodological reasons,

in order to demonstrate the efficacy and tolerability

of a treatment. The most important aspect of the

compliance of placebo-controlled studies with ethi-

cal standards is that patients are not subjected to the

risk of severe or irreversible harm.

Hans-Jurgen Moller

Chief Editor

Correspondence:

Prof. Hans-Jurgen Moller

Department of Psychiatry

Ludwig-Maximilians-University

Nussbaumstr. 7

80336 Munich

Germany

Tel: �/49 89 5160 5501

Fax: �/49 89 5160 5522

E-mail: [email protected]

References

EMEA. 2001. EMEA/CPMP position statement on the use

of placebo in clinical trials with regard to the revised Declara-

tion of Helsinki [www.emea.eu.int/pdfs/human/press/pos/1742

401en.pdf].

World Medical Association. 2000. Declaration of Helsinki, 5th

(Edinburgh) amendment [www.wma.net/e/policy/b3.htm].

Editorial 131

REVIEW

World Federation of Societies of Biological Psychiatry (WFSBP)Guidelines for Biological Treatment of Schizophrenia, Part 1: Acutetreatment of schizophrenia

PETER FALKAI1, THOMAS WOBROCK1, JEFFREY LIEBERMAN2, BIRTE GLENTHOJ3,

WAGNER F. GATTAZ4, & HANS-JURGEN MOLLER5, WFSBP Task Force on Treatment

Guidelines for Schizophrenia*

1Department of Psychiatry and Psychotherapy, University of Saarland, Homburg/Saar, Germany, 2Department of

Psychiatry, College of Physicians and Surgeons, Columbia University, New York State Psychiatric Institute, Lieber Center for

Schizophrenia Research, New York, USA, 3Department of Psychiatry, University of Copenhagen, Bispebjerg Hospital,

Denmark, 4Department of Psychiatry, University of Sao Paulo, Brazil, and 5Department of Psychiatry and Psychotherapy,

Ludwig-Maximilians-University, Munich, Germany

AbstractThese guidelines for the biological treatment of schizophrenia were developed by an international Task Force of the WorldFederation of Societies of Biological Psychiatry (WFSBP). The goal during the development of these guidelines was toreview systematically all available evidence pertaining to the treatment of schizophrenia, and to reach a consensus on a seriesof practice recommendations that are clinically and scientifically meaningful based on the available evidence. Theseguidelines are intended for use by all physicians seeing and treating people with schizophrenia. The data used for developingthese guidelines have been extracted primarily from various national treatment guidelines and panels for schizophrenia, aswell as from meta-analyses, reviews and randomised clinical trials on the efficacy of pharmacological and other biologicaltreatment interventions identified by a search of the MEDLINE database and Cochrane Library. The identified literaturewas evaluated with respect to the strength of evidence for its efficacy and then categorised into four levels of evidence (A�/

D). This first part of the guidelines covers disease definition, classification, epidemiology and course of schizophrenia, aswell as the management of the acute phase treatment. These guidelines are primarily concerned with the biologicaltreatment (including antipsychotic medication, other pharmacological treatment options, electroconvulsive therapy,adjunctive and novel therapeutic strategies) of adults suffering from schizophrenia.

Key words: Schizophrenia, acute phase treatment, evidence-based medicine, practice guidelines, biological treatment,

antipsychotics

EXECUTIVE SUMMARY OF

RECOMMENDATIONS

General recommendations

Specific treatment is indicated for patients who meet

diagnostic criteria for schizophrenia, a schizophrenic

episode or psychotic symptoms related to schizo-

phrenic disorder. An assessment of mental and

physical health to evaluate relevant psychiatric and

medical comorbid conditions, psychosocial circum-

stances and quality of life should be undertaken

regularly. When a person presents psychotic symp-

toms for the first time a careful diagnostic evaluation

* A. Carlo Altamura (Italy), Nancy Andreasen (USA), Thomas R.E. Barnes (UK), Helmut Beckmann (Germany), Jorge Ciprian-

Ollivier (Argentina), Tim Crow (UK), Anthony David (UK), Michael Davidson (Israel), Bill Deakin (UK), Helio Elkis (Brazil), Lars Farde

(Sweden), Wolfgang Gaebel (Germany), Bernd Gallhofer (Germany), Jes Gerlach (Denmark), Steven Richard Hirsch (UK), Carlos R.

Hojaij (Australia), Assen Jablensky (Australia), John Kane (USA), Takuja Kojima (Japan), Lars von Knorring (Sweden), Patrick McGorry

(Australia), Herbert Meltzer (USA), Driss Moussaoui (Marokko), Franz Muller-Spahn (Switzerland), Jean-Pierre Olie (France), A.

Pacheco Palha (Portugal), Mitsumoto Sato (Japan), Heinrich Sauer (Germany), Nina Schooler (USA), Daniel Weinberger (USA), Shigeto

Yamawaki (Japan).

Correspondence: Dr. med. Thomas Wobrock, M.D., Department of Psychiatry and Psychotherapy, University of Saarland, Kirrberger

Strasse, D-66421 Homburg/Saar, Germany. Tel: �/49 6841 1624216. Fax: �/49 6841 1624270. E-mail: thomas.wobrock@uniklinik-

saarland.de

The World Journal of Biological Psychiatry, 2005; 6(3): 132�/191

ISSN 1562-2975 print/ISSN 1814-1412 online # 2005 Taylor & Francis

DOI: 10.1080/15622970510030090

should be performed, including laboratory investiga-

tion and imaging techniques (cerebral CT or MRI)

in order to exclude organic brain disease. After the

initial assessment of the patient’s diagnosis and

establishment of a therapeutic alliance, a treatment

plan must be formulated and implemented. This

formulation involves the selection of the treatment

modalities, the specific type(s) of treatment, and the

treatment setting. Periodic reevaluation of the diag-

nosis and the treatment plan is essential. Engage-

ment of the family and other significant support

persons, with the patient’s permission, is recom-

mended to further strengthen the therapeutic effort.

The goals and strategies of treatment vary according

to the phase and severity of illness. In the acute

phase of treatment (lasting weeks to months), which

is defined by an acute psychotic episode, major goals

are to develop an alliance with the patient

and family, to prevent harm, control disturbed

behaviour, reduce the severity of psychosis and

associated symptoms (e.g., agitation, aggression,

negative symptoms, affective symptoms), determine

and address the factors that led to the occurrence of

the acute episode and to effect a rapid return to the

best level of functioning. Special attention should be

paid to the presence of suicidal ideation, intent or

plan, and the presence of command hallucinations.

The patient should be provided with information on

the nature and management of the illness, including

the benefits and side effects of the medication, in a

form that is appropriate to his or her ability to

assimilate the information. In the acute treatment

phase, the main emphasis is on pharmacotherapeutic

(and other somatic) interventions. Therefore anti-

psychotic therapy should be initiated as a necessary

part of a comprehensive package of care that

addresses the individual’s clinical, emotional and

social needs.

Specific treatment recommendations

In first-episode psychosis antipsychotic pharmacologi-

cal treatments should be introduced with great care

due to the higher risk of extrapyramidal symptoms

(EPS). Appropriate strategies include gradual intro-

duction of low-dose antipsychotic medication with

careful explanation. The first-line use of second-

generation antipsychotic medication (SGAs) (and

alternatively the use of first-generation antipsycho-

tics (FGAs) at the lower end of the standard dose

range) is the preferred treatment for a person

experiencing a first episode of schizophrenia. This

recommendation is mainly based on the better

tolerability and reduced risk of tardive dyskinesia

associated with the atypical antipsychotics. Skilled

nursing care, a safe and supportive environment, and

regular and liberal doses of benzodiazepines may be

essential to relieve distress, insomnia and beha-

vioural disturbances secondary to psychosis, while

antipsychotic medication takes effect.

In multiple episodes the most common contributors

to symptom relapse are antipsychotic medication

non-adherence, substance use and stressful life

events, although relapses are not uncommon as a

result of the natural course of the illness, despite

continuing treatment. If nonadherence is suspected,

it is recommended that the reasons for it be

evaluated and considered in the treatment plan. It

is recommended that pharmacological treatment

should be initiated promptly, because acute psycho-

tic exacerbations are associated with emotional

distress, and a substantial risk of dangerous beha-

viours. Given the advantages of second-generation

antipsychotics (SGAs), these antipsychotics gener-

ally seem preferable, although in principal all anti-

psychotics have their place in the treatment of acute

schizophrenia. The selection of an antipsychotic

medication is guided by the patient’s previous

experience of symptom response and side effects,

intended route of administration, the patient’s pre-

ferences for a particular medication, the presence of

comorbid medical conditions, and potential interac-

tions with other prescribed medications. The dose

may be titrated as quickly as tolerated to the target

therapeutic dose of the antipsychotic medication

(e.g., 300�/1000 chlorpromazine (CPZ) equivalents

for FGAs) while monitoring the patient’s clinical

status. Especially when using FGAs it is recom-

mended to keep the dose as low as possible to reduce

the risk of extrapyramidal side effects. Rapid

dose escalation, high loading doses and treatment

with high doses above the mentioned dose range

do not have proven superior efficacy, but have

been associated with increased side effects. For

patients presenting with high degrees of agitation

in an emergency setting there is evidence for

superior efficacy with the combination of benzodia-

zepines.

In assessing treatment-resistant schizophrenia (TRS)

or partial response, multidimensional evaluation

should consider persistent positive or negative

symptoms, cognitive dysfunction with severe impair-

ment, bizarre behaviour, recurrent affective symp-

toms, deficits in vocational and social functioning

and a poor quality of life. The target symptoms

should be precisely defined. It is important to

evaluate carefully whether there is insufficient im-

provement in the target symptoms, despite treat-

ment at the recommended dosage for a duration of

at least 6�/8 weeks with at least two antipsychotics,

one of which should be an atypical antipsychotic.

Adherence should be ensured, if necessary by

WFSBP Guidelines for Biological Treatment of Schizophrenia 133

checking drug concentrations. In individuals with

clearly defined TRS, clozapine should be introduced

as treatment of choice because of clozapine’s super-

ior efficacy in this regard. Treatment alternatives in

case of nonresponse may be other SGAs, augmenta-

tion strategies (antidepressants, mood stabilisers) in

relation to target symptoms, combination of anti-

psychotics (limited evidence for risperidone or

sulpiride in combination with clozapine) and, as

the last treatment option, electroconvulsive therapy

(ECT).

For patients presenting with catatonic features the

option of ECT may be considered earlier when

insufficient response on benzodiazepines is ob-

served.

Treatment of negative symptoms begins with asses-

sing the patient for syndromes that can cause

secondary negative symptoms. The treatment of

such secondary negative symptoms consists of treat-

ing their cause, e.g., antipsychotics for primary

positive symptoms, antidepressants for depression,

anxiolytics for anxiety disorders, or antiparkinsonian

agents, antipsychotic dose reduction or switch to a

SGA for extrapyramidal side effects. For primary

negative symptoms treatment with SGAs is recom-

mended. The greatest level of evidence is for

amisulpride, but it has not been clearly proven to

have more efficacy.

For patients with concomitant substance use

disorders, a comprehensive integrated treatment is

recommended in which the same clinicians or team

of clinicians provide treatment for both diseases.

There is limited evidence that SGAs, especially

clozapine, but also risperidone and olanzapine, are

beneficial for dual diagnosis patients. This may be

due to reduced severity of EPS and decrease of

craving.

Depressive symptoms that occur during the acute

psychotic phase usually improve as patients recover

from the psychosis. There is also evidence suggesting

that depressive symptoms are reduced by antipsy-

chotic treatment, with comparison trials finding

that SGAs may have greater efficacy in treating

depressive symptoms than FGAs. Antidepressants

may be added as an adjunct to antipsychotics when

syndromal criteria for major depressive episode are

met.

There is evidence to suggest that both first- and

second-generation antipsychotic medications may

reduce the risk of suicide . In several studies, clozapine

demonstrated the most consistent reduction of

suicide rates and persistent suicidal behaviour.

Further treatment strategies for medical and psy-

chiatric comorbid conditions are extensively dis-

cussed in the full text of the guideline.

Schizophrenia

Introduction

Schizophrenia is a major psychotic disorder (or

cluster of disorders) that usually appears in late

adolescence or early adulthood. Despite modern

treatment techniques schizophrenia still presents an

enormous burden to the patients and their relatives.

In most cases there is an impairment in occupational

or social functioning, characterised by social with-

drawal, loss of interest or ability to function at school

or work, change in personal hygienic habits or

unusual behaviour still in the prodromal phase

(Loebel et al. 1992; Hafner and an der Heiden

2003).

Schizophrenia presents different symptoms in

multiple domains in great heterogeneity across

individuals and also variability within individuals

over time. Due to systematic observation of psycho-

pathology, positive and negative phenomena can be

separated (Andreasen 1982; Crow 1985). Positive

symptoms in a broader manner include delusions or

delusional ideation, hallucinations, disturbance of

association, catatonic symptoms, agitations, feelings

of alien influence and suspiciousness. Negative

symptoms include restricted range and intensity of

emotional expression, reduced thought and speech

productivity and social withdrawal associated with a

reduced initiation of goal-directed behaviour. There-

fore, negative components could be defined as

affective flattening, alogia, anhedonia and avolition.

As a third category, disorganised symptoms include

disorganised speech, disorganised behaviour and

poor attention.

The currently used diagnostic criteria for schizo-

phrenia (DSM-IV, ICD-10) define schizophrenia as

a discrete category. Schizophrenia can be diagnosed

if organic brain disease has been excluded, and an

essential feature of symptoms or at least one core

symptom has to be present for a significant length of

time during a 1-month period (or for a shorter time

if successfully treated) (according to ICD-10), or

for at least 6 months (according to DSM-IV) (see

Table I).

According to DSM-IV or ICD-10, subtypes of

schizophrenia are defined by their predominant

symptoms. These subtypes include (1) the hebe-

phrenic subtype, in which flat or inappropriate affect

is prominent, (2) the catatonic subtype, in which

characteristic autosymptoms (catatonia) are perma-

nent, (3) the disorganised subtype, in which dis-

organised speech and behaviour dominate, (4) the

paranoid subtype, in which predominantly dilutions

or auditory hallucinations are present, (5) the

undifferentiated subtype, which is a non-specific

category, and (6) schizophrenia simplex, so-called

134 P. Falkai et al.

due to the course and permanent negative symptoms

of the disease. The subtypes of schizophrenia are

listed in Table II.

These guidelines and the presented recommenda-

tions focus on the acute, and continuation and

maintenance treatment of schizophrenia. The guide-

lines may help clinicians, service users and caregivers

become aware of the different treatments that are

available, and may be useful in deciding which

treatment to apply as they include the level of

evidence available for each treatment. The first

part (Part 1) of these guidelines covers epidemiol-

Table I. Diagnostic criteria of schizophrenia or schizophrenic episode.

ICD-10

Characteristic symptomatology:

1. One month or more, in which a significant portion of time is taken up by one very clear symptom or two less clear symptoms:

A. Thought echo, thought insertion or withdrawal, and thought broadcasting;

B. Delusions of control, influence, or passivity, clearly referred to body or limb movements or specific thoughts, actions, or sensations;

delusional perception;

C. Hallucinatory voices giving a running commentary on the patient’s behaviour, or discussing the patient among themselves, or other

types of hallucinatory voices coming from some part of the body;

D. Persistent delusions of other kinds that are culturally inappropriate and completely impossible, such as religious or political

identity, or superhuman powers and abilities (e.g. being able to control the weather, or being in communication with aliens

from another world);

Or

2. At least two of the following:

E. Persistent hallucinations in any modality, when accompanied either by fleeting or half-formed delusions without clear affective

content, or by persistent over-valued ideas, or when occurring every day for weeks or months on end;

F. Breaks or interpolations in the train of thought, resulting in incoherence or irrelevant speech, or neologisms;

G. Catatonic behaviour, such as excitement, posturing, or waxy flexibility, negativism, mutism, and stupor;

H. ‘‘Negative’’ symptoms such as marked apathy, paucity of speech, and blunting or incongruity of emotional responses, usually

resulting in social withdrawal and lowering of social performance; it must be clear that these are not due to depression or to

neuroleptic medication;

Or

3. I. A significant and consistent change in the overall quality of some aspects of personal behaviour, manifest as loss of interest,

aimlessness, idleness, a self-absorbed attitude, and social withdrawal.

Duration: One of A�/D, or two of E�/H present for 1 month, or I present for more than 1 year (Simple schizophrenia)

DSM-IV

A. Characteristic symptoms: Two (or more) of the following, each present for a significant portion of time during a 1-month period (or less if

successfully treated):

1. Delusions.

2. Hallucinations

3. Disorganised speech (frequent derailment or incoherence)

4. Grossly disorganised or catatonic behaviour

5. Negative symptoms, i.e., affective flattening, alogia or avolition.

Note: Only: one criterion A symptom required if delusions are bizarre or hallucinations consist of a voice keeping up a running commentary

on the person’s behaviour or thoughts, or two or more voices conversing with each other.

B. Social/Occupational functioning

For significant portion of time, since the onset of the disturbance, one or more major areas of functioning such as work, interpersonal

relations, or self-care are markedly below the level achieved prior to onset (or when the onset is in childhood or adolescence, the feilure to

achieve expected level of interpersonal, academic, or occupational achievement).

C. Continuous signs of the disturbance persist for at least 6 months. This 6-month period must include at least one month of symptoms (or

less if successfully treated) that meet Criterion A (i.e., active-phase symptoms) and may include periods of prodromal or residual

symptoms. During these prodromal or residual periods, the signs of the disturbance may be manifested by only negative symptoms or tow

or more symptoms listed in Criterion A present in an attenuated form (e.g., odd beliefs, unusual perceptual experiences).

Exclusion: The diagnosis is not made in presence of extensive depressive or manic symptoms unless it is clear that schizophrenic symptoms

antedated the affective disturbance. The disturbance is not due to substance intoxication, dependence or withdrawal, or overt brain

disease.

D. Schizoaffective and Mood Disorder exclusion: Schizoaffective Disorder and Mood Disorder with Psychotic features have been ruled out

because wither (1) no Major depressive, Manic, or Mixed episodes have occurred concurrently with the active phase symptoms; or (2) if

mood episodes have occurred during active phase symptoms, their total duration has been brief relative to the duration of the active and

residual periods.

E. Substance/general medical condition exclusion: The disturbance is mot due to the direct physiological effects of a substance (e.g., a drug

of abuse, a medication) or a general medical condition.

F. Relationship to a pervasive developmental disorder: If there is a history of Autistic Disorder or another Pervasive Developmental

Disorder, the additional diagnosis of schizophrenia is made only if prominent delusions or hallucinations are also present for at least a

month (or less if successfully treated).

WFSBP Guidelines for Biological Treatment of Schizophrenia 135

ogy, course and aetiology of schizophrenia, the

pharmacological properties, classification and side

effects of antipsychotics and adjunctive agents,

assessment before and during treatment and the

management of the acute phase treatment.

Epidemiology and course of schizophrenia

Schizophrenia is a relatively common illness. Life-

time prevalence varies for methodological reasons,

but the mean prevalence is estimated to be nearly

one case per 100 persons (1%) in the general

population. While limited to broad or narrow

diagnostic criteria, the mean incidence of schizo-

phrenia reported in epidemiological studies is

0.11�/0.24 per 1000 persons (Jablensky et al.

1992). The average rates for men and women are

nearly similar, but the mean age of onset is about 5

years later for women than for men (Hafner and an

der Heiden 2003).

In most cases the onset of schizophrenic symp-

toms, according to diagnostic criteria, is preceded by

a prodromal period characterised by early signs of

impairment in personal and social functioning. The

average length of the prodromal phase is between 2

and 5 years (Beiser et al. 1993; Hafner et al. 1993)

and is often neither recognised nor treated and

therefore called the duration of untreated illness

(DUI). The prodromal phase is followed by the

onset of acute schizophrenia, marked by character-

istic positive symptoms of hallucinations, delusions

and behavioural disturbances. In most industrialised

countries 1�/2 years pass before adequate treatment

is initiated (Johnstone et al. 1986; Hafner et al.

1993). The time from the first occurrence of

psychotic symptoms until treatment begins is de-

fined as the duration of untreated psychosis (DUP)

(McGlashan and Johanessen 1996). Research indi-

cates that delayed access to health services and

treatment is associated with slower or less complete

recovery (Loebel et al. 1992) and increased risk of

relapse in the subsequent 2 years (Johnstone et al.

1996). The acute phase refers to the presence of

florid psychotic features. After positive symptoms

have diminished, in a certain number of cases

negative symptoms similar to the symptoms in the

prodromal period remain.

The course and pattern of schizophrenia varies

considerably. During the stabilisation or recovery

phase, psychotic symptoms decrease in severity. The

stabilisation period lasts about 6 months and is

followed by the stable phase. In the stable phase

negative and residual positive symptoms that may be

present are relatively consistent in severity and

magnitude. A certain number of patients (nearly

20�/30%) display no prominent symptoms after the

first episode (Hafner and an der Heiden 2003).

Acute exacerbations may interrupt the stable phase

and require additional treatment or interventions.

Longitudinal neuropsychological assessment showed

that patients with schizophrenia have considerable

cognitive dysfunction in the first 4�/5 years of illness,

and after this period there is little evidence for

deterioration (Hoff et al. 1999). Approximately

50% of people with schizophrenia treated in stan-

dard services will relapse and require readmission

within the first 2 years, and up to 80% within a

5-year period (Robinson et al. 1999); about 10�/25%

will have no further admissions (Fenton et al. 1987;

Hegarty et al. 1994). Before relapse occurs warning

signs often appear, which usually consist of non-

psychotic symptoms followed by emotional distur-

bance and mild psychotic symptoms over a period of

4�/12 weeks (Birchwood et al. 1989; Gaebel et al.

1993). There is some evidence that the negative

symptoms may become steadily more prominent in

some individuals during the course of illness

(McGlashan and Fenton 1993; Moller et al. 2002).

Predictors associated, on average, with better out-

comes are later age at onset, female gender, married

marital status, a sociable premorbid personality,

good premorbid adjustment and functioning, higher

IQ, a psychoreactive trigger of onset, an acute onset

of symptoms, predominantly affective or positive

symptoms and no disorganised or negative symp-

toms at onset, fewer prior episodes, a phasic pattern

of episodes and remissions, lack of family history of

schizophrenia and a low level of expressed emotions

in the family (Hegarty et al. 1994; Davidson and

McGlashan 1997; Bottlender et al. 2000, 2002,

2003; Hafner and an der Heiden 2003).

Other, comorbid mental disorders and general

medical conditions are often found with schizophre-

nia. After the elevated suicide risk, cardiovascular

disorders, respiratory and infectious diseases, acci-

Table II. Subtypes of schizophrenia.

Subtypes ICD-10

F 20.0 Paranoid schizophrenia

F 20.1 Hebephrenic schizophrenia

F 20.2 Catatonic schizophrenia

F 20.3 Undifferentiated schizophrenia

F 20.4 Post schizophrenic depression

F 20.5 Residual schizophrenia

F 20.6 Simple schizophrenia

F 20.8 Other schizophrenia

F 20.9 Schizophrenia unspecified

DSM-IV

Schizophrenia, paranoid type

Schizophrenia, catatonic type

Schizophrenia, disorganized type

Schizophrenia, undifferentiated type

Schizophrenia, residual type

136 P. Falkai et al.

dents and traumatic injuries serve as main causes for

the excess mortality of schizophrenia (Brown et al.

2000). One of the most frequent comorbid diseases

is substance abuse disorder, which occurs in

15�/71% of patients with schizophrenia (Soyka et

al. 1993; Kovasznay et al. 1997; Bersani et al. 2002).

Factors influencing the risk of abusing drugs are

associated with the former and current social en-

vironment and premorbid personality (Arndt et al.

1992). Comorbid conditions can worsen the course

and complicate treatment (Linszen et al. 1994).

Aetiology

Schizophrenia is a disorder with a complex aetiology.

Research has attempted to determine the role of

specific biological variables, such as genetic and

biochemical factors und subtle changes in the brain

morphology. According to the neurodevelopment

hypothesis, schizophrenia appears as a result of

disturbed brain development during the pre- or

perinatal period, but the specificity of the nature of

early brain disruption or pathogenesis is not yet

defined and remains unclear (for review see Mar-

enco and Weinberger 2000). Genetic components

could only explain half of the risk to develop

schizophrenia, and pre- or perinatal complications

are responsible for 1�/2% of the risk (Gottesman and

Bertelsen 1989).

Therefore the favoured model of the illness is the

‘vulnerability-stress-coping-model’ (Nuechterlein and Daw-

son 1984). This concept proposes that vulnerability

will result in the development of symptoms when

environmental stresses are present and coping me-

chanisms fail. Vulnerability factors based on a

biological component with a genetic background

interact with complex physical, environmental and

psychological vulnerability factors. Biochemical the-

ories focus mainly on the ‘dopamine hypothesis’.

The dopamine hypothesis implies that there is an

increased production of this neurotransmitter or an

over-sensitivity of dopamine receptors in a certain

brain region (mesolimbic system), resulting in hy-

perexcitability and the appearance of positive symp-

toms and a hypodopaminergic state in frontal brain

regions, followed by negative symptoms (for review

see Sedvall and Farde 1995). This hypothesis is

supported by the successful treatment of psychotic

symptoms by agents blocking D2 receptors in the

mesolimbic system. Besides dopamine, other neuro-

transmitters like serotonin and glutamate seems to

be involved in the pathophysiology of schizophrenia.

The blockade of serotonin 5-HT2A receptors and the

preferential blockade of specific subtypes of dopa-

mine receptors was hypothesised to be a relevant

mechanism for the efficacy of atypical or second-

generation antipsychotics in treating negative symp-

toms (Moller 2003).

Assessment before and during treatment

A full assessment of health and social care needs

should be undertaken regularly, including assess-

ment of social functioning and quality of life. Every

patient should have a careful initial assessment,

including complete psychiatric and general medical

histories. In order to exclude most general medical

conditions that can contribute to psychotic symp-

toms, physical and mental status examinations,

including a neurological examination, are indicated.

Basic laboratory tests should be conducted and

should include CBC, measurements of blood elec-

trolytes, lipids and glucose, tests of liver, renal,

thyroid function, and additionally HIV, hepatitis,

and syphilis screening, when indicated. Ordinarily, a

urine screen should be performed to assess recent

use of substances. In addition to establishing base-

lines for the administration of psychotropic medica-

tion, these tests examine the patient for illnesses that

can mimic schizophrenia and illnesses that are often

comorbid with schizophrenia and require modifica-

tion of the treatment plan. It may be useful to

assess blood levels of antipsychotic medication to

establish whether the patient has been taking his

or her medication. Tests to assess other general

medical needs of patients should be considered

(e.g., measurement of the human chorionic gonado-

tropin b subunit in women of childbearing age).

Magnetic resonance imaging (MRI) or computed

tomography (CT) and EEG should be performed in

patients with a first episode and when the clinical

picture is unclear, or when there are abnormal

findings from a routine examination. Patients

with preexisting cardiac disease need to be carefully

monitored for ECG abnormalities since antipsycho-

tic medications can be cardiotoxic. Such knowledge

influences the choice of medication. Neuropsycho-

logical tests are generally not useful in making

a diagnosis of schizophrenia during the acute

phase but are useful after stabilisation to evaluate

cognitive deficits, which can affect the treatment

plan. Special circumstances (before and after the

application of ECT) may require neuropsychological

evaluation.

Special attention should be paid to the presence of

suicidal ideation, intent, or plan, and the presence of

command hallucinations, and precautions should be

taken whenever there is any question about a

patient’s intent to commit suicide, since suicidal

ideation is the best predictor of a subsequent

attempt in schizophrenia (Meltzer et al. 2003).

Similar evaluations are necessary when considering

WFSBP Guidelines for Biological Treatment of Schizophrenia 137

the likelihood that the individual will harm someone

else or engage in other forms of violence (Buckley

et al. 2003).

Economic outcome, health service and stigma

Schizophrenia presents an enormous burden to the

patients and their relatives and reduces the quality of

life, especially due to chronic impairment. Schizo-

phrenia has been estimated to be one of 30 leading

courses of disability worldwide (Murray and Lopez

1997). Four to 15% of people suffering from

schizophrenia commit suicide, and the excess mor-

tality among people with schizophrenia is approxi-

mately 50% above that of the general population

(Brown et al. 2000). The cumulative cost of the care

of adults with schizophrenia is high: about 5�/6% of

National Health Service inpatient costs in England

were attributable to schizophrenia (Knapp 1997).

Schizophrenia has been estimated to account for

2.5% of annual health care expenditures in the

United States of America (Rupp and Keith 1993).

Over 65% of patients with schizophrenia after a

second psychotic episode are unemployed, thus

contributing to the high degree of so-called indirect

costs of the disease (Guest and Cookson 1999).

People with schizophrenia experience stigmatisation,

and one of the major goals of the activities of the

schizophrenia networks and health organisations is

to support programs against stigma (Gaebel and

Baumann 2003). It is of great importance to develop

successful treatment strategies, including multidi-

mensional approaches, in order to reduce the burden

of the disease.

Goal and target audience of WFSBP guidelines

These guidelines are intended for use in clinical

practice by all physicians investigating, diagnosing

and treating patients with schizophrenia. Therefore

an update of contemporary knowledge of various

aspects of schizophrenia, especially treatment op-

tions, is provided. The aim of these guidelines is to

improve standards of care, diminish unacceptable

variations in the provision and quality of care, and

support physicians in clinical decisions. Although

these guidelines favour particular treatments on the

basis of the available evidence, the treating physician

remains responsible for his assessment and treat-

ment option. These guidelines are primarily con-

cerned with the biological (somatic) treatment of

adults and address recommendations in this field.

The specific aim of these guidelines is to evaluate

the role of pharmacological agents in the treat-

ment and management of schizophrenia, while the

role of specific psychological interventions and

specific service delivery systems is covered only

briefly. The effectiveness of somatic treatment is

considered.

The guidelines were developed by the authors and

arrived at by consensus with the WFSBP Task Force

on Schizophrenia, consisting of 37 international

experts in the field.

Methods of literature research and data extraction

In the development of these guidelines the following

guidelines, consensus reports and sources were

considered:

American Psychiatric Association. Practice guide-

line for the treatment of patients with schizophre-

nia (APA 1997) and American Psychiatric

Association. Practice guideline for the treatment

of patients with schizophrenia, 2nd ed. (APA

2004).

Deutsche Gesellschaft fur Psychiatrie, Psy-

chotherapie und Nervenheilkunde. Praxisleitlinien

Psychiatrie und Psychotherapie: Schizophrenie

(DGPPN 1998); Guidelines for Neuroleptic Re-

lapse Prevention in Schizophrenia (Kissling

1991).

National Institute for Clinical Excellence. Core

Interventions in the Treatment of Schizophrenia

London (NICE 2003) and National Institute for

Clinical Excellence. Guidance on the use of newer

(atypical) antipsychotic drugs for the treatment of

schizophrenia (NICE 2002).

Royal Australian and New Zealand College of

Psychiatrists. Australian and New Zealand clinical

practice guideline for the treatment of schizophre-

nia, Draft only (RANZCP 2003), and Summary

Australian and New Zealand clinical practice

guideline for the treatment of schizophrenia

(McGorry et al. 2003).

Scottish Intercollegiate Guidelines Network. Psy-

chosocial Interventions in the Management of

Schizophrenia (SIGN 1998).

Task Force of the World Psychiatric Association.

The Usefulness and Use of Second-Generation

Antipsychotic Medications�/an Update (Sartorius

et al. 2002).

The Expert Consensus Guideline Series. Optimiz-

ing Pharmacologic Treatment of Psychotic Dis-

orders (Kane et al. 2003).

138 P. Falkai et al.

The Mount Sinai conference on the pharma-

cotherapy of schizophrenia (Marder et al. 2002).

The Texas Medication Algorithm Project

(TMAP) schizophrenia algorithms (Miller et al.

1999).

Translating research into practice: the Schizo-

phrenia Patient Outcomes Research Team

(PORT) treatment recommendations (Lehman

et al. 1998).

World Health Organization. WHO Guide to

Mental Health in Primary Care (WHO 2000).

The Cochrane Library, Meta-analyses on the

efficacy of different drugs and interventions used

in schizophrenia (2004 Issues).

Reviews, meta-analyses and randomised clinical

trials contributing to interventions in schizophre-

nic patients identified by search in the Medline

data base (up to February 2004) and individual

clinical experience by the authors and the mem-

bers of the WFSBP Task Force on Schizophrenia.

Evidence-based classifications of recommendations

The evidence found in the literature research and

data extraction was summarised and categorised to

reflect its susceptibility to bias (Shekelle 1999).

Daily treatment costs were not taken into considera-

tion due to the variability of medication costs

worldwide. Each treatment recommendation was

evaluated and discussed with respect to the strength

of evidence for its efficacy, safety, tolerability and

feasibility. It has to be kept in mind that the strength

of recommendation is due to the level of efficacy and

not necessarily of its importance. Four categories

were used to determine the hierarchy of recommen-

dations (related to the described level of evidence):

Level A. There is good research-based evidence to

support this recommendation. The evidence was

obtained from at least three moderately large,

positive, randomised controlled (double-blind) trials

(RCTs). In addition, at least one of these three

studies must be a well-conducted, placebo-con-

trolled study.

Level B. There is fair research-based evidence to

support this recommendation. The evidence was

obtained from at least two moderately large, positive,

randomised, double-blind trials (this can be either

two or more comparator studies or one comparator-

controlled and one placebo-controlled study) or

from one moderately large, positive, randomised,

double-blind study (comparator-controlled or pla-

cebo-controlled) and at least one prospective mod-

erately large (sample size equal to or greater than 50

participants), open-label, naturalistic study.

Level C. There is minimal research-based evidence

to support this recommendation. The evidence was

obtained from at least one randomised, double-blind

study with a comparator treatment and one pro-

spective, open-label study/case series (with a sample

size equal to or greater than 10 participants) showed

efficacy, or at least two prospective, open-label

study/case series (with a sample size equal to or

greater than 10 participants) showed efficacy.

Level D. Evidence was obtained from expert opi-

nions (from authors and members of the WFSBP

Task Force on Schizophrenia) supported by at least

one prospective, open-label study/case series (sam-

ple size equal to or greater than 10 participants).

No level of evidence or Good Clinical Practice (GCP).

This category includes expert opinion-based

statements for general treatment procedures and

principles.

Indication and goals of treatment for schizophrenia

Schizophrenia is a heterogeneous condition that has

a varying course and outcome and affects many

aspects of the patient’s life. The care of most patients

with this disorder involves multiple efforts and a

multidisciplinary team approach to reduce the

frequency, duration and severity of episodes, reduce

the overall morbidity and mortality of the disorder,

and improve psychosocial functioning, indepen-

dence and quality of life. The treatment of schizo-

phrenia requires the full understanding of the

patient, his or her needs and goals, conflicts and

defences, coping mechanisms and resources. There-

fore all those involved in treatment need to under-

stand the biological, interpersonal, social and

cultural factors which exert an influence on the

patient’s recovery. The major goals of treatment

should be the participation of the patient and all

those involved in the treatment process, co-opera-

tion with relatives, co-ordination and co-operation of

the treatment institutions in terms of integrated care

and the inclusion of the non-professional help and

self-help systems. Facilitation of access to the

healthcare system for patients suffering from schizo-

phrenia is an additional important goal of profes-

sional work. After the initial assessment of the

patient’s diagnosis and clinical and psychosocial

circumstances, a treatment plan must be formulated

WFSBP Guidelines for Biological Treatment of Schizophrenia 139

and implemented. This formulation involves the

selection of the treatment modalities, the specific

type(s) of treatment, and the treatment setting. The

goals and strategies of treatment vary according to

the phase and severity of illness.

In the acute phase of treatment (lasting weeks to

months), defined by an acute psychotic episode,

major goals are to develop an alliance with the

patient and family, to prevent harm, control dis-

turbed behaviour, reduce the severity of psychosis

and associated symptoms (e.g., agitation, aggression,

negative symptoms, affective symptoms), determine

and address the factors that led to the occurrence of

the acute episode and effect a rapid return to the best

level of functioning. A psychiatrist may see a patient

who is in an acute psychotic state that renders the

patient dangerous to self and others. Under these

circumstances, it may be impossible to perform an

adequate evaluation at the time of the initial evalua-

tion. Given the relative safety of most antipsychotic

medications, the psychiatrist may begin treatment

with an appropriate medication, even in states where

involuntary use of medication must be approved by a

court, and perform the necessary evaluations as they

become possible.

During the stabilisation phase (usually lasting 3�/6

months), the goals of treatment are to reduce stress

on the patient and provide support to minimise the

likelihood of relapse, enhance the patient’s adapta-

tion to life in the community, facilitate continued

reduction in symptoms and consolidation of remis-

sion, and promote the process of recovery. If the

patient has improved with a particular medication

regimen, it is recommended to continue that regi-

men for at least 6 months (APA 2004). It is also

critical to assess continuing side effects that may

have been present in the acute phase and to adjust

pharmacotherapy accordingly to minimise adverse

side effects that may otherwise lead to medication

nonadherence and relapse.

The goals of treatment during the stable phase

(lasting months to years) are to ensure that symptom

remission or control is sustained, that the patient is

maintaining or improving his or her level of func-

tioning and quality of life, that increases in symp-

toms or relapses are effectively treated, and that

monitoring for adverse treatment effects continues.

For most persons with schizophrenia in the stable

phase, psychosocial interventions are recommended

as a useful adjunctive treatment to pharmacological

treatment to improve outcome. The main aim of

pharmacological intervention in the stable phase is

to prevent relapse and help keep a person stable

enough to live as normal a life as possible (main-

tenance therapy and relapse prevention).

Acute-phase treatment of schizophrenia

In the acute phase, the specific treatment goals are to

prevent harm, control disturbed behaviour, suppress

symptoms, effect a rapid return to the best level of

functioning, develop an alliance with the patient and

family, formulate short- and long-term treatment

plans, and connect the patient with appropriate

aftercare in the community (APA 2004). Whatever

treatments are offered, it is essential to engage the

patient in a collaborative, trusting and caring work-

ing relationship at the earliest opportunity (NICE

2002). Psychosocial interventions in this phase aim

at reducing overstimulating or stressful relationships,

environments, or life events and at promoting

relaxation or reduced arousal through simple, clear,

coherent communications and expectations, a struc-

tured and predictable environment, low perfor-

mance requirements, and tolerant, nondemanding,

supportive relationships with the psychiatrist and

other members of the treatment team (DGPPN

1998; APA 2004). The patient should be provided

with information on the nature and management of

the illness that is appropriate to his or her ability to

assimilate the information. A patient has to be

informed about the benefits and side effects of the

medication (NICE 2002). The psychiatrist must

realise that the degree of acceptance of medication

and information about it varies according to the

patient’s cognitive capacity, the degree of the

patient’s denial of the illness, and efforts made by

the psychiatrist to engage the patient and family in a

collaborative treatment relationship (APA 2004).

Indications for hospitalisation include the patient’s

being considered to pose a serious threat of harm to

self or others, being unable to care for self, needing

constant supervision and general medical or psy-

chiatric problems that make outpatient treatment

unsafe or ineffective. Involuntary hospitalisations are

indicated if patients refuse to be admitted, and if

they meet the requirements of the local jurisdiction.

Alternative treatment settings, such as partial hospi-

talisation, home care, family crisis therapy, crisis

residential care, and assertive community treatment,

should be considered for patients who do not need

formal hospitalisation for their acute episodes (APA

2004). In the acute treatment phase, the main

emphasis is on pharmacotherapeutic (and other

somatic) interventions. Therefore antipsychotic

therapy should be initiated as early as possible as a

necessary part of a comprehensive package of care

that addresses the individual’s clinical, emotional

and social needs. The clinician responsible for

treatment and key worker should monitor both

therapeutic progress and tolerability of the drug on

an ongoing basis. Monitoring is particularly impor-

140 P. Falkai et al.

tant when individuals have just changed from one

antipsychotic to another (NICE 2002).

Antipsychotics

Antipsychotic medications have formed the basis of

schizophrenia treatment for approximately 50 years.

In terms of their chemical structure, the antipsycho-

tic medications, frequently known as neuroleptic

agents , are a heterogeneous group of psychoactive

drugs (such as phenothiazines, thioxanthenes, bu-

tyrophenones, diphenylbutylpiperidines, benza-

mides, benzisoxazoles and dibenzepines). They are

used in the acute phase treatment, long-term main-

tenance therapy and in the prevention of relapse of

schizophrenia.

Classification and efficacy

Conventional or first-generation antipsychotics (FGA)

can be classified into high- and low-potency medica-

tions. The effective dose of a first-generation anti-

psychotic medication is closely related to its affinity

for dopamine receptors (particularly D2) and its

tendency to cause extrapyramidal side effects

(Creese et al. 1976). High-potency antipsychotics

have a greater affinity for D2 receptors than low-

potency medications and the effective dose required to

treat psychotic symptoms like delusions and hallu-

cinations is much lower than for low-potency anti-

pychotics. This dose relationship can be expressed in

terms of dose equivalence (e.g., 100 mg of low-

potency antipsychotic chlorpromazine has an anti-

psychotic effect that is similar to that of 2 mg of the

high-potency antipsychotic haloperidol). Dose

equivalence does not equate with equivalence of

tolerability and should be considered as a general

concept rather than a precise clinical guide. The side

effects of high-potency agents, e.g., extrapyramidal

symptoms are easier to manage than the sedation

and orthostatic hypotension associated with low-

potency agents. For this reason, doses of low-

potency medications for a sufficient antipsychotic

effect often are difficult to reach.

With the detection of clozapine as an effective

antipsychotic agent that does not induce catalepsy

and apomorphine antagonism, a new class of ‘atypi-

cal’ antipsychotics became established. Because

various antipsychotics can be found on a continuum

ranging from typical to atypical, the terms second-

generation antipsychotics (SGA) or new-generation

antipsychotics (NGA) to describe these new agents,

which induce considerably fewer extrapyramidal

symptoms (EPS) in a therapeutic dose range than

conventional neuroleptics, have been found to be

more suitable (Fleischhacker 2002).

First-generation antipychotics

Studies demonstrating the efficacy of FGAs (e.g.,

phenothiazines) in reducing psychotic symptoms in

acute schizophrenia were mostly carried out in the

1960s and 1970s, comparing one or more antipsy-

chotic medications with either placebo or a sedative

agent. Nearly all of these studies found that the

antipsychotic medication was superior for treating

schizophrenia (NIH Psychopharmacology Service

Center Collaborative Study Group 1964). A review

of more than 100 trials concluded that the vast

majority of double-blind studies found superior

efficacy for FGAs compared to placebo and, with

the exception of mepazine and promazine, all of

these agents were equally effective, although there

were differences in dose, potency and side effects of

the different drugs (Davis et al. 1989). Superiority

of FGAs over placebo was confirmed by other

reviews evaluating randomised, double-blind studies

(Baldessarini et al. 1990; Kane and Marder 1993;

Dixon et al. 1995). Cochrane reviews and NICE

reviews, including high-quality RCTs, have found

the efficacy of some FGAs, e.g., chlorpromazine

(Thornley et al. 2004), flupenthixol (Centre for

Outcomes Research and Effectiveness, Systematic

Review Flupenthixol 2002), fluphenazine (Centre

for Outcomes Research and Effectiveness, Systema-

tic Review Fluphenazine 2002), perazine (Leucht

and Hartung 2004b), perphenazine (Centre for

Outcomes Research and Effectiveness, Systematic

Review Perphenazine 2002), pimozide (Sultana and

McMonagle 2004), sulpiride (Soares et al. 2004),

thioridazine (Sultana et al. 2004), trifluoperazine

(Marques et al. 2004) and zuclopenthixolacetate

(Fenton et al. 2004), but not of all FGAs,

e.g., benperidol (Leucht and Hartung 2004a), to

be similar to other conventional antipsychotics, and

superior compared to placebo for, e.g., chlorproma-

zine, thioridazine, trifluoperazine, and with only

slight superiority pimozide and sulpiride. Further-

more, FGAs (e.g., phenothiazines in this review)

have proven efficacy in diminishing psychotic symp-

toms in long-term treatment and reducing relapse

rates compared to placebo (Davis 1975).

High-potency first-generation antipsychotic agents are

associated with a high risk of extrapyramidal side

effects, a moderate risk of sedation, a low risk of

orthostatic hypotension and tachycardia, and a low

risk of anticholinergic and antiadrenergic effects.

A review of the clinical effects of haloperidol for

the management of schizophrenia, given as an

example for a high-potency FGA, demonstrated

efficacy in short- and longer-term periods compared

to placebo, but also a high propensity to cause

movement disorders (Joy et al. 2004) (Level A).

WFSBP Guidelines for Biological Treatment of Schizophrenia 141

Low-potency first-generation antipsychotic agents are

associated with a lower risk of extrapyramidal

effects, a high risk of sedation, a high risk of

orthostatic hypotension and tachycardia, and a

high risk of anticholinergic and antiadrenergic

effects. Especially because of sedation and ortho-

static hypotension, the dose should be increased

gradually. A review of the efficacy of chlorpromazine

compared to placebo as an example of a low-potency

FGA supported the view that there is slight superior

improvement under chlorpromazine therapy, focus-

ing on continuous outcome data, but more side

effects like sedation (Thornley et al. 2004) (Level A).

Second-generation antipsychotics

Although the efficacy of FGA in reducing positive

symptoms and relapse was impressive the problems

of only partially responding negative symptoms,

cognitive dysfunction, reduced quality of life and

functional impairment remained unsolved. A major

shortcoming of FGA’s contribution to the men-

tioned problems was that they produce significant

adverse effects, most notably extrapyramidal effects

(EPS) and subjective dysphoria in many patients,

leading to reduced tolerability and adherence. These

facts stimulated the development of second-genera-

tion antipsychotics (SGAs), the most significant

advantage of which is the lower propensity for

EPS, especially tardive dyskinesia. In addition, the

SGAs demonstrated better efficacy in treating nega-

tive symptoms, cognitive disturbances, and depres-

sive symptoms, a clinical profile that is often

described as a broader spectrum of clinical efficacy

(Moller 2000a,b). SGAs carry some risks like dis-

turbances of glucose utilisation, lipid metabolism

and weight gain, which are already known from

some of the conventional antipsychotics, but which

may be even more pronounced in some SGAs. At

the time of development of this guideline the

following second-generation antipsychotics (SGAs)

are available either in most European countries or on

the Australian and US market: amisulpride , aripipra-

zole , clozapine , olanzapine , quetiapine , risperidone ,

ziprasidone and zotepine . The available administra-

tion forms of these SGAs differ. All these medica-

tions can be administered in oral forms as pill or

tablets. Amisulpride and risperidone are available as

oral concentrates, olanzapine and risperidone in

quick-dissolving tablets, olanzapine and ziprasidone

as short-acting intramuscular preparations, and

risperidone as a long-acting injectable preparation.

A meta-analysis evaluating the efficacy of SGAs

compared to placebo including three RCTs of

risperidone, two RCTs of olanzapine and four

RCTs of quetiapine, calculated the effect size for

the categorical response rate as about 0.67, and

greater than 0.82 for continuous measure in favour

of the SGAs, and concluded that there is statistical

superiority of a SGA in therapeutic dose compared

to placebo (Woods et al. 2001). Placebo response

rates varied from 8 to 58% across the trials, in part

explained by response definitions used in the studies.

In addition, the review suggested the superiority of

ziprasidone, not marketed at the time of the analysis,

compared to placebo, similar to the other mentioned

SGAs.

Comparing the efficacy of FGAs versus SGAs. There is

still an ongoing controversial debate as to whether

SGAs, as a group, are superior to FGAs in their

efficacy and effectiveness in the treatment of schizo-

phrenia. Recent meta-analyses reported the crucial

points in randomised, controlled studies (Sartorius

et al. 2002). In a systematic overview and meta-

regression analysis of randomised controlled trials,

substantial heterogeneity was observed in the study

results comparing SGAs to FGAs, which was

partially accounted for by the dose of the FGAs

used. When the dose was about 12 mg/day of

haloperidol (or equivalent), atypical antipsychotics

were found to have no benefits in terms of efficacy or

overall tolerability, but to cause fewer extrapyramidal

side effects (Geddes et al. 2000). In a meta-analysis

of randomised efficacy trials comparing SGAs and

FGAs, and comparing different SGAs, effect sizes of

clozapine, amisulpride, risperidone and olanzapine

were greater than those of FGAs, and the effect of

zotepine was marginally geater, while other SGAs

revealed no clear superiority (Davis et al. 2003). No

difference in efficacy was detected among amisul-

pride, risperidone and olanzapine when directly

compared to each other. No evidence was found

that the haloperidol dose (or all FGA comparators

converted to haloperidol-equivalent doses) affected

these results. In a review of studies evaluating

efficacy and tolerability of olanzapine, risperidone,

quetiapine and sertindole, superiority to placebo was

reported (Leucht et al. 1999). Quetiapine and

sertindole were found to be comparable to haloper-

idol, while olanzapine and risperidone showed

slightly superior efficacy in the treatment of global

schizophrenic symptoms. In addition, olanzapine

and risperidone were found to demonstrate slight

superiority in improvement of negative symptoms.

All SGAs were noted to be associated with less

frequent EPS measured as the use of antiparkinso-

nian medications compared to haloperidol. A meta-

analysis of all randomised controlled trials in which

SGAs had been compared with low-potency

(equivalent or less potent than chlorpromazine)

FGAs found that, as a group, SGAs were moderately

142 P. Falkai et al.

more efficacious than low-potency antipsychotics,

largely irrespective of the comparator doses used

(Leucht et al. 2003). Furthermore the observation

has been made that low-potency FGAs in doses

lower than 600 mg/day chlorpromazine (CPZ)

equivalents might not induce more EPS than

SGAs.

Antipsychotic monotherapy should be preferred

and the minimum effective dose should be used. The

optimal dose for each patient has to be found by

clinical judgment. Recommended acute phase treat-

ment dosages for commonly used antipsychotics are

given in Table III.

Pharmacokinetics and pharmacogenetics

First-generation antipsychotic medications can be ad-

ministered in oral forms, as intravenous applications,

as short-acting intramuscular preparations, or as

long-acting injectable preparations. Short-acting in-

tramuscular medications reach a peak concentration

30�/60 minutes after the medication is administered,

whereas oral medications reach a peak after 2�/3

hours (e.g., Dahl 1990). As a result, the calming

effect of the first-generation antipsychotics may

begin more quickly when the medication is adminis-

tered parenterally. However, this calming effect on

agitation is different from the antipsychotic effect,

which may require several days or weeks. Oral

concentrates are typically better and more rapidly

absorbed than pill preparations, and often approx-

imate intramuscular administration in their time to

peak serum concentrations. Phenothiazines are po-

tent inhibitors of cytochrome P450 enzyme CYP

2D6 and may therefore impair the elemination of

substrates for this isoform. Zuclopenthixol is meta-

bolised by CYP 2D6, and haloperidol by CYP 2D6,

CYP 1A4 and CYP 3A4 (Spina et al. 2003).

Second-generation antipsychotics show similar phar-

macokinetics to those of FGAs. SGAs are rapidly

and completely absorbed after oral administration

but often undergo extensive first-pass hepatic meta-

bolism (Burns 2001). Time to peak plasma concen-

trations ranges from 1 to 10 hours. Atypical agents

are highly lipophilic, highly protein-bound, and tend

to accumulate in the brain and other tissues. The

liver extensely metabolises SGAs, predominantly by

cytochrome P450 enzymes. SGAs are only weak in

vitro inhibitors of CYP isoenzymes at therapeutic

concentrations and therefore not expected to inter-

fere with elimination of coadministered medication

(Spina et al. 2003). Due to large interindividual

variations in biotransformation and presence of

active metabolites that are not readily measured,

there is often little correlation between dose, serum

concentration and clinical effects (Burns 2001).

A number of medication interactions can have

clinically important effects for patients treated with

FGAs and SGAs (APA 2004). Coadministration of

drugs that inhibit or interfere with cytochrome P450

enzymes (e.g., heterocyclic antidepressants, some b-

blockers, cimetidine, caffeine, erythromycin, fluvox-

amine, fluoxetine, paroxetine, sertraline) may lead to

a significant increase in plasma levels and increase

Table III. Recommended dosage (orally) of selected antipsychotics in acute phase treatment.

Antipsychotic

Starting dose

(mg/day) DI1

Target dose first-episode

(mg/day)

Target dose multi-episode

(mg/day)

Maximal dosage

(mg/day)2

SGA

Amisulpride 200 (1)�/2 100�/300 400�/800 1200

Aripiprazole (10)�/15 1 15�/(30) 15�/30 30

Clozapine3 25 2�/(4) 100�/250 200�/450 900

Olanzapine 5�/10 1 5�/15 5�/20 20

Quetiapine 50 2 300�/600 400�/750 750

Risperidone 2 1�/2 1�/4 3�/6�/(10) 16

Ziprasidone 40 2 40�/80 80�/160 160

Zotepine 25�/50 2�/(4) 50�/150 100�/250 450

FGA

Chlorpromazine 50�/150 2�/4 300�/500 300�/1000 1000

Fluphenazine 0.4�/10 2�/3 2.4�/10 10�/20 20�/(40)

Flupenthixol 2�/10 1�/3 2�/10 10�/20 60

Haloperidol 1�/10 (1)�/2 1�/4 3�/15 100

Perazine 50�/150 1�/2 100�/300 200�/600 1000

Perphenazine 4�/24 1�/3 6�/36 12�/42 56

Pimozide 1�/4 2 1�/4 2�/12 16

Zuclopenthixol 2�/50 1�/3 2�/10 25�/50 75

1DI (dose intervals): recommended distribution of the daily dose �/ once�/1, twice�/2 etc.2Maximal approved dosage in many countries. In clinical practice some SGAs are even dosed higher.3Clozapine is usually not introduced to first-episode patients.

WFSBP Guidelines for Biological Treatment of Schizophrenia 143

side effects like sedation, hypotension and EPS;

inducers of CYP enzymes (e.g., phenytoin, nicotine,

rifampicin) can significantly reduce antipsychotic

levels. Barbiturates and carbamazepine decrease

plasma levels through effects on cytochrome P450

enzymes and reduce antipsychotic effect. In parti-

cular, changes in smoking status may especially

affect clozapine levels.

Further information about the pharmacokinetic

properties of the SGAs (Prior and Baker 2003; Raggi

et al. 2004) are presented in Table IV.

Pharmagenetic aspects focus mainly on different

genotypes of cytochrome P450 enzymes responsible

for variations in drug metabolism. In particular,

CYP 2 D6 activity can vary up to 1000-fold in the

population due to 70 or more genetic variants that

confer a decreased, normal or increased enzyme

activity (Albers and Ozdemir 2004). In 5�/8% of

Caucasians and 1�/5% of African-Americans and

Asians, the activity of the CYP 2D6 enzyme is very

low or absent and they are therefore so-called ‘poor

metabolisers’ (APA 2004). For this reason, blood

levels of psychotropic agents metabolised by CYP

2D6 (e.g., risperidone) may increase and these

substances may cause undesirable side effects even

when administered in therapeutic doses. In contrast,

ultrarapid metabolisers (due to multiple copies of

the CYP 2D6 gene) have subtherapeutic drug

exposure and may be incorrectly labelled as non-

compliant or treatment resistant (Albers and Ozde-

mir 2004).

Side effects

Differences in the risk of specific side effects of

antipsychotic medications are often predictable

from the receptor binding profiles of the various

agents. Some side effects result from receptor-

mediated effects within the central nervous system

(e.g., extrapyramidal side effects, hyperprolactine-

mia, sedation) or outside the central nervous system

(e.g., constipation, hypotension), whereas other side

effects are of unclear pathophysiology (e.g., weight

gain, hyperglycemia). Shared side effects of FGAs

and SGAs include neurological effects (i.e. acute and

chronic extrapyramidal effects, neuroleptic malig-

nant syndrome), sedation, cardiovascular effects

(i.e., hypotension, tachycardia and conduction ab-

normalities), anticholinergic and antiadrenergic

effects, weight gain and glucose and lipid metabolic

abnormalities, and sexual dysfunction, all in a

class- and substance-specific (often dose-related)

frequency and intensity. Table V gives an overview

of the estimated frequency of some important side

effects of the SGAs and haloperidol. Table VI gives

recommendations for monitoring.

Neurological side effects

Extrapyramidal side effects. Extrapyramidal side ef-

fects can be divided into acute and chronic cate-

gories. Acute extrapyramidal side effects are signs

and symptoms that occur in the first days and weeks

of antipsychotic medication administration, are dose

dependent, and are reversible with medication dose

reduction or discontinuation (Goetz and Klawans

1981). Acute dystonia typically occurs with high-

potency FGAs and more frequently in young, male

patients, but has also been reported in association

with SGAs, e.g., risperidone (Rupniak et al. 1986;

Leucht et al. 1999). Antipsychotic-induced parkinson-

ism is thought to affect 10�/80% (depending on

population and doses) of patients undergoing ther-

apy with high-potency FGAs (Bollini et al. 1994).

Akathisia occurs with a mean frequency of approxi-

mately 20�/25% in patients undergoing treatment

with FGAs (Braude et al. 1983; Grebb 1995). Tardive

dyskinesia (TD) may persist after medication dis-

Table IV. Pharmacokinetics of selected antipsychotics.

Agent

Maximal plasma level

(in hours)

Elimination half-time

(in hours) CYP enzymes*

FGAs

Chlorpromazine 2�/4 30 1A2, 2D6, 3A4

Haloperidol 3�/6 14�/20 2D6, 3A4

SGAs

Amisulpride 1�/3 12�/20 ?

Aripiprazole 3�/5 75�/146 (94) 2D6, 3A4

Clozapine 1.5�/3.6 16�/23 1A2, 3A4, (2C19, 2D6)

Olanzapine 5�/8 21�/54 1A2, (2C19, 2D6)

Quetiapine 1.0�/1.8 6.8 3A4

Risperidone 0.8�/1.4 3.6 2D6, (3A4)

Ziprasidone 3.8�/5.2 3.2�/10 3A4, (1A2, 2D6)

Zotepine 2.8�/4.5 8�/16 (12) 3A4, (1A2, 2D6)

*Secondary involved enzymes in parentheses. (Burns 2001, modified).

144 P. Falkai et al.

continuation and occurs with an incidence of

4(�/8)% per treatment year with conventional anti-

psychotics (Glazer 2000). After 4 years of therapy

with high-potency FGAs, approximately 20% of

patients have tardive dyskinesia and the rate is higher

(up to 50%) in elderly patients (Fenton 2000; Glazer

2000; Jeste 2000). Risk factors are age, female

gender, the presence of drug-induced parkinsonian

symptoms, diabetes mellitus, affective disturbances

and higher dose and longer duration of antipsychotic

therapy (Morgenstern 1993). SGAs induce fewer

extrapyramidal symptoms (EPS) in a therapeutic

dose range than FGAs and show a significant

reduction in the risk of tardive dyskinesia compared

to FGAs (Leucht et al. 1999; Correll et al. 2004).

Studies provided evidence that clozapine- and prob-

ably quetiapine-induced EPS are not dose dependent

(Buchanan et al. 1995; Cheer and Wagstaff 2004).

Neuroleptic malignant syndrome. Neuroleptic malig-

nant syndrome (NMS) is characterised by dystonia,

rigidity, fever, autonomic instability such as tachy-

cardia, delirium, myoglobinuria and increased levels

of creatine kinase, leukocytes and hepatic enzymes.

The prevalence of NMS is uncertain; it probably

occurs in less than 1% of patients treated with FGAs

(Adityanjee et al. 1999) and is even more rare among

patients treated with SGAs (e.g., Caroff et al. 2000).

Risk factors for NMS include acute agitation, young

age, male gender, preexisting neurological disability,

physical illness, dehydration, rapid escalation of

antipsychotic dose, use of high-potency medications

and use of intramuscular preparations (Pelonero

et al. 1998).

Epileptic seizures. Epileptic seizures occur in an

average of 0.5�/0.9% of patients receiving antipsy-

Table V. Selected side effects of commonly used antipsychotics.

Antipsychotic medication

Side effect Haloperidol Amisulpride Clozapine Olanzapine Risperidone Quetiapine Ziprasidone Aripiprazole

Akathisia/ Parkinsonism �/�/�/ 0�/�/ 0 0�/�/ 0�/�/�/ 0�/�/ 0�/�/ �/

Tardive dyskinesia �/�/�/ (�/) 0 (�/) (�/) ? ? ?

Seizures �/ 0 �/�/ 0 0 0 0 (�/)

QT-prolongation �/ (�/) (�/) (�/) (�/) (�/) �/ 0 (?)

Glucose abnormalities (�/) (�/) �/�/�/ �/�/�/ �/�/ �/�/ 0 0

Lipid abnormalities (�/) (�/) �/�/�/ �/�/�/ �/�/ �/�/ 0 0

Constipation �/ �/�/ �/�/�/ �/�/ �/�/ �/ 0 0

Hypotension �/�/ 0 (�/) (�/) �/�/ �/�/ �/ �/

Agranulocytosis 0 0 �/ 0 0 0 0 0

Weight gain* �/ �/ �/�/�/ �/�/�/ �/�/ �/�/ 0�/�/ �/

Prolactin elevation �/�/�/ �/�/�/ 0 (�/) �/�/ (�/) (�/) 0

Galaktorrhoea �/�/ �/�/ 0 0 �/�/ 0 0 0

Dysmenorrhoea/Amenorrhoea �/�/ �/�/ 0 0 �/�/ (�/) 0 0

Sedation �/�/�/ 0�/(�/) �/�/�/ �/�/�/�/ �/ �/�/ 0�/(�/) 0

Malignant neuroleptic

syndrome

(�/) ? (�/) (�/) (�/) (�/) ? (�/)

Note: Frequency and severity of side effects refers to information obtained by drug companies, FDA, additional literature and other

guidelines (e.g., APA 2004).

0�/no risk; (�/)�/occasionally, may be no difference to placebo; �/�/mild (less 1%); �/�/�/sometimes (less 10%); �/�/�/�/frequently

(�/10%); ?�/no statement possible due to lacking data.

*Weight gain during 6�/10 weeks: �/�/low (0�/1.5 kg); �/�/�/medium (1.5�/3 kg); �/�/�/�/high (�/3 kg).

Table VI. Monitoring for patients on second generation antipsychotics*

Baseline 4 Weeks 8 Weeks 12 Weeks Quarterly Annually

Personal/family history x x

Weight (BMI) x x x x x x

Waist circumference x x

Blood pressure x x x x

Fasting plasma glucose x x x

Fasting lipid profile x x x

Blood cell count x x x x

ECG x

BMI, body mass index; ECG, electrocardiogram.

*More frequent assessments may be warranted based on clinical status.

Note: Assessments refer mainly to American Diabetes Association and American Psychiatric Association, Diabetes Care 2004;27:596�/601.

WFSBP Guidelines for Biological Treatment of Schizophrenia 145

chotic medications (Devinsky et al. 1991), whereby

clozapine is associated with the highest incidence

rate (ca. 3%) and cumulative risk (ca. 10%) after 4

years of treatment (e.g., Devinsky et al. 1991;

Buchanan et al. 1995). The dose and initial dose

increases of clozapine appear to represent risk

factors for epileptic seizures (Devinsky et al. 1991).

Dose-dependent epileptic seizures at relevant rates

(7�/17%) have also been described with zotepine

(Hori et al. 1992).

Cognitive side effects. Although antipsychotic medica-

tions can effectively improve cognitive functions in

schizophrenic patients, memory problems and cog-

nitive disorders represent possible side effects of

antipsychotic therapy, which are particularly asso-

ciated with the anticholinergic effect of antipsychotic

medications and the use of anticholinergic agents

such as biperiden. Drug-induced cognitive disorders

have been more frequently reported in treatment

with typical antipsychotic medications (Buchanan et

al. 1994; Keefe et al. 1999; Meltzer and McGurk

1999; Purdon et al. 2000; Harvey and Keefe 2001;

Bilder et al. 2002; Green et al. 2002; Velligan et al.

2002).

Sedation. Sedation is a common side effect of FGAs,

as well as several of the SGAs, and may be related to

antagonist effects of those drugs on histaminergic,

adrenergic and dopaminergic receptors (e.g., Kinon

and Lieberman 1996). Sedation occurs more fre-

quently with low-potency typical antipsychotic med-

ications and clozapine. Sedation is most pronounced

in the initial phases of treatment, since most patients

develop some tolerance to the sedating effects with

continued administration.

Obesity and weight gain

Individuals with schizophrenia are more likely to be

overweight or obese than the population at large

(Marder et al. 2004). Combined with other risk

factors like smoking, reduced physical activity,

diabetes and hyperlipidemia the risk for cardiovas-

cular morbidity and mortality is increased. Besides

life style factors, such as poor diet and lack of

exercise, treatment with FGAs and SGAs can

contribute to weight gain (Marder et al. 2004).

Weight gain is a multifactorial occurrence during

treatment with neuroleptic agents. According to a

meta-analysis, the mean weight gain associated with

atypical antipsychotic medications is highest with

clozapine and olanzapine, and lower with risperi-

done, whereas ziprasidone hardly affects body

weight (Allison et al. 1999). The side effect of

weight gain has to be given due consideration, as it

affects therapy adherence, somatic sequelae, mortal-

ity, stigmatisation and quality of life (Allison et al.

2003). Therefore clinicians should sensitise patients

and their caregivers to the health risk associated with

excess weight and should encourage patients to self-

monitor their weight. Body mass index (BMI) and

waist size can serve as useful risk indicators (Marder

et al. 2004).

Metabolic side effects

Diabetes. There is evidence that schizophrenia itself

is an independent risk factor for impaired glucose

tolerance, which is a known risk factor for develop-

ing type 2 diabetes, regardless of whether patients

receive antipsychotic medication (Bushe and Holt

2004; Ryan et al. 2003). The interactions between

schizophrenia and diabetes are likely to be multi-

factorial and include genetic and environmental

factors. Besides endocrine stress sytems like the

sympathetic�/adrenal�/medullary system and the

hypothalamic�/pituitary�/adrenal axis, life style fac-

tors such as poor diet, obesity and lack of exercise

are involved in the genesis of diabetes (Dinan 2004;

Marder et al. 2004). Pharmaco-epidemiological

studies revealed a higher rate of diabetes in patients

receiving atypical antipsychotics compared with

those not receiving antipsychotics or with those

receiving conventional agents (Haddad 2004). Stu-

dies attempting to establish whether the association

with diabetes varies between different atypical anti-

psychotics are inconclusive (Koro et al. 2002a;

Wirshing et al. 2002; Bushe and Leonard 2004;

Haddad 2004). Nevertheless clozapine and olanza-

pine were thought to be the agents most commonly

associated with diabetes (Marder et al. 2004). As a

consequence, a baseline measure of (fasting) plasma

glucose levels should be collected for all patients

before starting a new antipsychotic, alternatively

hemoglobin A1c should be measured (Marder et al.

2004). Patients and their caregivers should be

informed about the symptoms of diabetes and

patients should be monitored at regular intervals

for the presence of these symptoms (GCP).

Hyperlipidemia. Retrospective reports and pharma-

coepidemiological studies found a significantly

greater extent of elevations of lipids (trigylcerides)

in patients taking certain atypical antipsychotic

medications (olanzapine and clozapine) than in

patients receiving other antipsychotics (e.g., halo-

peridol, quetiapine, risperidone) (Koro et al. 2002b;

Wirshing et al. 2002). Similar to diabetes, hyperli-

pidemia is linked to a multifactorial genesis and is

associated with obesity and life-style factors like poor

diet and lack of exercise. Elevated triglyceride and

146 P. Falkai et al.

cholesterol levels are associated with coronary heart

disease, including ischaemic heart disease and myo-

cardial infarction. Therefore total cholesterol, low-

density lipoprotein (LDL) and HDL cholesterol,

and triglyceride levels should be measured (Marder

et al. 2004). If the LDL level is greater than 130 mg/

dl the patient should be referred to an internist to

evaluate whether treatment with a cholesterol-low-

ering drug should be initiated.

Hyperprolactinemia and sexual dysfunction

Antipsychotics�/particularly FGAs, amisulpride and

risperidone�/can cause hyperprolactinemia by block-

ing dopamine D2 receptors (Mota et al. 2003;

Marder et al. 2004). Hyperprolactinemia can lead

to galactorrhoea, menstrual, cyclical and sexual

disturbances in women, and reproductive and sexual

dysfunction and galactorrhoea/gynaecomastia in

men (Dickson and Glazer 1999). Plasma prolactin

concentrations may remain elevated for up to 2

weeks after the cessation of oral therapy with a

conventional antipsychotic, and for up to 6 months

after cessation of depot therapy (Cutler 2003).

Studies of SGAs (with the exception of amisulpride

and risperidone) suggested that the transiently

elevated prolactin levels tended to return to the

normal range within a few days (Marder et al. 2004).

Effects on dopaminergic, adrenergic, cholinergic

and serotonergic mechanisms may also lead to

sexual dysfunction, whereby it is difficult to distin-

guish between this and disease-related impairment

of sexual activity (Baldwin and Birtwistle 1997;

Fortier et al. 2000). Peripheral a-adrenergic block-

ade can be responsible for priapism (Cutler 2003).

Of the SGAs, clozapine, olanzapine and ziprasidone

were associated with only little or modest prolactin

elevation (Cutler 2003), and quetiapine did not lead

to hyperprolactinemia (Arvanitis et al. 1997). A clear

association between prolactin elevation and sexual

dysfunction has not been established (Aizenberg et

al. 1995; Kleinberg et al. 1999). There is still an

ongoing debate as to whether hyperprolactinemia

increases the risk of breast cancer, and the studies

are inconclusive to date (Marder et al. 2004).

Hyperprolactinemia is suspected to cause osteoporo-

sis if it impairs sex steroid production.

Cardiovascular side effects

Hypotension and orthostatic hypotension are related to

the a-antiadrenergic effects of antipsychotic medica-

tions, and are therefore particularly associated with

low-potency typical antipsychotic medications and

some SGAs, e.g., clozapine (Buchanan et al. 1995;

APA 2004). Patients who experience severe postural

hypotension must be cautioned against getting up

quickly and without assistance as falls can result in

hip fractures and other accidents, particularly in

elderly patients. Gradual dose titration, starting with

a low dose, and monitoring of orthostatic signs

minimises the risk of complications due to ortho-

static hypotension. Tachycardia is particularly rele-

vant in pre-existing cardiac disease. Tachycardia can

result from the anticholinergic effects of antipsycho-

tic medications, but may also occur as a result of

postural hypotension. Tachycardia unrelated to

orthostatic blood pressure changes that result from

anticholinergic effects may occur in up to 25% of

patients treated with clozapine (APA 2004). Cardiac

side effects of antipsychotic medications registered as

ECG-abnormalities include prolongation of the QT

interval, abnormal T-waves, prominent U-waves and

widening of the QRS complex (Haddad and Ander-

son 2002). The average QTc interval in healthy

adults is approximately 400 ms. QTc prolongation

(QTc intervals above 500 ms) is associated with an

increased risk of torsade de pointes and transition to

ventricular fibrillation (Glassman and Bigger 2001).

At varying rates, all antipsychotics may cause (dose-

dependent) cardiac side effects; of the FGAs, this

predominantly applies to tricyclic neuroleptic agents

of the phenothiazine type (e.g., chlorpromazine,

promethazine, perazine and, especially, thioridazine)

and to pimozide. In addition, high-dose intravenous

haloperidol has been associated with a risk of QTc

prolongation (Al Khatib et al. 2003). Of the SGAs,

sertindole and ziprasidone were found to lengthen

the QT interval in a significant manner (Glassman

and Bigger 2001; Marder et al. 2004). Prior to

commencement, and at subsequent increases of

dose, of antipsychotic therapy with thioridazine,

mesoridazine and pimozide, ECG monitoring is

obligatory. This recommendation also applies to

ziprasidone if cardiac risk factors, like known heart

disease, congenital long QT syndrome, personal

history of syncope and family history of sudden

death at an early age, are present. If these risk factors

are present, thioridazine, mesoridazine and pimo-

zide should not be given (Marder et al. 2004). The

application of two or more antipsychotics or other

adjunctive agents could increase the risk of cardiac

side effects (e.g., QTc prolongation). Case reports

indicate that the use of clozapine is associated with a

risk of myocarditis in 1 per 500 to 1 per 10,000

treated patients (Killian et al. 1999; Warner et al.

2000; La Grenade et al. 2001). If the diagnosis is

probable, clozapine should be stopped immediately

and the patient referred urgently to a specialist for

internal medicine (Marder et al. 2004).

WFSBP Guidelines for Biological Treatment of Schizophrenia 147

Other side effects

Hematological effects , like inhibition of leukopoiesis,

occur in patients being treated with chlorpromazine,

for example, as benign leukopenia in up to 10% and

as agranulocytosis in approximately 0.3% of patients

(APA 2004). The risk of agranulocytosis (defined as

an absolute neutrophil count less than 500/mm3) has

been estimated at 0.05�/2.0% of patients per year of

treatment with clozapine (Buchanan et al. 1995).

The risk is highest in the first 6 months of treatment,

and therefore weekly white blood cell (WBC) and

neutrophil monitoring is required. After 18 weeks,

the monitoring rate may be reduced to every 2�/4

weeks, as the risk of agranulocytosis appears to

diminish considerably (an estimated rate of three

cases per 1000 patients). WBC counts must remain

above 3000/mm3 during clozapine treatment, and

absolute neutrophil counts must remain above 1500/

mm3. With maintenance treatment, patients should

be advised to report any sign of infection immedi-

ately (e.g., sore throat, fever, weakness or lethargy)

(APA 2004).

Allergic and dermatological effects , including photo-

sensitivity, occur infrequently but are most common

with low-potency phenothiazine medications. Pa-

tients should be instructed to avoid excessive sun-

light and use sunscreen (APA 2004). Hepatic effects

like elevated hepatic enzymes may be triggered by a

number of antipsychotic medications, but this is

usually asymptomatic. Direct hepatotoxicity or cho-

lestatic jaundice occur extremely rarely and are

particularly associated with low-potency phenothia-

zines (APA 2004). In studies involving olanzapine,

reversible, mainly slight elevations in hepatic en-

zymes have been reported (Beasley et al. 1996a).

Ophthalmological effects due to pigment accumulation

in the lens and cornea, retinopathies, corneal oe-

dema, accommodation disturbances and glaucoma

have also been described as side effects of antipsy-

chotic medication. To prevent pigmentary retinopa-

thies, corneal opacities and cataracts, patients

maintained on thioridazine and chlorpromazine

should have periodic ophthalmological examinations

(approximately every 2 years for patients with a

cumulative treatment of more than 10 years), and a

maximum dose of 800 mg/day of thioridazine is

recommended (APA 2004). As cataracts were ob-

served in beagles that were given quetiapine, psy-

chiatrists should ask about the quality of distance

vision and about blurry vision, and should refer to an

ocular evaluation yearly or every 2 years (Marder et

al. 2004). Urinary tract problems , such as urinary

retention and urinary incontinence, may be particu-

larly provoked by antipsychotic medications with

marked anticholinergic components such as phe-

nothiazines and those with cholinergic effects. Dry

mouth and eyes and constipation may result from

adrenergic and anticholinergic stimulation, often

described during treatment with FGAs. Sialorrhoea

and drooling occur relatively frequently with cloza-

pine treatment and are most likely due to decreased

saliva clearance related to impaired swallowing

mechanisms, or possibly as a result of muscarinic

cholinergic antagonist activity at the M4 receptor or

to a-adrenergic agonist activity (Rabinowitz et al.

1996).

Adjunctive medications

Benzodiazepines. The efficacy of benzodiazepines in

schizophrenia has been evaluated as monotherapy

and as adjuncts to antipsychotic medications (APA

2004). In a review of double-blind studies of

benozidazepines as monotherapy, superior effects

compared to placebo (reductions in anxiety, agita-

tion, global impairment or psychotic symptoms)

were reported in most, but not all studies (Wolkowitz

and Pickar 1991). Seven of 16 double-blind studies

evaluating benzodiazepines as adjuncts to psychotic

medication showed positive effects on anxiety, agita-

tion, psychosis or global impairment; five of 13

demonstrated efficacy in treating psychotic symp-

toms (Wolkowitz and Pickar 1991). It was noted that

benzodiazepines may improve the response to anti-

psychotic medications, but this effect may be limited

to the acute phase and not be sustained (e.g.,

Altamura et al. 1987; Csernansky et al. 1988). The

efficacy in patients with prominent agitation was one

of the most consistent findings in retrospective and

open-label studies (e.g., Salzman et al. 1991; Wolk-

owitz and Pickar 1991) (Level A). Common side

effects of benzodiazepines include sedation, ataxia,

cognitive impairment and a tendency to cause

behavioural disinhibition in some patients.

Mood stabilisers and anticonvulsants. A recent meta-

analysis including 20 double-blind RCTs using

dichotomous response criteria (e.g., symptom re-

duction at least 50% of baseline level) found no

advantage of lithium as a sole agent compared to

placebo or to antipsychotics, but a slight superiority

of lithium versus placebo augmentation to antipsy-

chotics (focusing primarily on nonresponders)

(Leucht et al. 2004a). In addition, lithium as a sole

agent was inferior to antipsychotics if patients with

schizoaffective disorder were excluded from analysis.

Overall there is evidence for superior efficacy of

lithium augmentation to antipychotics, especially in

patients with mood symptoms (Level A) and treat-

ment-resistant schizoprenia (Level B). Patients

should be monitored for adverse effects that are

148 P. Falkai et al.

commonly associated with lithium (e.g., polyuria,

tremor) and with its interaction with an antipsycho-

tic medication (e.g., EPS, confusion, deterioration,

other signs of neuroleptic malignant syndrome),

including short-time evaluation of blood levels,

particularly during the initial period of combined

treatment (APA 2004).

A recent review identified five randomised, con-

trolled studies examining the efficacy of valproate as

an adjunct to antipsychotics in schizophrenia (Basan

et al. 2004). One of these trials was single-blind

(Hesslinger et al. 1999), the others double-blind

(Fisk and York 1987; Dose et al. 1998; Wassef et al.

2000; Casey et al. 2003a). The overall results of this

meta-analysis were inconclusive. In summary, there

is limited evidence that adding valproate to anti-

psychotic treatment may be successful in reducing

specific symptoms, e.g., hostility (Citrome et al.

2004), and therefore it may useful to introduce

valproate to special patient populations, but not

to use it in general when treating schizophrenia

(Level C).

A meta-analysis including 10 double-blind RCTs

found that carbamazepine , compared to placebo and

compared to antipsychotics as the sole treatment for

schizophrenia, revealed no beneficial effect (Leucht

et al. 2004b). Carbamazepine augmentation of

antipsychotics versus placebo was superior com-

pared with antipsychotics alone in terms of overall

improvement, but there were no differences for

mental state outcomes (Leucht et al. 2004b). There

is limited evidence that adding carbamazepine to

antipsychotic treatment may be successful in redu-

cing specific symptoms, e.g., aggressive behaviour or

affective symptoms (Luchins 1987; Okuma et al.

1989), but not for its general use in treating

schizophrenia (Level C).

In treatment-resistant schizophrenia, the addition

of lamotrigine (200 mg/day) to an ongoing clozapine

treatment was effective in reducing positive and

general psychopathological symptoms in one RCT

(Tiihonen et al. 2003); in another RCT, adding

lamotrigine (400 mg/day) to conventional antipsy-

chotics, risperidone, olanzapine or clozapine, im-

proved positive and general psychopathology when

analysing the completers, but no difference was seen

in the ‘primary last observation carried forward’

analysis (Kremer et al. 2004). There is evidence that

lamotrigine as an adjunctive treatment, especially to

clozapine, can reduce schizophrenic psychopathol-

ogy (Level C), keeping the potential elevated risk

of leucopenia and agranulocytosis in mind (Kossen

et al. 2001).

In an open, non-randomised trial no beneficial

effect of topiramate added to an ongoing treatment

with clozapine, olanzapine, risperidone or flu-

penthixol was seen in treatment-resistant schizo-

phrenia (Dursun and Deakin 2001).

Antidepressants. Studies of antidepressants in schizo-

phrenia focused mainly on their efficacy in treating

comorbid depression or negative symptoms (APA

2004). A meta-analysis suggested some evidence for

superior improvement of depressive symptoms com-

pared to placebo, while deterioration of psychotic

symptoms or worsening of adverse effects, especially

EPS, was not noted (Whitehead et al. 2004),

although this was described in some individual

studies. An earlier review of the use of antidepres-

sants in patients with schizoaffective disorder or

schizophrenia with mood symptoms came to a

positive conclusion and recommended treatment

with antidepressive agents if indicated (Levinson et

al. 1999). There is also some evidence for efficacy of

antidepressants in negative symptoms of schizophre-

nia (APA 2004; Moller 2004a) (Level B). Since most

of the studies were performed in combination with

first-generation antipsychotics, findings may be dif-

ferent with second-generation antipsychotics (APA

2004). Clomipramine (Berman et al. 1995) and

fluvoxamine (Reznik and Sirota 2000) showed

advantages in treating obsessive�/compulsive symp-

toms in schizophrenia, derived from two small open

studies (Level D). In a crossover study, citalopram

revealed efficacy in patients with a history of aggres-

sion in that it reduced the frequency of incidents

(Vartiainen et al. 1995) (Level D).

Others. A meta-analysis of five double-blind RCTs,

in which b-blockers were added to standard drug

treatment in schizophrenia, found no overall efficacy

for this strategy; nevertheless, several small open

studies and case reports provided benefits in the

outcome of treatment refractory patients (Cheine

et al. 2004).

Based on pathophysiological considerations and

experimental research the glutamatergic agents , gly-

cine , d-cycloserine , and d-serine have been studied as

additional treatments to antipsychotics in schizo-

phrenia, with controversial results (Tsai et al. 1998,

1999; Goff et al. 1999; Potkin et al. 1999; Evins et

al. 2000; Javitt et al. 2001; Moller 2003; APA 2004).

The mainly negative studies of glutamatergic sub-

stances as adjuncts to clozapine may be due to their

similar action on negative symptoms via the gluta-

matergic pathway.

Baclofen , a GABAB agonist, did not show any

relevant benefit in open and placebo-controlled

double-blind trials in monotherapy or when added

to standard antipychotics, following an initial suc-

cessful case report (Soares et al. 2004; Wassef et al.

2003b). Piracetam added to haloperidol at a dose

WFSBP Guidelines for Biological Treatment of Schizophrenia 149

of 3200 mg/day demonstrated beneficial effects in

a placebo-controlled randomised study (Noorbala

et al. 1999). Acetylcholinesterase inhibitors (e.g.,

donepezil) revealed, in case reports (Risch et al.

2001) and in an uncontrolled study, positive results

on a variety of cognitive measures (Buchanan et al.

2003), but not in a randomised, placebo-controlled

trial in patients with chronic schizophrenia

(Friedman et al. 2002). Oestrogen adjunction (0.05

mg/day) to ongoing haloperidol treatment showed

greater improvement in positive symptoms and

general psychopathology in women with chronic

schizophrenia (Kulkarni et al. 2001; Akhondzadeh

et al. 2003), but not in acute schizophrenia (0.625

mg/day conjugated oestrogen added to haloperidol)

(Louza et al. 2004). Augmentation with unsaturated

essential fatty acids (EFA) (e.g., eicosapentaenoic

acid) demonstrated controversial results in four

randomised placebo-controlled studies with EPA

(Emsley et al. 2003). A Cochrane review stated no

clear effects of v-3-fatty acids (Joy et al. 2004), but

supported the suggestion of further research in this

direction.

Electroconvulsive therapy (ECT)

Electroconvulsive therapy (ECT) was introduced in

1938. It involves the induction of seizures for

therapeutic purposes by the administration of a

variable frequency electrical stimulus to the brain

via electrodes applied to the scalp. The procedure is

usually modified by the use of short-acting anaes-

thetics and muscle relaxants (Tharyan and Adams

2004). In clinical practice the following are recom-

mended: generalised motor seizures of 25�/30 sec-

onds duration, monitoring via EEG and by

observing motor convulsions in a forearm isolated

from muscle relaxants, a stimulus intensity approxi-

mately 2.5-fold the seizure threshold, treatment

frequencies of twice or three times weekly and of

between 12�/20 treatments (Tharyan and Adams

2004). The evaluation preceding ECT should con-

sist of a general medical evaluation to identify risk

factors (including history and physical examination,

assessment of vital signs, basic laboratory tests and

ECG), anaesthesia evaluation addressing the nature

and extent of anaesthetic risk and obtaining in-

formed consent. Although there are no absolute

contraindications to ECT, recent myocardial infarc-

tion, cardiac arrhythmias or pacemakers, congestive

heart failure or severe coronary heart disease,

abdominal and intracranial aneurysm, and intracra-

nial space-occupying lesions are relative contraindi-

cations and require extreme caution and

consultation of a specialist in internal medicine

(APA Task Force 2001). During the informed

consent process, information about the procedure,

including anaesthesia, should be given and the

potential risks and benefits of ECT and alternative

therapeutic approaches should be described. In

terms of electrode placement, two RCTs comparing

bitemporal to unilateral nondominant hemisphere

electrode placements in patients with schizophrenia

did not show significant differences regarding mental

state, global improvement or cognitive functioning

(Thayran and Adams 2004). Although findings in

patients with depression suggest that unilateral and

perhaps bifrontal electrode placement may be asso-

ciated with fewer cognitive effects, and that efficacy

with unilateral electrode placement may depend on

the extent to which the stimulus intensity exceeds

the seizure threshold, the applicability of these

observations to patients with schizophrenia is un-

certain (APA 2004). An RCT assessing three differ-

ent stimulus intensities with bitemporal ECT found

that rates of remission and effects on cognition were

comparable, but the low-dose remitter group re-

ceived more ECT treatments and required more

days to meet remitter status than the 2- and 4-fold

seizure threshold groups (Chanpattana et al. 2001).

Clinical case series, primarily from the older litera-

ture, suggest that achieving full clinical benefit for

patients with schizophrenia may require a longer

course of acute treatment than for patients with

mood disorders (e.g., Kalinowsky 1943), and one

RCT showed advantage for 20 treatments over 12

treatments (Baker et al. 1960).

The efficacy of acute treatment with ECT in

patients with schizophrenia has been described in

case series, uncontrolled studies and controlled

trials. Therefore, publication details of reviews and

HTA reports have been provided (Hawkins et al.

1995; Fink and Sackeim 1996; Rey and Walter

1997; Walter et al. 1999; APA Task Force 2001;

Greenhalgh (UK ECT Group) et al. 2002; Tharyan

and Adams 2004). One result of the cited reviews is

that antipsychotic treatment alone generally pro-

duced better short-term outcomes compared with

ECT alone (Level A). There is also evidence from at

least three studies that ECT leads to a significant

better global impression compared to sham (pla-

cebo) treatment (Thayran and Adams 2004) (Level

A). Nevertheless, there are different opinions, and

other reviewers did not find a significant advantage

for ECT compared with sham treatment regarding

mental state (APA 2004). Combined treatment with

ECT and first-generation antipsychotic medications

(FGAs) was observed to be more effective than

treatment with ECTalone in most but not all studies

(APA 2004).

A review on efficacy of ECT in adolescents

identified no controlled trials and showed improve-

150 P. Falkai et al.

ment rates of 63% for depression, 80% for mania,

42% for schizophrenia and 80% for catatonia. In

conclusion, ECT in the young seems similar in

effectiveness and side effects to ECT in adults,

bearing the lack of systematic evidence in mind

(Rey and Walter 1997).

Side effects of ECT on the cardiovascular system

are common but are typically benign and self-limited

(APA Task Force 2001). Seldomly ECT may be

associated with more serious cardiac arrhythmias,

ischaemia and infarction, although the type, severity

and likelihood of cardiac complications are generally

related to the type and severity of preexisting cardiac

disease (e.g., Rice et al. 1994). ECT treatment and

its anaesthesia are often associated with a transient

postictal confusional state, at times accompanied by

postictal agitation. Cognitive side effects may also be

observed with ECT, although there is much indivi-

dual variation in the extent and severity of such

effects (APA Task Force 2001). For most patients,

however, the retrograde memory impairment typi-

cally resolves in a few weeks to months after

cessation of treatment (APA 2004). Rarely do

patients report more pervasive or persistent cognitive

disruption that involves more distant memories (e.g.,

Donahue 2000). A comprehensive review concluded

that there is no credible evidence that ECT causes

structural brain damage (Devanand et al. 1994).

ECT more often leads to improvements in concen-

tration and attention (and in consequence memory

function), parallel with clinical improvement (Prudic

et al. 2000). Other side effects that are commonly

noted after ECT include headache, generalised

muscle aches, and nausea and/or vomiting. These

effects usually resolve spontaneously or with analge-

sic or antiemetic medications (Datto 2000).

In catatonic schizophrenia, ECT is often the

treatment of choice.

In summary, apart from in catatonia, electrocon-

vulsive therapy (ECT) should only be used in

exceptional cases in treatment-refractory schizo-

phrenia, as no advantages have been consistently

demonstrated compared with pharmacological treat-

ments (Level C). Most studies of ECT did not

conduct a comparison to monotherapy with atypical

agents as an alternative. ECT should be considered

in catatonia (or severe affective symptoms), as there

is limited evidence in trials and clinical knowledge to

confirming its efficacy in such cases, and alternative

options are rare (Level C).

Repetitive transcanial magnetic stimulation (rTMS)

Repetitive transcranial magnetic stimulation (rTMS)

is not an approved treatment for neuropsychiatric

disorders to date, nevertheless this novel somatic

technique has been studied in many neuropsychia-

tric diseases (Burt et al. 2002). In addition, there are

a few controlled trials for ameliorating psychotic

symptoms in schizophrenia. Repetitive TMS stimu-

lates cortical neurons noninvasively by magnetic

induction, using a brief, high-intensity magnetic

field. Advantages of this technique compared to

ECT are better tolerability, fewer side effects and no

need for anaesthesia. Improvement in auditory

hallucinations after stimulation of the left

temporal�/parietal cortex augmenting antipsychotic

treatment was observed in a randomised, double-

blind, sham-controlled trial and in two randomised,

double-blind, crossover studies, each with small

sample sizes (Hoffman et al. 2000, 2003; Rollnik

et al. 2000). Another small, randomised, controlled

study showed no effect of rTMS stimulating the

right dorsolateral prefrontal cortex (Klein et al.

1999). Further research activities are required to

evaluate efficacy and potential benefits of rTMS in

schizophrenia.

Psychotherapy and psychosocial interventions

As mentioned earlier, these guidelines focus on

biological (somatic) treatments of schizophrenia,

therefore psychotherapeutic approaches alone or in

combination with pharmacotherapy will only be

described briefly in these guidelines. The aims of

psychological treatment methods in schizophrenic

diseases are to reduce the individual vulnerability,

alleviate the adverse influence of external stressors,

improve the quality of life, reduce the disease

symptoms and promote and improve the patient’s

communication skills and ability to cope with the

disease. Psychotherapy has to pay attention to the

biological factors involved in schizophrenia and must

be aimed at enabling the patient to cope with the

disease and its consequences (acceptance of relapses,

self-management, coping with problems). Psy-

chotherapeutic and psychosocial interventions be-

come more important in the long-term treatment of

schizophrenia and are mostly restricted to counsel-

ling/supportive psychotherapy, psychoeducation and

family interventions or cognitive behavioural therapy

in the acute phase treatment. For this reason,

evidence-based evaluation referring to main topics

regarding guidelines, meta-analysis and systematic

reviews will be presented in the long-term treatment

part of these guidelines.

First-episode schizophrenia

A patient with a first episode of symptoms char-

acterizing schizophrenia may be experiencing a

schizophrenic episode or the onset of schizophreni-

WFSBP Guidelines for Biological Treatment of Schizophrenia 151

form disorder, or may be having an episode of

another illness or disorder that can cause similar

symptoms. Therefore careful initial evaluation and

assessment has to be conducted as described above.

Observation of a patient, even for a short while, can

provide clues to the nature of the diagnosis. In

naturalistic and controlled studies, patients with

first-episode psychosis were more treatment respon-

sive than patients with multiple episodes of psychosis

but, at the same time, quite sensitive to side effects,

especially EPS (e.g., Lieberman et al. 1996). In-

cidence of tardive dyskinesia with low-dose haloper-

idol was seen at least as often as in other samples

treated with standard doses of FGAs, and additional

clinical features could not identify subjects at risk for

tardive dyskinesia (Oosthuizen et al. 2003). Predic-

tors of reduced treatment response in one study were

male sex, obstetric complications in birth history,

more severe hallucinations and delusions, attention

deficits and development of parkinsonism during

antipsychotic treatment (Robinson et al. 1999). In

other studies, depressive symptoms at baseline pre-

dicted fewer negative symptoms in the course and

better outcome (Oosthuizen et al. 2002), and

dystonia was associated with younger age, prominent

negative symptoms and higher disease severity

(Aguilar et al. 1994). Earlier antipsychotic treatment

(or shorter duration of untreated psychosis) was

associated with better outcomes in first-episode

schizophrenia, whereby poor premorbid function

could indicate an illness subtype less likely to

respond to antipsychotic treatment regardless of

when it is initiated (Perkins et al. 2004). Cannabis

use, common in first-episode schizophrenics, was

noted to cause confusion and delay in treating the

psychotic episode.

Efficacy of first-generation antipsychotics. In a 5-week

RCT comparing flupenthixol (mean dose 20 mg/

day) to pimozide (mean dose 18.8 mg/day), compar-

able efficacy in improvement of positive symptoms

was noted, while response of negative symptoms

varied. Pimozide produced a higher elevation of

prolactin levels (The Scottish Schizophrenia Re-

search Group 1987). A randomised trial found that

a higher dosage of fluphenazine (20 mg/day) re-

vealed better global improvement but more disabling

side effects than lower doses (5 and 10 mg/day) (Van

Putten et al. 1991). A review evaluating pharma-

cotherapy of first episode schizophrenia reported

that in open trials with subsequent increase of

dosage in case of nonresponse, haloperidol led to

the best improvement (72% of cases) at doses

between 2 and 5 mg daily compared to higher

dosages (10�/20 mg/day) (Remington et al. 1998).

Open, non-comparative 6-week trials of fluphena-

zine (20 and 40 mg/day) and haloperidol (20�/40

mg/day) found that approximately 70% of first-

episode patients were stabilised within this period

(Remington et al. 1998). In a randomised trial,

higher doses of haloperidol (20 mg/day) demon-

strated only intermittently (shorter than 2 weeks)

better efficacy on psychopathology compared to

lower dosages (5 and 10 mg/day), before outcome

worsened due to EPS (Van Putten et al. 1990). More

recent open (Oosthuizen et al. 2001) and rando-

mised, controlled (Oosthuizen et al. 2004) studies

showed low doses of haloperidol (2 mg/day or less)

to be at least as effective and better tolerated (lower

EPS) than higher doses of haloperidol.

Efficacy of second-generation antipychotics. Only a few

randomised controlled studies are available that

compare the efficacy of SGAs with that of

FGAs. Risperidone compared to haloperidol (mean

6 mg/day) in flexible dosing showed comparable

improvement in psychotic symptoms and less EPS in

a 6-week trial, demonstrating that already low

dosages (B/6 mg/day) may be successful in control-

ling symptoms (Emsley et al. 1999). In another RCT

with two different doses of risperidone (2 vs. 4 mg/

day), patients with the higher dosage demonstrated

increased impairment in motor fine-tuning tasks but

no better improvement in psychotic symptoms

(Merlo et al. 2002). In a non-comparative open trial

in first-psychotic-episode patients, low-dose risper-

idone (lower than 6 mg/day) was concluded to be

effective and well tolerated, and significant improve-

ments could be maintained over 1 year of treatment

(Huq et al. 2004). These results underline conclu-

sions drawn by other open studies (Kopala et al.

1997; Yap et al. 2001). A small, randomised, open

trial of low-dose risperidone compared to low-dose

zuclopenthixol revealed no evidence for differential

effects on psychopathology or cognitive function, if

corrected for the covariates EPS and anticholinergic

medication (Fagerlund et al. 2004). Low-dose

zuclopenthixol did, however, cause significantly

more EPS compared to risperidone. First-episode

patients treated with olanzapine had superior im-

provement in overall and positive symptoms and

clinical response (Sanger et al. 1999), overall and

negative symptoms (Lieberman et al. 2003a) and

displayed a lower rate of EPS compared to haloper-

idol (Sanger et al. 1999; Lieberman et al. 2003a). In

a randomised double-blind study comparing the

neurocognitive effects of olanzapine and low doses

of haloperidol, a beneficial effect on neurocognitive

function could be observed for olanzapine, but the

difference in benefit to low-dose haloperdiol was

small (Keefe et al. 2004). In a prospective, com-

parative, open, non-randomised trial comparing

152 P. Falkai et al.

olanzapine to conventional agents, superior efficacy

in clinical response, positive and negative symptoms,

agitation and depression, and lower frequency of

EPS were observed with olanzapine (Bobes et al.

2003). Patients with first-episode schizophrenia

treated with clozapine yielded more rapid improve-

ment and remission, demonstrated better improve-

ment in clinical global impressions and showed a

reduced EPS-rate compared to chlorpromazine

(Lieberman et al. 2003b). In a Cochrane Review

including two RCTs (Emsley et al. 1999; Sanger et

al. 1999) the authors consistently found no superior

efficacy of SGAs versus FGAs in first-episode

schizophrenia, nevertheless lower EPS rates (re-

duced use of anticholinergics) were observed in

patients treated with risperidone or olanzapine

compared to haloperidol, and olanzapine revealed

superior improvement in global psychopathology

(Rummel et al. 2003). Another review suggested

SGAs as first-line treatment for first-episode patients

(Bradford et al. 2003).

In a randomised double-blind study comparing

olanzapine (mean dose 15 mg/day) to risperidone

(mean dose 4 mg/day) no difference in efficacy on

positive and negative symptoms, and in the fre-

quency and severity of adverse events, could be

detected (van Bruggen et al. 2003).

In summary there is evidence for efficacy of FGAs

(particularly haloperidol, flupenthixol, pimozide,

chlorpromazine, all Level C) and SGAs (particularly

clozapine, Level C , olanzapine and risperidone, both

Level B) in the treatment of patients with first-

episode schizophrenia.

General recommendations. In some recent guidelines,

initial treatment is recommended in an outpatient or

home setting if possible, because this approach can

minimise trauma, disruption and anxiety for the

patient and family, who are usually poorly informed

about mental illness and have fears about and

prejudices against inpatient psychiatric care (NICE

2002; RANZCP 2003). In other guidelines these

benefits are weighed against the advantages of a

hospital setting, which allows more careful monitor-

ing of the psychotic symptoms as well as any side

effects, including acute dystonia, akathisia or neu-

roleptic malignant syndrome arising from treatment

with antipsychotic medications (DGPPN 1998; APA

2004). Inpatient care is required if there is a

significant risk of self-harm or aggression, if the level

of support in the community is insufficient, or if the

crisis is too great for the family to manage, even with

home-based support. Inpatient care should be pro-

vided in the least restrictive environment (RANZCP

2003). Unfortunately, the preceding recommenda-

tions can be followed only in an ideal situation.

Often, with psychotic and possibly violent patients,

an adequate diagnostic interview or physical exam-

ination cannot be conducted. Nevertheless, no

matter how agitated the patient, the admitting staff

should do everything they can to make sure that the

patient is not suffering from another disorder that

requires emergency help (APA 2004).

Pharmacological treatments should be introduced

with great care in medication-naive patients. Skilled

nursing care, a safe and supportive environment, and

regular and liberal doses of benzodiazepines could

be given to relieve distress, insomnia and beha-

vioural disturbances secondary to psychosis, while

antipsychotic medication takes effect (RANZCP

2003).

Choice of antipsychotic medication. Based on the

results of the cited studies and clinical experience,

first-line use of SGAs (except clozapine) is recom-

mended for individuals with newly diagnosed schi-

zophrenia, mostly on the basis of better tolerability

and reduced risk of EPS, especially tardive dyskine-

sia (NICE 2002; RANZCP 2003; APA 2004). Early

use of clozapine may be considered if suicide risk is

prominent or persistent (RANZCP 2003), but

usually clozapine should not be the drug of choice

in first-episode schizophrenia. In the longer term,

the risk�/benefit ratio may change for some patients,

for example if weight gain or sexual side effects

associated with the SGA develop (RANZCP 2003).

Dosage. First SGAs, and second-line FGAs, at the

lower end of the standard dose range are the

preferred treatments for a person experiencing a

first episode of schizophrenia (DGPPN 1998; NICE

2002; RANZCP 2003; APA 2004). Based on the

mentioned findings, this dosage recommendation is

mostly confirmed for haloperidol and risperidone

(Level B), whereas for other antipsychotics there is

only sparse evidence (Level D).

Acute exacerbation (relapse)

The most common contributors to symptom relapse

are antipsychotic medication non-adherence, sub-

stance use, and stressful life events, although re-

lapses are not uncommon as a result of the natural

course of the illness, despite continuing treatment. If

nonadherence is suspected, it is recommended that

the reasons for it be evaluated and considered in the

treatment plan (DGPPN 1998; APA 2004).

Efficacy of first-generation antipsychotics. In a review,

with the exception of mepazine and promazine, all

FGAs demonstrated superior efficacy in acute treat-

ment compared to placebo, while there were differ-

WFSBP Guidelines for Biological Treatment of Schizophrenia 153

ences in dose, potency and side effects of the

different drugs (Davis et al. 1989; Baldessarini et

al. 1990; Kane and Marder 1993; Dixon et al.

1995). Haloperidol was concluded as being effica-

cious in managing acute schizophrenic episodes

compared to placebo, but also demonstrated high

propensity to cause movement disorders (Joy et al.

2004). In two RCTs, perazine showed similar

efficacy compared to haloperidol, but less EPS

under perazine treatment was observed (Schmidt

et al. 1982; Klimke et al. 1993). Two RCTs

comparing perazine with the atypical antipsychotic

zotepine showed controversial results; one of them

found a superiority of zotepine in improving psy-

chopathology (Wetzel et al. 1991), whereas the

other found an advantage of perazine in this regard

(Dieterle et al. 1991). According to these studies and

one additional trial comparing perazine with ami-

sulpride, perazine produced low EPS and had an

extrapyramidal side-effect risk similar to the men-

tioned atypical agents (Ruther and Blanke 1988).

Efficay of second-generation antipsychotics. Patients

selected in the placebo-controlled studies on amisul-

pride displayed predominantly negative symptoms

(e.g., Pailliere-Martinot et al. 1995). Compared to

haloperidol, fluphenazine or flupenthixol, a compar-

able or greater improvement of overall symptoms

(Pichot et al. 1988; Delcker et al. 1990, Moller et al.

1997; Puech et al. 1998; Wetzel et al. 1998; Colonna

et al. 2000), comparable improvement in positive

symptoms (Delcker et al. 1990; Moller et al. 1997;

Puech et al. 1998; Wetzel et al. 1998; Carriere et al.

2000, Colonna et al. 2000) and greater improvement

in negative symptoms (Moller et al. 1997; Puech et

al. 1998; Colonna et al. 2000) was found with

amisulpride. In a small, 6-week RCT comparing

amisulpride (400�/800 mg/day) with risperidone (4�/

8 mg/day) in patients with predictive positive symp-

toms, no significant difference was observed in terms

of efficacy or overall tolerability (Hwang et al. 2003).

In a review, amisulpride was judged as being at least

as efficacious as haloperidol and flupenthixol in

treating acute exacerbations of schizophrenia with

an optimal dose range of 400�/800 mg/day (Freeman

1997). In summary, there is convincing evidence for

the efficacy of amisulpride treatment in acute

episodes of schizophrenia (Level B).

In a meta-analysis pooling data from five double-

blind controlled studies comparing aripiprazole with

placebo and haloperidol in acutely relapsed patients,

aripiprazole showed a favourable safety and toler-

ability profile (Marder et al. 2003). Compared to

placebo, significantly superior improvement in ne-

gative symptoms was observed only with 15 mg/day

aripiprazole and 10 mg/day haloperidol, and not

with 30 mg/day aripiprazole (Kane et al. 2002), and

in another study with dosages of 20 and 30 mg/day

aripiprazole, and 6 mg/day risperidone (Potkin et al.

2003). Compared to haloperidol or risperidone,

randomised, double-blind controlled studies dis-

played comparable improvement of overall, positive

and negative symptoms with aripiprazole (Kane et

al. 2002; Potkin et al. 2003). In summary, there is

convincing evidence for the efficacy of aripirazole in

acute schizophrenic episodes (Level B).

For acutely ill schizophrenic patients, clozapine

showed superiority in psychopathological improve-

ment compared to placebo in two RCTs with small

sample sizes (Shopsin et al. 1979; Honigfeld et al.

1984). In European double-blind, randomised multi-

centre studies the efficacy of clozapine was compar-

able to haloperidol, chlorpromazine, trifluoperazin

and clopenthixol (Fischer-Cornelssen and Ferner

1976). In addition, other small RCTs revealed

comparable efficacy (Chiu et al. 1976; Guirguis et

al. 1977; Gelenberg and Doller 1979) or superior

overall improvement compared to chlorpromazine

(Shopsin et al. 1979). Clozapine was as efficacious as

haloperidol (Klieser et al. 1994) or risperidone

(Heinrich et al. 1994) in acute schizophrenic episodes

and superior compared to chlorpromazine in first-

episode patients (Lieberman et al. 2003). In sum-

mary, there is evidence for the efficacy of clozapine in

acute schizophrenic episodes (Level B), but because

of its side effect profile, especially haematological

adverse effects, it is not recommended in first-line

treatment of acute schizophrenia.

In acutely ill schizophrenic patients, olanzapine

treatment was superior to placebo in improvement

of overall and positive symptoms in a dose range of

5�/20 mg/day (Beasley et al. 1996a,b; Hamilton et al.

1998). In comparison to haloperidol, similar or

greater improvement of overall, positive and negative

symptoms was observed (Beasley et al. 1996b, 1997;

Tollefson et al. 1997; Hamilton et al. 1998; Revicki

et al. 1999; Ishigooka et al. 2001; Lieberman et al.

2003). In summary, there is convincing evidence for

the efficacy of olanzapine for treatment of acute

schizophrenic episodes (Level A).

In acutely relapsed patients, treatment with que-

tiapine compared to placebo was superior in im-

provement of overall and positive symptoms in

dosages of 150�/750 mg/day (Fabre et al. 1995;

Borison et al. 1996; Arvanitis et al. 1997; Small et al.

1997). There was also a significantly superior

improvement in negative symptomatology, but not

in all dose ranges (Fabre et al. 1995; Borison et al.

1996; Arvanitis et al. 1997; Small et al. 1997).

Greater or comparable improvement of overall,

positive and negative symptoms was observed with

quetiapine compared to haloperidol or chlorproma-

154 P. Falkai et al.

zine in RCTs treating acutely ill patients (Arvanitis

et al. 1997; Peuskens and Link 1997; Copolov et al.

2000). Therefore, evidence was revealed for the use

of quetiapine in the treatment of acute schizophrenic

episodes (Level A).

For acute-phase schizophrenia, in a dose range of

6�/16 mg/day, risperidone demonstrated efficacy in

placebo-controlled trials in the treatment of overall

symptoms and positive symptoms (Borrison et al.

1992; Chouinard et al. 1993; Marder and Meibach

1994). In one short-term study, only 6 mg/day

risperidone revealed a superior improvement in

negative symptoms compared to placebo in a

sample of chronic schizophrenic patients (Choui-

nard et al. 1993). Compared with haloperidol (Claus

et al. 1992; Ceskova and Svestka 1993; Chouinard et

al. 1993; Min et al. 1993; Marder and Meibach

1994; Peuskens et al. 1995; Blin et al. 1996),

perphenazine (Hoyberg et al. 1993) and flupenthixol

(Huttunen et al. 1995), there was a comparable or

better response in global psychopathology and

positive symptoms with risperidone in most fixed

doses or with flexible dosing. There was a dose-

dependent EPS rate. In most studies risperidone

produced significantly less EPS than conventional

comparator agents. In summary, there is convincing

evidence for the efficacy of risperidone treatment in

acute episodes of schizophrenia (Level A).

In double-blind, randomised trials, 80�/160 mg/

day ziprasidone was superior in improving overall

and positive symptoms compared to placebo in

acutely ill patients (Keck et al. 1998; Daniel et al.

1999). In one study there was also superior im-

provement in negative symptomatology (Daniel et

al. 1999). Compared with haloperidol, ziprasidone

showed comparable improvement in overall, positive

and negative symptoms in acute phase RCTs (Goff

et al. 1998; Hirsch et al. 2002). In summary there is

convincing evidence for the efficacy of ziprasidone in

treatment of acute schizophrenic episodes (Level A).

One RCT, comparing 150�/300 mg/day zotepine to

placebo, demonstrated superior improvement in

overall, positive and negative psychopathology

(Cooper et al. 2000a). Compared to haloperidol

(Fleischhacker et al. 1989; Barnas et al. 1992; Petit

et al. 1996; Hwang et al. 2001), chlorpromazine

(Cooper et al. 2000a) and perazine (Dieterle et al.

1991; Wetzel et al. 1991), zotepine showed compar-

able or superior improvement in global psycho-

pathology, positive and negative symptoms. In

summary, there is good evidence for the efficacy

of zotepine for treatment of acute schizophrenia

(Level A).

General recommendations. It is recommended that in

multiple-episode patients, antipsychotic pharmaco-

logical treatment should be initiated promptly,

provided it will not interfere with diagnostic assess-

ment, because acute psychotic exacerbations may be

associated with emotional distress, disruption to the

patient’s life, and a substantial risk of dangerous

behaviours to self, others or property (APA 2004).

Antipsychotic monotherapy is recommended across

the guidelines (e.g., DGPPN 1998; NICE 2002;

RANZCP 2003; APA 2004) in the initial treatment

of acute schizophrenic episode (Level D).

Choice of antipsychotic medication. The choice of

antipsychotic drug should be made jointly by the

individual and the clinician responsible for treatment

based on an informed discussion of the relative

benefits of the drugs and their side-effect profiles

(NICE 2002; RANZCP 2003; APA 2004). When

full discussion between the clinician responsible for

treatment and the individual concerned is not

possible, in particular in the management of an

acute schizophrenic episode, the oral SGAs should

be considered as the treatment options of choice

because of the lower potential risk of extrapyramidal

symptoms (EPS) (NICE 2002) and the at least

comparable efficacy compared to FGAs (RANZCP

2003; APA 2004). Nevertheless, selection of an

antipsychotic medication is frequently guided by

the patient’s previous experience with antipsycho-

tics, including the degree of symptom response, past

experience of side effects, preferred route of medica-

tion administration, the presence of comorbid med-

ical conditions, and potential interactions with other

prescribed medications (DGPPN 1998; APA 2004).

All available SGAs, except clozapine (RANZCP

2003; APA 2004), should be considered as treat-

ment options for individuals currently receiving

conventional antipsychotic drugs who, despite ade-

quate symptom control, are experiencing unaccep-

table side effects, and for those in relapse who have

previously experienced unsatisfactory management

or unacceptable side effects with conventional anti-

psychotic drugs. It is not recommended that, in

routine clinical practice, individuals change to one of

the oral atypical antipsychotic drugs if they are

currently achieving good control of their condition

without unacceptable side effects with conventional

antipsychotic drugs (Level D) (DGPPN 1998; NICE

2002; APA 2004).

Dosage. An overview of studies comparing different

dosages of FGAs concluded that daily doses lower

than 300 mg CPZ equivalents were inadequate for

optimal treatment and doses above 940 mg CPZ

equivalents produced no better responses than in the

range of 540�/940 mg CPZ equivalents (Davis et al.

1989) (Level A). Another review found superior

WFSBP Guidelines for Biological Treatment of Schizophrenia 155

improvement in nearly two-thirds of trials using

doses of 300 mg CPZ equivalents or less, and

consistent superiority for daily doses of 500 mg

CPZ equivalents and more compared to placebo

(Baldessarini et al. 1990) (Level A). In addition the

best dose-dependent response was found in a range

of 2�/10 mg/day haloperidol (Baldessarini et al.

1990). No significant advantage was found for

dosages greater than 10�/20 mg/day haloperidol in

acute treatment compared to lower doses (Kane and

Marder 1993; Dixon et al. 1995). Recent RCTs of

FGAs for acute treatment focusing on dosing

strategies have consistently found that modest doses

(mostly less than 10 mg/day haloperidol or equiva-

lent, or plasma levels B/18 ng/ml haloperidol) were

as efficacious or more efficacious than higher doses

(Coryell et al. 1998; Stone et al. 1995; Volavka et al.

2000) (Level A). Moderate doses of FGAs were

noted to improve comorbid depression (Koreen et

al. 1993; Volavka et al. 1996; Krakowski et al. 1997),

whereas higher doses were associated with a greater

risk of EPS and dysphoria (Bollini et al. 1994;

Krakowski et al. 1997). In a systematic review of

16 RCTs with 19 different randomised dose com-

parisons of haloperidol , using low doses (between 3

and 7.5 mg/day compared to 7.5�/15 and 15�/35 mg/

day, respectively) did not result in loss of efficacy,

but were associated with a lower rate of clinically

significant extrapyramidal adverse effects than

higher doses (Wairach et al. 2004) (Level A). A

further review stated that the near-maximal efficacy

dose for haloperidol was between 3 and 10 mg/day,

nevertheless high doses of haloperidol were found to

be no less effective than medium doses (Davis and

Chen 2004) (Level A). Individual adaptation of the

dose, and not standard dosing, seems to be the best

treatment strategy (Klieser and Lehmann 1987;

Dixon et al. 1995).

In summary, the recommendation of daily dosages

between 300 and 1000 mg CPZ equivalents for

FGAs in the treatment of an acute symptom episode

for a minimum of 6 weeks remains stable across the

guidelines and over time (e.g., APA 1997, 2004;

DGPPN 1998; Lehman et al. 1998; Working Group

for the Canadian Psychiatric Association 1998;

NICE 2002), whereby the minimum effective dose

should be used. The optimal dose for each patient

has to be found by clinical judgment.

For SGAs randomised, placebo-controlled studies

which compared two or more doses of an antipsy-

chotic were used in a review to calculate the dose�/

response curve (for each FGA or SGA , and as a

group based on dose equivalence). The near-max-

imal effective dose was defined as the threshold dose

necessary to produce all or almost all the clinical

responses for each drug. It was found that the

near-maximal efficacy dose for risperidone was 4

mg/day, for ziprasidone 120 mg/day, for aripiprazole

10 mg/day, for clozapine greater than 400 mg/day

and for olanzapine probably greater than 16 mg/day

(Davis and Chen 2004). Evidence for dose adjust-

ment based on clinical routine, which differs from

initial recommendations of approval and marketing

studies, came from national surveys. From 1997

until 2001 the mean dose of risperidone used for

inpatients in the New York state system decreased

from 7.1 to 4.9 mg/day (Citrome et al. 2002).

Additionally, based on a retrospective survey, a less

rapid titration of risperidone (0.5�/2 mg/day) was

recommended to keep patients compliant with their

medication (Luchins et al. 1998). On the other

hand, doses of olanzapine in the New York state

system for inpatients increased; in 2001 nearly 26%

of olanzapine patients received doses greater than 20

mg/day (Citrome et al. 2002). Based upon a review

of published findings and clinical experience, a more

rapid initiation schedule for quetiapine was pro-

posed than currently provided for treatment in

hospitalised patients with acute schizophrenia. Ad-

ditionally, higher doses (up to 1600 mg/day) of

quetiapine had been well tolerated in some patients

due to its favourable tolerability profile (Arango and

Bobes 2004). A review concluded that there is still

not enough data available to ensure a clear dose�/

response relationship of all approved SGAs and to

make recommendations of optimal dosing strategies

(Kinon et al. 2004). The dose may be titrated as

quickly as tolerated to the target therapeutic dose of

the antipsychotic medication, and unless there is

evidence that the patient is having uncomfortable

side effects (APA 2004). There is broad agreement

in reviews (e.g., Davis et al. 1989; Baldessarini et al.

1990; Kane and Marder 1993; Dixon et al. 1995;

Davis and Chen 2004; Kinon et al. 2004)

and guidelines (e.g., DGPPN 1998; NICE 2002;

APA 2004) that massive loading doses of antipsy-

chotic medication referred to as ‘rapid neuroleptisa-

tion’ do not provide any advantage over standard

dosing in initial treatment and may be associated

with a higher risk of EPS (Level D). Therefore

this treatment strategy should not be used in the

treatment of the acute episode for people with

schizophrenia.

Specific clinical features influencing the

treatment plan

Treatment of predominantly positive symptoms

In common practice patients with an acute schizo-

phrenic episode present with predominantly positive

symptoms. This topic was already discussed in a

156 P. Falkai et al.

previous section of this guideline. It is recommended

that in multiple-episode patients antipsychotic phar-

macological treatment should be initiated promptly,

provided it will not interfere with diagnostic assess-

ment, because acute psychotic exacerbations may be

associated with emotional distress, disruption to the

patient’s life, and a substantial risk of dangerous

behaviours to self and others. For first-episode

patients, short-term observation and administration

of low-dose benzodiazepines may be helpful in

establishing the diagnosis before antipsychotic treat-

ment is introduced. Other psychoactive medications

are commonly added to antipsychotic medications

when patients continue to demonstrate active psy-

chotic symptoms despite adequate medication trials.

For persisting positive symptoms despite pharma-

cotherapy, treatment-resistance should be consid-

ered (see this section in the guideline).

In summary no specific treatment recommenda-

tions are given for patients with predominantly

positive symptoms and the reader is referred to the

section discussing strategies for treating acute re-

lapse or other specific clinical features influencing

the treatment plan.

Treatment of agitation

Schizophrenic patients show agitated, aggressive or

violent behaviour, mostly related to psychotic symp-

toms (e.g., persecutory delusions, mania or halluci-

nations), or as a result of other symptoms, such as

threatening and anxiety, when internal controls are

compromised (Angermeyer 2000). Factors relating

to the patient’s environment or the institutions

involved in treatment, such as crowded wards, lack

of privacy and long waiting times, contribute to the

occurrence of aggressive behaviour. The prediction

of aggressive and violent behaviour during hospita-

lisation is difficult; however, an association was

seen with hostility and thought disorders (Steinert

2002). Physician and staff confronted with an

acutely ill, aggressive patient with schizophrenia

should provide structure, reduce stimulation, try to

verbally reassure and calm the person, and to de-

escalate the situation at the earliest opportunity

(Osser and Sigadel 2001). If possible, oral adminis-

tration of medications is preferable to parenteral

administration. The lowest effective dose should

be given, and, if necessary, increased incrementally.

Emergency management of violence in schizophre-

nia may include sedation, and as the last option

restraint and seclusion. Similarly, in this context

the use of drugs to control disturbed behaviour

(rapid tranquillisation) is often seen as a last resort,

where appropriate psychological and behavioural

approaches have failed or are inappropriate. The

aim of drug treatment in such circumstances is to

calm the person, and reduce the risk of violence and

harm, rather than treat the underlying psychiatric

condition. Psychiatrists, and the multidisciplinary

team, who use rapid tranquillisation should be

trained in the assessment and management of service

users specifically in this context: this should include

assessing and managing the risks of drugs (benzo-

diazepines and antipsychotics), using and maintain-

ing the techniques and equipment needed for

cardiopulmonary resuscitation, and prescribing

within therapeutic limits and using flumazenil

(benzodiazepine antagonist) (DGPPN 1998; NICE

2002; APA 2004).

Two RCTs found that the combination of haloper-

idol (5 mg) and lorazepam (4 mg) intramuscularly

produces an overall superior and faster clinical

response than haloperidol alone (Bieniek et al.

1998; Garza-Trevino et al. 1989). Comparing

monotherapy of benzodiazepines with antipychotics

alone, lorazepam (or flunitrazepam) and haloperidol

administered intramuscularly demonstrated similar

efficacy in controlling agitation and general response

to treatment (Battaglia et al. 1997; Foster et al.

1997; Dorevitch et al. 1999). In one study, loraze-

pam 2 mg was superior in improvement of gobal

impression compared to haloperidol 5 mg (Foster et

al. 1997). The administration of midazolam 15 mg

was superior in terms of sedation (and therefore

reducing agitation) compared to the combination of

haloperidol (5 mg) and promethazine (50 mg), both

intramuscularly, in an open, randomised, controlled

study (TREC 2003).

While the use of SGAs , with a lower liability for

extrapyramidal side effects, show promise for rapid

tranquilisation, a study comparing olanzapine (10

mg intramuscularly) to haloperidol (7.5 mg intra-

muscularly) observed similar efficacy in reducing

agitation at 2 and 24 hours after the first injection

(Wright et al. 2001). Olanzapine demonstrated a

favourable side effect profile, e.g., reduced addi-

tional need for benzodiazepines, less dystonia and

EPS, and less need to receive anticholinergic drugs

(Altamura et al. 2003). There is a risk of sudden

death following intramuscular application of olanza-

pine and benzodiazepines, therefore the combined

use should be avoided. An open-label study demon-

strated equal efficacy of ziprasidone 20�/80 mg

intramuscularly compared to 10�/40 mg haloperidol

intramuscularly in acute schizophenic patients with

agitation (Swift et al. 2003), and a randomised

study showed comparable efficacy of ziprasidone

40 mg and haloperidol 10 mg intramuscularly

(Brook et al. 2000). In addition, a dose-finding

study showed superiority for ziprasidone 20 mg

compared to ziprasidone 2 mg intramuscularly in

WFSBP Guidelines for Biological Treatment of Schizophrenia 157

reducing acute agitation (Daniel et al. 2001). Rapid

sedation may also be achieved through administra-

tion of low potency antispychotics (e.g., levomepro-

mazine, chlorprothixene) or zuclopenthixolacetate

(DGPPN 1998), but this strategy is not recom-

mended anymore in recent guidelines (e.g., APA

2004).

If oral treatment is accepted the combination of

oral risperidone (2 mg) and lorazepam (2 mg)

appears to be comparable to intramuscular haloper-

idol (5 mg) and lorazepam (2 mg) (Currier and

Simpson 2001).

Recommendations. Lorazepam and conventional neu-

roleptic agents showed comparable efficacy in the

acute treatment of aggression and psychomotor

agitation (Level C). Due to the more favourable

side effect profile of lorazepam, initial treatment

should be performed with 2�/4 mg lorazepam in

patients in whom no decision has yet been taken on

whether to follow a medicinal or non-medicinal

strategy or on which type of antipsychotic treatment

to use. Administration of diazepam or other benzo-

diazepines (apart from lorazepam) or of low-potency

neuroleptic agents, such as chlorprothixene or levo-

promazine, is not recommended in the treatment of

agitation and excitation due to inferior efficacy or

inferior tolerability. In patients whose aggressive

behaviour is clearly due to psychotic symptoms, a

combination treatment of lorazepam with a neuro-

leptic agent can be undertaken (Level C). Due to

better tolerability, an atypical neuroleptic agent,

such as olanzapine or ziprasidone, preferably in

parenteral form, may be used where possible in

preference to a conventional neuroleptic agent (Level

C). There is a risk of sudden death following

intramuscular application of olanzapine and benzo-

diazepines, therefore the combined use should be

avoided. Similarly also caution has to be recom-

mended when combining clozapine with benzodia-

zepines (Rupprecht et al. 2004). Clinicians must also

be aware of cardiac abnormalities, especially when

using ziprasidone intramuscularly.

If extremely rapid sedation is urgently required,

haloperidol and lorazepam can be parenterally

administered. If treatment is not sufficient to treat

the symptoms of excitation/tension or anxiety, addi-

tional treatment with carbamazepine, valproate or

lithium may be considered (Level D). Measures such

as restraint and seclusion should only be used in

exceptional emergency situations. They should be

carefully documented and explained to the patient.

In all cases, the patient should be allowed to express

his or her opinions and discuss his or her experience.

The physician should see a secluded or restrained

patient as frequently as needed to monitor any

changes in the patient’s physical or mental status

and to comply with local law.

Treatment of predominantly negative symptoms

Negative symptoms in schizophrenia can be differ-

entiated into primary negative symptoms, suggesting

a core symptomatology in schizophrenia, and sec-

ondary negative symptoms as a consequence of

positive symptoms (e.g., social withdrawal because

of paranoid ideas), due to EPS (e.g., neuroleptic-

induced akinesia), depressive symptoms (e.g., post-

psychotic or pharmacogenic depression) or environ-

mental factors (e.g., social understimulation due to

hospitalism) (Carpenter et al. 1985).

Due to their pharmacological profile, especially

the preferential blockade of 5-HT2A receptors,

SGAs were concluded to possess favourable ability

to treat negative symptoms compared to FGAs (e.g.,

Moller 2003). Unfortunately most trials were carried

out in patients experiencing acute exacerbations or

presenting with a mixture of positive and negative

symptoms, and therefore the improvement in nega-

tive symptoms could be interpreted as a decrease in

secondary negative symptoms. Overlap between

adverse drug effects (EPS) and depression makes

interpretation of the study results difficult. In addi-

tion most comparator doses of FGAs were retro-

spectively judged as too high and associated with a

high EPS rate. Even when more elaborate statistical

approachs, e.g., path analysis, are used, interpreta-

tion requires caution (Moller et al. 1995). On the

other hand the efficacy of atypical antipsychotics in

treating negative symptoms may be underestimated

in meta-analysis due to methodological pitfalls

inherent to such analyses (Moller 2003).

Efficacy of first-generation antipsychotics. In most

studies there is also improvement of negative symp-

toms with FGAs but the trials focus mainly on

positive symptoms (Dixon et al. 1995). Compared to

placebo there is evidence for the efficacy of FGAs in

treating negative symptoms (e.g., Davis et al. 1989).

But even when ratings for negative symptoms are

given, primary and secondary negative symptoms

are not distinguished. Comparison studies to SGAs

(as mentioned later) demonstrate a tendency for

lower doses to be favourable in treating negative

symptoms. There were no studies in patients with

predominantly negative symptoms.

Efficacy of second-generation antipsychotics. Amisul-

pride in dosages up to 800 mg/day showed superior

improvement compared to haloperidol (20 mg/day)

(Moller et al. 1997) and similar efficacy compared to

risperidone (8 mg/day) (Peuskens et al. 1999) in

158 P. Falkai et al.

treating negative symptoms in short-term RCTs. In

a 1-year, double-blind randomised maintenance

study with flexible doses, amisulpride was associated

with greater improvement in negative symptomatol-

ogy compared to haloperidol (Colonna et al. 2000).

Selecting patients with predominantly negative

symptoms a randomised double-blind long-term

trial comparing six dose levels of amisulpride with

haloperidol revealed better, but no significant im-

provement in negative symptoms after 1-year treat-

ment in favour of amisulpride (Speller et al. 1997).

Two RCTs showed superiority of amisulpride com-

pared to placebo in improvement of negative symp-

toms, at dosages of 100�/300 mg/day over 6 weeks

(Boyer et al. 1995) and 100 mg/day over 6 months

(Loo et al. 1997). Additionally, two RCTs displayed

better efficacy of amisulpride in the treatment of

negative symptoms compared to placebo in a patient

sample suffering predominantly from persistent

negative symptomatology (Palliere-Martinot et al.

1995; Danion et al. 1999). In meta-analyses (Leucht

et al. 1999, 2002; Leucht 2004), amisulpride,

olanzapine and risperidone were found to reveal

superiority in treating negative symptoms compared

to FGAs, derived from studies of acutely ill patients.

In three small RCTs comparing amisulpride with

FGAs in patients with predominantly negative

symptoms (Pichot and Boyer 1989; Saletu et al.

1994; Speller et al. 1997), there was only a trend in

favour of amisulpride but no statistical difference.

Nevertheless, amisulpride is the only SGA that has

been studied extensively in this patient population,

and especially in regard to the placebo-controlled

studies there is evidence that treatment with ami-

sulpride is effective at a dose range of 50�/300 mg/

day in improving negative symptoms (Leucht 2004)

(Level A).

Aripiprazole at a dosage of 15 mg/day, but not 30

mg/day and haloperidol 10 mg/day showed superior

efficacy in improvement of negative symptoms

compared to placebo in an RCT of 4 weeks duration

(Kane et al. 2002). In another short-term RCT,

efficacy of aripiprazole (20 and 30 mg/day) in

treating negative symptoms was comparable to

risperidone (6 mg/day) (Potkin et al. 2003). In a

placebo-controlled RCT, aripipazole demonstrated

superior efficacy in improvement of negative symp-

toms over 6 months (Pigott et al. 2003). Pooled data

of two 52-week RCT comparing aripiprazole 30 mg/

day to haloperidol 10 mg/day showed superior

improvement in negative symptomatology in favor

of aripiprazole (Kasper et al. 2003). In summary,

although there is evidence for efficacy in treating

negative symptoms (Level A), there is no clear

experience with aripiprazole in patients with pre-

dominantly negative symptoms.

Clozapine was found to be effective in open, non-

comparative trials in treatment refractory patients

with more-or-less predominant negative symptoms

(Meltzer et al. 1989; Meltzer 1992; Lindenmayer et

al. 1994) and in a double-blind trial compared to

chlorpromazine (Kane et al. 1988). In another

double-blind RCT, with a small sample size, no

difference between clozapine and haloperidol in the

improvement of negative symptoms could be de-

tected (Breier et al. 1994). In a double-blind multi-

comparative RCT already mentioned above,

clozapine (target dose 500 mg/day) revealed statis-

tically significant superiority in impoving negative

symptoms compared to haloperidol (target dose 20

mg/day) in patients with suboptimal response to

previous treatment, but this effect was described as

clinically modest (Volavka et al. 2002). While one

meta-analysis found that there is slight significant

evidence for superiority of clozapine compared to

FGAs in the treatment of negative symptoms

(Wahlbeck et al. 2004), another meta-analytic review

reported an advantage of clozapine in this regard

evaluating its efficacy in treatment-resistant patients

(Chakos et al. 2001). In summary, although there is

evidence for efficacy in treating negative symptoms

(Level A), there is only little experience with

clozapine in patients with predominantly negative

symptoms.

Olanzapine displays evidence for superior efficacy

in treating negative symptoms in acute-phase RCTs

compared to placebo and haloperidol (Beasley et al.

1996a,b, 1997; Tollefson et al. 1997). Nevertheless,

in an extension study after 24 weeks there was no

longer a statistically significant difference between

olanzapine and haloperidol in reducing negative

symptoms (Hamilton et al. 1998). A path-analysis

of these studies found that most of the changes in

negative symptoms could not be explained by other

compounds (positive symptoms, depression, EPS)

(Tollefson et al. 1997). While one short-term

double-blind randomised study demonstrated super-

iority of olanzapine (mean dose 17.2 mg/day) in

improving negative symptoms compared to risper-

idone (mean dose 7.2 mg/day) (Tran et al. 1997),

another RCT (mean dose olanzapine 12.4 mg/day

versus risperidone 4.8 mg/day) could not replicate

this finding, probably due to lower dosage of

risperidone associated with less EPS (Conley and

Mahmoud 2001). In summary, although there is

evidence for efficacy in treating negative symptoms

(Level A), there is no clear experience with

olanzapine in patients with predominantly negative

symptoms.

Quetiapine produced significant superior improve-

ment in negative symptoms compared to placebo

only in higher dosage (750 mg/day) acute phase

WFSBP Guidelines for Biological Treatment of Schizophrenia 159

treatment in one RCT (Small et al. 1997) and in

another trial in the whole dose range (75�/750 mg/

day), with best results at the 300-mg/day dose

(Arvanitis et al. 1997). Compared to haloperidol

(12 mg/day) there was no significant difference

(Arvanitis et al. 1997), and compared to chlorpro-

mazine (750 mg/day) a trend towards better efficacy

in negative symptomatology could be observed

(Peuskens and Link 1997). Overall, there is evidence

for similar efficacy, but not for significant advantages

compared to FGAs (Cheer and Wagstaff 2004) in

the treatment of negative symptoms (Level A).

Risperidone showed significantly superior improve-

ment in negative symptoms compared to haloperidol

and placebo only at a dosage of 6 mg/day (Peuskens

et al. 1995). A re-analysis using the path-analytical

approach revealed a direct effect of treatment on

negative symptoms in this study (Moller 2003). In

partially refractory schizophrenic patients, similar

efficacy was found in improvement of negative

symptoms compared to clozapine and perphenazine

(Moller 2003). In double-blind randomised compar-

isons with olanzapine there were similar results or

inferior improvement in negative symptomatology

(Tran et al. 1997; Conley and Mahmoud 2001). In

maintenance treatment compared to haloperidol,

beneficial effects of negative symptomatology were

reported for risperidone (Csernansky et al. 2002). A

meta-analyis of the pooled results from six double-

blind RCTs comparing risperidone to FGAs found

that risperidone showed significantly superior im-

provement in negative symptoms (Carman et al.

1995). In summary, there is evidence for efficacy in

treating negative symptoms (Level A), but no clear

experience in patients with predominantly negative

symptoms.

Ziprasidone showed superior efficacy for improve-

ment of negative symptoms compared to placebo in

one short-term RCT (Daniel et al. 1999), but not in

a statistically significant manner in another RCT

(Keck et al. 1998), both in acutely ill patients. In a

double-blind randomised extension study over 1

year, including patients with chronic schizophrenia

presenting predominantly negative symptoms, a

statistically significant superior improvement in

negative symptoms was observed in favour of

ziprasidone (dosage 40, 80 and 160 mg/day) com-

pared to placebo at endpoint (Arato et al. 2002). In

summary, there is evidence for efficacy in treating

negative symptoms (Level A), and limited evidence

in this regard in patients with predominantly nega-

tive symptoms (Level C).

Zotepine revealed inconsistent efficacy for superior

improvement of negative symptoms compared to

FGAs in earlier RCTs (Moller 2003), but more

recent double-blind RCTs demonstrated significant

advantages of zotepine compared to haloperidol

(Petit et al. 1996) or chlorpromazine (Cooper et al.

2000a). A placebo-controlled study in patients with

predominant negative symptoms failed to demon-

strate efficacy of zotepine (Moller et al. 2004). A

relapse-prevention double-blind RCT compared to

placebo displayed no significant differences in regard

to negative symptomatology over 26 weeks (Cooper

et al. 2000b). In summary, there is evidence for

efficacy in treating negative symptoms (Level A),

and no evidence in patients with predominantly

negative symptoms.

Efficacy of antidepressive agents. Despite atypical

antipsychotic agents, antidepressants are used as

adjunctive treatment in patients with predominantly

negative symptoms (APA 2004). The role of this

strategy still remains unclear, because the available

studies (most of them performed with SSRI) are

inconsistent and often lack high methodological

standards (Moller 2004a). An earlier RCT indicated

that, e.g., imipramine added to long-acting FGA,

may provide benefits in negative symptoms in stable

outpatients (Siris et al. 1991), whereas addition of

desipramine or amitriptyline in acutely decompen-

sated patients was associated with poorer antipsy-

chotic response without improving depression

(Kramer et al. 1989). Marprotiline revealed no

significant difference in a double-blind crossover

study (Waehrens and Gerlach 1980).

In six placebo-controlled studies of SSRIs for

negative symptoms, one reported a modest advan-

tage of fluoxetine 20 mg/day added to long-acting

injectable antipsychotic medication (Goff et al.

1995) and another reported significant superior

improvement in negative symptoms with fluoxetine

(Spina et al. 1994), while four found no advantage

for SSRIs, compared with placebo, in patients

receiving fluoxetine combined with ongoing cloza-

pine (Buchanan et al. 1996) and fluoxetine (Arango

et al. 2000), citalopram (Salokangas et al. 1996), or

sertraline (Lee et al. 1998) added to first-generation

antipsychotics. Four controlled studies of adjunctive

fluvoxamine (100 mg/day) have demonstrated posi-

tive results (Silver and Nassar 1992; Silver and

Shmugliakov 1998; Silver et al. 2000, 2003), while

there was no benefit for marprotiline (100 mg/day)

added to antipsychotic treatment (Silver and Shmu-

gliakov 1998). In a double-blind placebo-controlled

study, mirtazapine demonstrated superior improve-

ment in negative symptomatology after 6 weeks

(Berk et al. 2001). In contrast, reboxetine (8 mg/

day) showed no effects on negative smptoms in a

double-blind placebo-controlled trial (Schutz and

Berk 2001). A placebo-controlled trial found no

advantage for adjunctive selegiline compared to

160 P. Falkai et al.

placebo (Jungemann et al. 1999). Overall, the

evidence for efficacy of antidepressants for negative

symptoms of schizophrenia is limited (Level C),

especially when taking into consideration the fact

that differentiating the improvement in depressive

symptoms from negative symptoms is difficult in

some cases. Since most of the studies were per-

formed in combination with first-generation anti-

psychotics, it is possible that the findings might

be different with second-generation antipsy-

chotics, although this possibility seems unlikely

(APA 2004). Adding�/at least TCA�/to acutely

decompensated patients, might worsen the psychotic

symptoms.

Efficacy of other medications. Earlier reports indicated

that lithium augmentation to antipsychotics im-

proved negative symptoms specifically (Small et al.

1975; Growe et al. 1979), but this finding could not

be confirmed in later trials and meta-analyses (e.g.,

Leucht et al. 2004). There is some evidence for

adding glutamatergic agents, e.g., d-cycloserine

(Moller 2003; APA 2004), and the combination of

adjunctive d-serine with FGAs or risperidone in

treating negative symptoms (Tsai et al. 1998). In

addition, there is no clear evidence for the efficacy of

oestrogen augmentation or augmentation with cog-

nitive enhancers, but pilot studies demonstrated

encouraging results for improvement (Moller

2003). For further treatment strategies please refer

to the section below (see Treatment-resistant schizo-

phrenia).

Recommendations. For the treatment of negative

symptoms, second-generation antipsychotics should

be preferred (Level A). Of the atypical compounds,

amisulpride seems to have advantages and this

antipychotic is the only one that has been investi-

gated in several studies in this special schizophrenic

population (Level A). Clozapine may be superior

compared to other antipsychotics when treating

negative symptoms in the context of treatment-

resistant schizophrenia (Level B). In cases of inade-

quate response comedication with SSRIs (Level B)

and possibly mirtazapine (Level C) may be bene-

ficial. The pharmacokinetic interactions with SSRIs

have to be considered carefully. Add-on therapies

with glutamatergic agents or oestrogen may be

discussed as experimental approaches.

Treatment of cognitive symptoms

Neurocognitive deficits have been recognised as an

important feature, or even a core deficit, of schizo-

phrenia. Cognitive functioning is a correlate of

global and specific functional outcome in schizo-

phrenia and cognitive impairments account for

significant variance in measures of functional

status (Green 1996). SGAs have been reported to

have more beneficial effects on cognitive functioning

than FGAs, nevertheless the methodology used to

assess cognitive deficits in schizophrenia has been

deficient in many clinical studies (Harvey and Keefe

2001).

Efficacy of first- and second-generation antipsychotics. In

reviews and most studies, FGAs demonstrated no or

only minor beneficial effects on cognition (e.g.,

Cassens et al. 1990; Sharma 1999), whereby in-

appropriately large dose ranges, combined with EPS

or concomitant anticholinergic medication, may

have had a negative effect on cognition. A meta-

analysis of 20 clinical trials (consisting of 11 switch-

ing studies, four comparative randomised open

studies and five randomised double-blind studies)

revealed evidence that SGAs show superior improve-

ment in essential aspects of cognition compared to

FGAs (Harvey and Keeefe 2001) (Level A). This

could be confirmed for some cognitive domains in a

randomised double-blind study comparing olanza-

pine, risperidone, clozapine and haloperidol in

patients with a history of suboptimal response to

conventional antipsychotics (Bilder et al. 2002). A

systematic review showed superior beneficial effects

on neurocognition in patients treated with SGAs

(clozapine, risperidone, olanzapine, quetiapine

and zotepine) compared to FGAs, although some

studies provided conflicting results and there was a

variety of methodological limitations (Weiss et al.

2002). In addition, a randomised double-blind

study demonstrated comparable cognitive-enhan-

cing effects relative to previous treatment (mostly

haloperidol or risperidone) in acutely ill inpatients

treated with olanzapine or ziprasidone (Harvey et al.

2004).

In contrast to these results, risperidone (6 mg/day)

compared to low-dose haloperidol (5 mg/day)

showed no superior improvement of neurocognitive

deficits over a 2-year period in a randomised double-

blind study (Green et al. 2002). In a randomised

double-blind trial in first-episode psychosis olanza-

pine (mean 9.6 mg/day) demonstrated only a small

advantage with respect to neurocognitive deficits

compared to low-dose haloperidol (mean 4.6 mg/

day) (Keefe et al. 2004).

Recommendations. In schizophrenic patients with

cognitive deficits, SGAs provide an at least modest

beneficial effect on neurocognitive functions com-

pared to FGAs (Level A), although some studies

revealed conflicting results. Adjunctive medications,

previous treatments and doses of FGAs have to be

WFSBP Guidelines for Biological Treatment of Schizophrenia 161

taken into consideration before switching to SGAs to

improve neurocognition.

Treatment of predominantly catatonic symptoms

Significant catatonic symptoms tend to be present in

close to 10% of patients admitted to psychiatric

facilities (Blumer 1997). While the catatonic subtype

of schizophrenia is diagnosed in approximately 5%

of all first-episode patients (Jablensky et al. 1992),

malignant catatonia is extremely rare. The differ-

entiation between catatonia and neuroleptic malig-

nant syndrome may be impossible in some cases,

mentioned as ‘catatonic dilemma’ (Lausberg and

Hellweg 1998).

Efficay of first-generation antipsychotics. Antipsycho-

tics, especially FGAs , demonstrated poor efficacy in

treating catatonia (e.g., Zemlan et al. 1986; Hawkins

et al. 1995) (Level C). Additionally, patients with

past or present catatonic symptoms are particularly

vulnerable to neuroleptic malignant syndrome

(NMS) (Lausberg and Hellweg 1998). Therefore

antipsychotics with preferentially D2-blocking prop-

erties may not be beneficial for treating catatonia,

and even worsen catatonic symptoms (Blumer

1997).

Efficay of second-generation antipsychotics. Some case

reports suggest that SGAs may be more effective in

treating catatonic symptoms than FGAs, e.g., ami-

sulpride (French and Eastwood 2003), clozapine

(e.g., Lausberg and Hellweg 1998; Gaszner and

Makkos 2004), olanzapine (Martenyi et al. 2001),

risperidone (Poyurousky et al. 1997; Kopala and

Caudle 1998; Hesslinger et al. 2001; Valevski et al.

2001) and zotepine (Harada et al. 1991) (Level D).

Others. Benzodiazepines revealed efficacy in acute

catatonic reactions (e.g., Rosebush et al. 1990;

Ungvari et al. 1994; Bush et al. 1996; Lee et al.

2000) and showed benefits added to antipsychotics

in chronic catatonia (Ungvari et al. 1999) (Level C).

Patients with prominent catatonic features seemed

to derive particular benefit from treatment with

ECT ; nevertheless, the evidence is limited by the

inclusion of patients with mood disorder diagnoses

and consists primarily of case reports, case series and

open prospective trials (e.g., Bush et al. 1996;

Petrides et al. 1997; Suzuki et al. 2003) (Level C).

Reviewing case reports and case series, ECT also

appeared as efficacious as lorazepam (improvement

85 vs. 79%) and was more likely to provide a positive

outcome in cases of malignant catatonia (Hawkins et

al. 1995). Findings from the above cited studies

confirm the clinical impression that ECT is bene-

ficial in patients with catatonic schizophrenia who

have not responded to first-line treatment with

lorazepam (Level D) (APA 2004).

Recommendations. Benzodiazepines, e.g., lorazepam,

appear to offer a safe, effective, first-line treatment of

catatonia (Level C). ECT should be considered

when rapid resolution is necessary (e.g., malignant

catatonia) or when an initial lorazepam trial fails

(Level C). When initiating antipsychotic treatment

SGAs, e.g., clozapine, should be preferred in

patients presenting catatonic symptoms, because of

a reduced risk of developing neuroleptic malignant

syndrome, and suggested higher efficacy (Level D).

Treatment-resistant schizophrenia

Depending upon the definition of treatment-resis-

tant schizophrenia (TRS), about 10�/30% of pa-

tients have little or no response to antipsychotic

medications, and up to an additional 30% of patients

have partial responses to treatment, meaning that

they exhibit improvement in psychopathology but

continue to have mild to severe residual hallucina-

tions or delusion (e.g., Brenner et al. 1990; Essock et

al. 1996a). Even if a patient’s positive symptoms

remit with antipsychotic treatment, other residual

symptoms, including negative symptoms and cogni-

tive impairment, often persist. Treatment resistance

is often associated with long periods of hospitalisa-

tion. However, chronic hospitalisation may also

occur in the presence of less severe psychotic

symptoms and it is not a reliable indicator of poor

response to antipsychotics. The use of widespread

criteria for TRS, including functional level, led to

prevalences of 55�/65% following treatment with

SGAs, a figure which would probably be even higher

if cognitive deficits and poor quality of life were also

included (e.g., Helgason 1990; Hegarty et al. 1994).

Treatment may be completely or partially unsuccess-

ful for a variety of reasons. The patient may receive a

suboptimal dose of antipsychotic, either because an

inadequate dose has been prescribed or because of at

least partial non-adherence, or the prescribed anti-

psychotic may be partially or fully ineffective (APA

2004). Substance use may also cause or contribute

to treatment resistance. Nevertheless TRS may be

associated with neurobiological factors (e.g., brain-

morphological abnormalities) or may depend on

environmental factors (e.g., unfavourable familial

atmosphere, high expressed emotions). Multidimen-

sional evaluation of TRS should consider persistent

positive or negative symptoms, cognitive dysfunction

with severe impairment, bizarre behaviour, recurrent

affective symptoms and suicidal behaviour, deficits

in vocational and social functioning and a poor

162 P. Falkai et al.

quality of life. Therefore, in suspected TRS, the

target symptoms should be precisely defined. Due to

modified criteria in recent guidelines, treatment

resistance is assumed if there is either no improve-

ment at all or only insufficient improvement in the

target symptoms, despite treatment at the recom-

mended dosage for a duration of at least 6�/8 weeks

with at least two antipsychotics, one of which should

be an atypical antipsychotic (NICE 2002; APA

2004). Compliance should be ensured, if necessary

by checking drug concentrations.

Efficacy of first-generation antipsychotics. There is only

limited evidence for the efficacy of FGAs for TRS. A

review summarizing more than 100 trials comparing

two or more different FGAs concluded that only one

study found any FGA to be more effective than

another (Janicak et al. 1993). As a result, in terms of

efficacy, FGAs were considered interchangeable

(Conley und Buchanan 1997). In most controlled

trials of FGAs in patients with drug-resistant symp-

toms, the percentage of responders was fewer than

5% (Kane et al. 1988) or in the range of 3�/7%

(Kinon et al. 1993), changing the FGA once or twice

if response failed. So the primary reason for choos-

ing between drugs was to reduce side effects, provide

different dosing strategies, or offer different routes of

administration (Conley und Buchanan 1997). Based

on neurobiochemical considerations, data from

blocking 80�/90% of D2 receptors by 400 mg CPZ

equivalents (e.g., Farde et al. 1992) and clinical

impressions from acute treatment studies (e.g.,

Baldessarini et al. 1988; Bollini et al. 1994; Dixon

et al. 1995) suggested that higher dosages of FGA

produce no therapeutic benefit for TRS, but led to a

higher rate and severity of EPS and other disabling

side effects (Kane 1994; Moller 1996).

Efficacy of second-generation antipsychotics. A meta-

analysis summarising 12 RCTs comparing the

efficacy and tolerability of SGAs versus FGAs for

TRS found that there were more favourable out-

comes when treated with clozapine reflected by

superior improvement in overall psychopathology,

categorical response rate, EPS and adherence rate,

and when treated with olanzapine with regard to

categorical response and adherence rates (Chakos et

al. 2001).

The efficacy of clozapine for TRS could be under-

lined in a meta-analysis using Cochrane criteria

(Wahlbeck et al. 1999) and by evaluating all available

10 RCTs comparing clozapine to other antipsycho-

tic agents in another meta-analysis (NICE 2002). In

this review, six double-blind, randomised controlled

studies in TRS lasting 6�/52 weeks demonstrated

superior efficacy in improvement of global psycho-

pathology and global response rate using mean doses

of 176�/600 mg/day clozapine compared to chlor-

promazine (mean dose approximately 1200 mg/day)

and haloperidol (doses between 16�/28 mg/day)

(Kane et al. 1988, 2001; Kumra et al. 1996; Hong

et al. 1997; Buchanan et al. 1998; Rosenheck et al.

1997). In addition, an open, randomised, controlled

long-term study (2 years) revealed evidence for

superior global response rate with clozapine (mean

dose 496 mg/day) compared to standard FGAs

(mean dose 1386 mg/day CPZ equivalents) (Essock

et al. 1996b). Another systematic review of clozapine

in TRS including 18 prospective, controlled clinical

trials (15 double-blind), 23 prospective observa-

tional studies and nine retrospective observational

studies, found response to treatment in experimental

designs in 36% and in retrospective analysis up to

54%, whereas dropout rates ranged from 33 to 20%,

respectively (Brambilla et al. 2002). As a result of

this review, clozapine was concluded also to demon-

strate effectiveness and overall good tolerability in

the treatment of TRS under clinical practice. Ad-

ditionally superior effectiveness in adherence, quality

of life and participation in psychosocial therapies

was reported in further trials (Lieberman et al. 1994;

Alvarez et al. 1997; Ciaparelli et al. 2003). The time

course of improvement with clozapine was different

and ranged from 4 to 8 weeks (Conley et al. 1997;

Rosenheck et al. 1999), whereas in other trials

improvement was still observed after periods of 6�/

12 months (Spina et al. 2000). In summary, there is

good research-based evidence to recommend treat-

ment with clozapine for treatment-resistant schizo-

phrenic patients (Level B).

For risperidone (6 mg/day) compared to haloper-

idol (15 mg/day), significant superior improvement

in total psychopathology was only observed in the

first 4 weeks of an 8-week RCT (Wirshing et al.

1999). Predictors for response to risperidone were

predominant positive symptoms and EPS at study

entry. A 12-week study in patients with acute

exacerbations and TRS revealed better improvement

in negative and overall symptoms favouring risper-

idone 6 mg/day compared to haloperidol 20 mg/day

(Zhang et al. 2001). Additionally treatment-resistant

patients demonstrated superior improvement in

verbal memory function (Green et al. 1997) and in

perception of emotion (face recognition) (Kee et al.

1998) under treatment with risperidone compared

to haloperidol. In a double-blind 14-week RCT

comparing risperidone (mean dose 11.6 mg/day),

clozapine (mean dose 527 mg/day), olanzapine

(mean dose 30.4 mg/day) and haloperidol (mean

dose 25.7 mg/day) superior improvement in overall

psychopathology (PANSS total score) was observed

in the 3 SGAs, but considering EPS reduction as a

WFSBP Guidelines for Biological Treatment of Schizophrenia 163

covariate, significance was reached only for cloza-

pine and olanzapine (Volavka et al. 2002). Compar-

ing risperidone (3�/10 mg/day) to clozapine (150�/

400 mg/day), one RCT showed a similar reduction

in psychotic symptoms and no difference in EPS rate

(Bondolfi et al. 1998). Two other RCTs revelaed

superiority for clozapine in improvement of total

psychopathology, positive and depressive symptoms

(mean doses 5.9 mg/day risperidone vs. 404 mg/day

clozapine) (Breier et al. 1999), and improvement in

global impression, overall and positive symptoms

(mean doses 9.0 mg/day risperidone vs. 642 mg/day

clozapine) (Azorin et al. 2001). In open, prospective

cross-over studies 40�/80% of risperidone non-re-

sponders and solely 0�/15% of clozapine non-

responders improved by switching to the other agent

(Cavallaro et al. 1995, Still et al. 1996), or similar

efficacy was observed (Daniel et al. 1996, Konrad et

al. 2000). As a result of these trials there may be an

advantage for switching from risperidone to cloza-

pine in regard to response rate. Additionally, open,

comparative studies pointed towards a superior

efficacy of clozapine (e.g., Flynn et al. 1998;

Lindenmayer et al. 1998; Wahlbeck et al. 2002). In

summary, there is evidence for superiority of risper-

idone compared to haloperidol, but there may be

disadvantages compared to clozapine in the treat-

ment of therapy refractory schizophrenia (Level B).

A meta-analysis identified two RCTs comparing

olanzapine to FGAs in TRS and calculated an

approximately 1.7-higher probability of response

favouring olanzapine (Chakos et al. 2001). Patients

with olanzapine treatment suffered from significantly

less EPS. In an 8-week RCT, only slight superior

response (7 vs. 0%) was observed in olanzapine-

treated (25 mg/day) patients compared to chlorpro-

mazine (1200 mg/day) (Conley et al. 1998). Re-

analysis of an RCT focusing on TRS (defined as one

unsuccesful trial with FGA) revealed a higher

response rate (47 vs. 35%) for olanzapine (mean

dose 11.1 mg/day) compared to haloperidol (mean

dose 10.0 mg/day) (Breier and Hamilton 1999). In

this study, superior improvement in overall, positive,

negative and depressive symptoms was observed in

the olanzapine group. As mentioned earlier, olanza-

pine demonstrated superior improvement in total

psychopathology and negative symptoms compared

to haloperidol without confounding with EPS (Vo-

lavka et al. 2002), nevertheless clinical effects of all

assessed atypicals (clozapine, olanzapine, risperi-

done) differed only slightly from haloperidol in this

randomised, double-blind trial. In another RCT

compared to clozapine (mean dose 304 mg/day)

there was a similar improvement of overall symp-

toms with olanzapine treatment (mean dose 20.5

mg/day), while olanzapine patients revealed less

adverse effects (Tollefson et al. 2001). Switching

from clozapine to olanzapine (5�/25 mg/day) led to

response in more than 40% of the patients in

prospective studies (Henderson et al. 1998; Dossen-

bach et al. 2000), while in one study the greatest

reduction in psychotic symptoms was observed in

weeks 3�/6 (Dossenbach et al. 2000). Further switch

trials under naturalistic conditions, mainly from

risperidone, revealed similar results (Lindenmayer

et al. 2001, 2002; Rodriguez-Perez et al. 2002; Chiu

et al. 2003; Karagianis et al. 2003). In summary,

there is evidence for superiority of olanzapine

compared to haloperidol or chlorpromazine, and

limited evidence of similar efficacy compared to

clozapine in the treatment of therapy refractory

schizophrenia (Level B).

In an RCT lasting 12 weeks, patients with only

partial response to fluphenazine treatment (20 mg/

day) demonstrated a significantly higher response

rate (52 vs. 38%) to quetiapine (600 mg/day) than to

haloperidol (20 mg/day) (Emsley et al. 2000).

Switching from haloperidol, olanzapine or risperi-

done to quetiapine (mean dose 505 mg/day) revealed

clinical improvement in 69% of the patients (De

Nayer et al. 2003).

Treatment with aripiprazole (mean dose 28.8 mg/

day) compared to perphenazine (mean dose 39.1

mg/day) showed a similar response in TRS concern-

ing recent definition (unsuccessful trial with one

SGA and risperidone or olanzapine) in a randomised

double-blind trial (Ebrecht et al. 2004). Significant

improvement in total psychopathology was observed

in both treatment conditions. In an open trial

switching stable chronic patients to aripiprazole (30

mg/day), improvement in global impression, overall,

positive and negative symptoms was noted (Casey et

al. 2003).

Switching stable but symptomatic outpatients

from conventional antipsychotics, olanzapine or

risperidone to ziprasidone (mean dose 91 mg/day)

was found to be well tolerated and associated with

symptom improvement (Weiden et al. 2003).

In a double-blind RCT comparing zotepine (150�/

450 mg/day) and clozapine (150�/450 mg/day)

similar improvement in positive and negative symp-

toms, and some cognitive domains was observed

during 6-week treatment (Meyer-Lindenberg et al.

1997). Two open studies found at least moderate

global improvement under 1-year treatment with

zotepine (50�/500 mg/day), whereas response (at

least 20% decrease in BPRS total score) was noted

in nearly 80% of the patients occurring in the first 12

weeks (Harada et al. 1992).

In summary there is very limited evidence for

efficacy of aripiprazole, quetiapine, ziprasidone and

164 P. Falkai et al.

zotepine in the treatment of therapy refractory

schizophrenia (Level D).

General recommendations. The first step in the clinical

management of treatment-resistant schizophrenia

(TRS) is to establish that antipsychotic drugs have

been adequately tried in terms of dosage, duration

and adherence. Other causes of non-response should

be considered in the clinical assessment, such as co-

morbid substance misuse, poor treatment adher-

ence, the concurrent use of other prescribed medi-

cines, polypharmacy including pharmacokinetic and

pharmacodynamic interactions, physical illness and

poor social environment and support (DGPPN

1998; NICE 2002, McGorry et al. 2003; APA

2004). Especially when applying SGA (lower range)

it may be useful to consider the patient’s weight to

adjust the dose (Kane et al. 2003a). The fact that

there is some delay between initiation of treatment

and full clinical response complicates the evaluation

of treatment response, and sometimes symptoms can

continue to improve for up to 6 months. If the

patient shows partial response an expert consensus

agrees to extend the duration of the trial to 4�/10

weeks for the initially switched antipsychotic, and

approximately 5�/11 weeks for the second antipsy-

chotic prescribed (Kane et al. 2003a). Target symp-

toms have to be defined. If the target symptoms of

schizophrenia have been unresponsive to conven-

tional antipsychotics, treatment with an SGA should

be provided; based on available evidence to date

olanzapine or risperidone may be worth considering

as first line (NICE 2002) (Level B). In individuals

with clearly defined treatment resistance (including

at least one adequate treatment with SGA), cloza-

pine should be introduced as treatment of choice

because of clozapine’s superior efficacy in this regard

(Level B) (DGPPN 1998; NICE 2002; APA 2004).

For non-adherent patients, special efforts to enhance

the therapeutic alliance should be made including

psychotherapeutic and psychosocial interventions

e.g., adherence therapy, psychoeducation and family

interventions. Depot antipsychotics can be consid-

ered if there has been poor adherence, but clozapine

may also be preferable because the improving

psychopathology and the required monitoring often

enhances adherence (McGorry et al. 2003). Cogni-

tive behavioural therapy should be offered in con-

junction with clozapine to improve social

functioning, quality of life, and to reduce positive

and negative symptoms (see section Psychotherapy)

(Level C). Besides clozapine, there are limited

options for patients suffering from severe and

significant residual symptoms, even after antipsy-

chotic monotherapy has been optimised without

success. This may be the reason why various

augmentation strategies and polypharmacy are of-

fered, despite the fact that there is only limited or no

evidence supporting this approach. Augmentation

and combination strategies are reviewed and dis-

cussed below. Before switching to another agent,

expert consensus agrees on increasing the dose of the

current antipsychotic, unless side effects lead to

earlier switching (Kane et al. 2003).

Switching strategies

There are three main strategies for switching from

one agent to another: cross-titration (gradually

tapering off the dose of the first antipychotic while

gradually increasing the dose of the second), overlap

and taper (continuing the same dose of the first

antipsychotic while gradually increasing the second

to a therapeutic level and then tapering the first),

and abrupt change of the antipsychotics. Some

studies compared these strategies, especially the

two first mentioned switching modalities, and found

no difference in efficacy and tolerability (Kane et al.

2003a). Most guidelines prefer cross-titration, with

the exception of switching to clozapine (e.g., APA

2004). In this case tapering off of the first anti-

psychotic should be completed before introducing

clozapine due to potential haematological side ef-

fects. Nevertheless many experts prefer also cross-

titration, probably reflecting the need for relatively

slow titration of clozapine (Kane et al. 2003a).

Switching to clozapine

As previously discussed, a substantial number of

studies have consistently demonstrated that cloza-

pine is superior to conventional neuroleptic agents in

the treatment of refractory schizophrenia (Level B).

In this case no general dosage recommendations can

be given; however, although the mean dose in the

studies was 400�/500 mg clozapine daily, some

patients responded well to only 100�/200 mg/day

and others benefited from doses of up to 900 mg/

day. According to clinical practice a target dose of up

to 400 mg/day may be selected. After nonreponse,

despite continuing treatment over 4�/6 weeks and

obtaining sufficient blood levels of clozapine, the

next step should consist of increasing the dosage up

to 900 mg/day and closely monitoring for adverse

effects (Level D). If there is still no adequate

response augmentation or combination strategies

should be considered (Level D).

Switching to other SGAs

As mentioned, a few studies have produced evidence

of the superiority of SGAs other than clozapine,

mainly risperidone and olanzapine, compared with

WFSBP Guidelines for Biological Treatment of Schizophrenia 165

typical neuroleptic agents in treatment-refractory

schizophrenia (Level B). There is some evidence

from switch studies that switching from clozapine to

other SGAs, e.g., olanzapine in the case of inade-

quate response may provide benefits (Level D).

More future research is urgently needed to develop

evidence-based switching strategies, especially from

clozapine.

Combining antipsychotics

Despite the frequent combination of two and more

antipsychotics in clinical practice (e.g., in more than

50% of hospitalised patients), there are few rando-

mised controlled studies available evaluating the

efficacy of combining strategies (Freudenreich and

Goff 2002). In patients with TRS adding an anti-

psychotic to clozapine was mentioned as a reasonable

treatment option. To date, four double-blind, pla-

cebo-controlled randomised trials have been pub-

lished evaluating combination therapy with

clozapine. While combining clozapine with chlor-

promazine revealed no benefit (Potter et al. 1989),

adding sulpiride to clozapine led to a more than 20%

decrease in psychopathology (BPRS total score) and

superior improvement compared to placebo (Shiloh

et al. 1997). In addition to earlier open studies and

case reports, a recent double-blind RCT showed

significantly superior improvement of positive, ne-

gative and total psychopathology when adding risper-

idone for up to 6 mg/day to clozapine (mean dose

400 mg/day) (Josiassen et al. 2005). In contrast to

these results another RCT revealed a significantly

inferior improvement of positive symptoms by add-

ing risperidone compared to placebo (Yagcioglu et

al. 2005). In open trials pimozide (mean dose 4 mg/

day) combined to clozapine demonstrated improve-

ment in psychotic symptoms, but also increased or

changed side effects (Freudenreich and Goff 2002;

Miller and Craig 2002). It was discussed if an

increased serum level of clozapine may be respon-

sible for better outcome (e.g., in combination with

risperidone) (Tyson et al. 1995). The effect of

increasing clozapine levels could not be replicated

in another study over 4 weeks in TRS (Henderson

und Goff 1996). A significant decrease in psycho-

pathology was found without changes of clozapine

serum level. Further case series revealed evidence for

the efficacy of adding risperidone to clozapine

(McCarthy and Terkelsen 1995; Morea et al. 1999;

Raskin et al. 2000; Taylor et al. 2001). Case reports

and open label studies revealed success with the

combination of clozapine and olanzapine (Gupta et

al. 1998), clozapine with amisulpride (Allouche et al.

1994) and clozapine with ziprasidone (Kaye 2003).

In a case series, improvement of positive and

negative symptoms was reported in patients receiv-

ing olanzapine in combination with sulpiride (Raskin

et al. 2000). However, in a randomised controlled

study with small sample size of patients with TRS,

the addition of sulpiride (600 mg/day) to olanzapine

monotherapy over 8 weeks did not reveal relevant

benefits in positive and negative symptomatology,

but led to improvement of depressive symptoms

(Kotler et al. 2004). A small case series showed

efficacy of combining olanzapine and risperidone

(Lerner et al. 2000).

Recommendations. Overall there is only very limited

evidence for the efficacy of combining antipsychotics

in TRS. The combination of clozapine with risper-

idone, and clozapine with sulpiride may reflect the

best treatment options, based on successful RCT

(Level C). Nevertheless in clinical practice it may be

useful to select the antipsychotic with which the best

previous response could be obtained for further

combination (Level D). On a theoretical pharmaco-

dynamic basis, antipsychotics with different receptor

profiles may be selected for combination therapy

(e.g., clozapine combined with amisulpride) (Level

D). Last but not least clinicians should be aware of

potentiating side effects when combining two or

more antipsychotics and may select agents with

decreased interaction risk.

Augmentation strategies

Augmentation strategies have to be selected in

regard to the defined target symptoms in TRS.

Depending on the type of predominantly presented

residual symptom (e.g., aggressive behaviour, anxi-

ety, positive, negative, cognitive or mood symp-

toms), augmentation includes adding lithium,

anticonvulsants, benzodiazepines and b-blocking

agents, and in some exceptional cases undertaking

more or less experimental approaches with N-

methyl-d-aspartate (NMDA) receptor agonists

(e.g., d-serine, glycine, d-cycloserine) or cholinergic

agonists (APA 2004).

To avoid risking side effects and potential drug

interactions, effectiveness of current treatment with

the chosen adjunctive medications has to be eval-

uated at short time intervals, and adjunctive medica-

tions that do not produce clinical benefits should be

discontinued. Randomised controlled studies with

greater sample sizes of augmentation strategies for

TRS are still lacking, therefore recommendations are

based mainly on clinical experience (Freudenreich

and Goff 2002).

166 P. Falkai et al.

Mood stabilisers and Anticonvulsants. A meta-analysis

evaluating the efficacy of lithium in schizophrenic

patients included 20 randomised, controlled studies.

Eleven of these trials studied the efficacy of lithium

as add-on treatment to ongoing antipsychotic

medication (Leucht et al. 2004). Overall the combi-

nation strategy revealed better response, but super-

iority compared to antipsychotic monotherapy could

not be found consistently in all trials and disap-

peared if patients with mood symptoms were ex-

cluded. In small case series augmentation with

lithium showed improvement in TRS (Simhandl et

al. 1996).

As mentioned earlier, adding valproate to haloper-

idol, olanzapine and risperidone in randomised

controlled trials demonstrated inconsistently positive

effects only on different specific response aspects,

e.g., negative symptoms and global impression

(Wassef et al. 2000), hostility (Dose et al. 1998;

Citrome et al. 2004) and more rapid onset of action

(Casey et al. 2003a), while another study reported

no benefit in any outcome measure (Hesslinger

1999). In one small case series improvement in

patients with TRS was reported (Morinigo et al.

1989).

The addition of carbamazepine revealed contro-

versial results. In combination with haloperidol, a

worsening of symptoms was reported, possibly due

to pharmacokinetic reasons (decrease of haloperidol

plasma level) (Hesslinger 1999). A meta-analysis

concluded that there is no evidence for the efficacy

of combining carbamazepine with antipsychotics in

schizophrenia (Leucht et al. 2002). Overall in

the eight evaluated placebo-controlled trials signifi-

cantly better improvement was noted only in a

minority of patients in the active substance group,

and there was no superiority in regard to response

criteria (reduction above 50% of BPRS total score).

Patients augmented with carbamazepine showed less

EPS and similar tolerability compared to placebo. In

a small case series augmentation with carbamazepine

showed improvement in TRS (Simhandl et al.

1996).

Lamotrigine was observed, in case reports, to

reduce psychotic symptoms in combination with

clozapine (Dursun et al. 1999; Dursun and Deakin

2001; Saba et al. 2002). When lamotrigine was

added to risperidone or haloperidol, case reports

were controversial (Dursun and Deakin 2001; Ko-

livakis et al. 2004). No significant improvement was

observed in addition to olanzapine or flupenthixol

(Dursun and Deakin 2001). In two RCTs the

addition of lamotrigine (200 mg/day) versus placebo

to an ongoing clozapine treatment (Tiihonen et al.

2003), or to conventional antipsychotics, risperi-

done, olanzapine or clozapine in a dose of 400 mg/

day, was slightly effective (Kremer et al. 2004) in

reducing positive and general psychopathological

symptoms in treatment-resistant schizophrenic pa-

tients. In both trials lamotrigine was well tolerated.

In an open, non-randomised trial no beneficial

effect of topiramate added to an ongoing treatment

with clozapine, olanzapine, risperidone or flu-

penthixol was seen in treatment-resistant schizo-

phrenia (Dursun and Deakin 2001).

Antidepressive agents. When augmenting antidepres-

sive agents, the potential exacerbation of psychotic

symptoms due to increased adrenergic and dopami-

nergic transmission has to be considered (Siris et al.

2000). Beyond treating depressive symptoms (see

previous section), persisting negative symptomatology

is the target of adding antidepressive agents. It is

noteworthy that, in a recent review, fluvoxamine and

fluoxetine demonstrated improvement of negative

symptoms independent of the influence on depres-

sive mood (Silver 2003). In addition, e.g., a recent

double-blind, placebo-controlled study augmenting

mirtazapine 30 mg/day to ongoing clozapine treat-

ment showed significant reduction of negative

symptoms and total psychopathology in 8 weeks

(Zoccali et al. 2004). The efficacy in treating

schizophrenic patients with predominantly negative

symptoms is reported elsewhere in this guideline (see

Section 3.3.3). Other studies were mainly conducted

only in postpsychotic depression, and therefore no

recommendation for treatment-resistant schizophre-

nia could be extrapolated.

Benzodiazepines. Additional benzodiazepine treat-

ment consistently showed no superiority to antipsy-

chotic monotherapy in patients with TRS. Two

double-blind RCTs demonstrated superior improve-

ment in total psychopathology (Lingjaerde et al.

1979; Wolkowitz et al. 1992), while other trials

revealed no benefit (Holden et al. 1968; Hanlon et

al. 1970; Ruskin et al. 1979; Pato et al. 1989). In the

subsample of patients with catatonia the benzodia-

zepines oxazepam and clonazepam revealed efficacy

in acute catatonia, while for lorazepam no super-

iority could be observed in a placebo-controlled

RCT in chronic catatonic patients (Ungvari et al.

1999). In general, lorazepam may provide some

advantages for combination approaches, because

good absorption of the oral preparation and less

muscle relaxation than with other benzodiazepines

was observed (APA 2004).

b-Blocking agents. A systematic review and meta-

analysis concluded that b-blockers did not reveal

clear evidence for their use as adjunctive medication

in schizophrenia and also in patients with TRS

WFSBP Guidelines for Biological Treatment of Schizophrenia 167

(Cheine et al. 2004). A placebo-controlled RCT

demonstrated only slight, non-significant improve-

ment for the combination with nadolol (80�/120 mg/

day) over 3 weeks (Allan et al. 1996). In a 16-week,

randomised clinical trial, patients showed no sig-

nificant difference when augmented with proprano-

lol (up to 1920 mg/day) (Myers et al. 1981). Further

RCTs reported no clinical improvement with aug-

mentation of oxprenolol (160�/640 mg/day) (Karniol

and Portela 1982), or propranolol in the dose range

of 80�/640 mg/day (Pugh et al. 1983) or in a mean

dose of 450 mg/day (Yorkston et al. 1977) over 12-

week treatment.

Recommendations. In summary, there is very limited

evidence that augmentation with mood stabilisers or

anticonvulsants reveals benefits for patients with

TRS. Therefore, administration of valproic acid,

carbamazepine or lamotrigine should only be con-

sidered after other therapy options have been

exhausted. There is evidence that lamotrigine as an

adjunctive treatment, especially to clozapine, can

reduce schizophrenic psychopathology (Level C).

Adding valproate may be a therapy option if aggres-

sion and hostility is predominantly present (Level C),

and lithium may reveal benefits if depressive symp-

toms are predominant (Level D). Carbamazepine

should not be administered together with clozapine

due to a potential increased risk of haematological

side effects. In excited, anxious and catatonic

patients, the use of adjunctive benzodiazepines is

beneficial (Level C). Patients presenting with persist-

ing negative (Level C) or depressive symptoms (Level

B) may be treated with antidepressants.

Electroconvulsive therapy (ECT) in TRS

In patients with treatment-resistant schizophrenia

(TRS), reports and case series suggest that ECT

may augment response to FGAs (e.g., Konig and

Glatter-Gotz 1990; Sajatovic and Meltzer 1993;

Chanpattana et al. 1999) and increase therapeutic

benefit in combination with SGAs (e.g., Kupchik et

al. 2000; Hirose et al. 2001). Therefore the combi-

nation of neuroleptics and ECT may be recom-

mended (APA 2004) (Level C). Reviewing

randomised controlled studies (RCTs) there is only

one study using stringently defined criteria for TRS

comparing ECT with sham-ECT, both groups on

concurrent antipsychotics. In this study no signifi-

cant advantage for ECT was found (Thayran and

Adams 2004). Nevertheless another RCTexamining

continuation ECT for TRS revealed responder rates

for remission above 50% (Chanpattana et al. 1999).

There is only inconsistent evidence that mood

symptoms or a diagnosis of schizoaffective disorder

suggest a better response to ECT (e.g., Folstein et al.

1973; Dodwell and Goldberg 1989) (Level D). Some

reports suggest that greater benefits are observed in

patients with younger age (Chanpattana et al. 1999),

predominantly positive symptoms (Landmark et al.

1987), shorter illness and episode durations (50�/

70% for patients who have been ill for less than 1

year but less than 20% for patients who have been

continuously ill for more than 3 years) (e.g.,

Kalinowsky and Worthing 1943; Chanpattana et al.

1999; Thayran and Adams 2004), or fewer paranoid

or schizoid premorbid personality traits (Dodwell

and Goldberg 1989) (Level D). The available

evidence suggests that antipsychotic medications

should be continued during and after ECT when it

is used in the treatment of schizophrenia (APA

2004) (Level D). A review of case series stated that

the addition of ECT to clozapine for patients who

did not respond to either traditional medication or a

trial of clozapine alone had been reported as well

tolerated and effective in 67% of cases (Kupchik et

al. 2000) (Level D).

In summary there is only limited evidence for the

efficacy of ECT in TRS. Therefore, ECT may be a

treatment option in patients not responding to

clozapine or when a trial with clozapine is not

recommended, e.g., because of harmful side effects

(APA 2004) (Level C). The available evidence

suggests that antipsychotic medications should be

continued during and after ECT (APA 2004) (Level

D). After successful ECT treatment in TRS, main-

tenance ECT has to be considered (APA 2004)

(Level C).

Combining with psychotherapy

A review stated that cognitive behaviour therapy

techniques may have value in improving positive

symptoms and therefore effectiveness in treatment-

refractory schizophrenia (Cormac et al. 2004). In

addition, cognitive remediation may be useful as a

therapeutic strategy to reduce the severity of cogni-

tive deficits (Bark et al. 2003). For persistent positive

symptoms or disabling negative symptomatology, a

CBT programme may lead to superior improvement

compared to standard care or supportive counselling

(Lewis et al. 2002). In another RCT, cognitive

behavioural strategies (individual or group therapy),

including elements of family intervention, proved

efficacy in improvement of delusions and hallucina-

tions (Drury et al. 2000). To date, some other

randomised controlled trials are underway to assess

efficacy of CBT in acutely ill patients (Wiedemann

and Klingberg 2003). CBT was effective compared

to standard care in increasing adherence and treat-

ment (Kemp et al. 1998). In a meta-analysis it was

168 P. Falkai et al.

concluded that some RCTs showed efficacy in

symptom improvement for family intervention com-

pared to standard care (Falloon 2003). In summary,

especially CBTand family intervention are helpful as

adjuncts to antipsychotic treatment in patients with

TRS (Level B).

Treatment in special circumstances

Psychiatric comorbidity

Suicidal behaviour. Common estimates are that 10%

of people with schizophrenia will complete suicide,

and that the rate of suicide attempts is 2�/5-fold

higher (e.g., Siris 2001). Cohort studies have shown

that approximately 10% of patients with first-epi-

sode schizophrenia attempted suicide within 1 year,

whereby hallucinations and previous suicidal beha-

viour represented the greatest risk factors (Norden-

toft et al. 2002). Demographically associated with

suicidality in schizophrenia are young age, being

early in the course of the illness, male gender,

coming from a high socioeconomic family back-

ground, being highly intelligent, having high expec-

tations, not being married, lacking social supports,

having awareness of symptoms, and being recently

discharged from the hospital. Also associated are

reduced self-esteem, stigma, recent loss or stress,

hopelessness, isolation, treatment non-adherence

and substance abuse (e.g., Siris 2001). Clinically,

the most common correlates of suicidality in schizo-

phrenia are depressive symptoms and the depressive

syndrome, although severe psychotic and panic-like

symptoms may contribute as well (e.g., Siris 2001).

Suicide attempts by schizophrenic patients are more

frequently fatal, indicating that more violent meth-

ods are used (Beautrais 2001). Different genetic and

neurobiological factors are thought to influence

suicidal behaviour, perhaps independently of genetic

factors due to schizophrenia (Meltzer 2002). Suicide

is the leading cause of premature death among

patients with schizophrenia (e.g., Meltzer 2002).

Suicidal ideas or threats should be judged in the

context of a patient’s history as provided by the

patient and by relatives and the current therapist, if

they are available (APA 1997). There should be

close monitoring of vulnerable patients during times

of personal crisis, significant environmental changes,

or heightened distress or depression during the

course of illness. The frequency of outpatient visits

may need to be increased during vulnerable periods,

including recent discharge from the hospitals (APA

1997).

In some open and randomised controlled studies

comparing SGAs with FGAs, a protective effect may

be observed for the prevention of suicide. This effect

could not be demonstrated consistently (Keck et al.

2000; Tondo et al. 2001; Barak et al. 2004). There-

fore it was concluded to date that no significant

difference has emerged between atypical and typical

neuroleptic agents as substance groups in the pre-

vention of suicidal behaviour (Level C). If there is an

increased risk of suicide, clozapine has proved to be

superior to other neuroleptic agents in the reduction

of suicidal behaviour in randomised controlled and

open studies (Meltzer and Okayli 1995; Meltzer et

al. 2003; Wagstaff and Perry 2003) (Level B).

Recommendations. For patients at high risk for suicide

hospitalisation should be considered, and suicide

precautions should be instituted. It is important to

maximise the somatic treatment of psychosis and

depression and address the patient’s suicidality

directly, with an empathic and supportive approach

(APA 2004). Clozapine therapy should be consid-

ered if there is a significant and continuously

increased risk of suicide (Level B).

Depression and anxiety

Depressive symptoms may occur in all phases of

schizophrenia, e.g., prodromal phase, first episode,

during the early course and after remission, and

depression may contribute to the residual symptoms

of schizophrenia, whereby the proportion of patients

with schizophrenia who also manifest depression

ranges from 7 to 75% (Siris 2000). Depressive

symptoms have to be distinguished from side effects

of antipsychotic medications (including medication-

induced dysphoria, akinesia and akathisia), and the

primary negative symptoms of schizophrenia (APA

2004). If antipsychotic medication-induced dys-

phoria is suspected, then antipsychotic dose reduc-

tion may be effective, or alternatively a switch to an

antipsychotic with a lower risk of inducing extra-

pyramidal symptoms (e.g., SGA) may be consid-

ered. Some FGAs (e.g., thioridazine) (Dufresne et

al. 1993) and SGAs have been suggested to be

effective in treating depressive symptoms in schizo-

phrenia. SGAs are suggested to be superior in this

regard compared to FGA; however, evidence is

limited (e.g., Tollefson et al. 1998; Peukens et al.

2000; Moller 2005a) (see also Section 2.3.1.2. Efficacy

of SGA). Depressive symptoms in acute episodes

may improve parallel with psychosis due to anti-

psychotic treatment (Moller 2005b). Due to poten-

tial worsening of psychosis by antidepressive agents

during the acute phase, antidepressants are advo-

cated primarily as adjunctive treatment in the stable

phase of schizophrenia (Mulholland and Cooper

2000). Treatment with antidepressants added as an

adjunct to antipsychotics is indicated when the

WFSBP Guidelines for Biological Treatment of Schizophrenia 169

symptoms meet the syndromal criteria for major

depressive disorder or are severe and causing sig-

nificant distress (e.g., when accompanied by suicidal

ideation) or interfering with function (DGPPN

1998; APA 2004). Tricyclic antidepressants

(TCAs) have been primarily examined in the treat-

ment of postpsychotic depression (Siris et al. 2000)

(Level B). Other antidepressants, e.g., SSRIs and

dual reuptake inhibitors, have also been found to be

useful in the treatment of depression in schizophre-

nia (Siris 2000) (Level B). Nevertheless, one RCT

observed no significant advantage with sertraline

compared to placebo and demonstrated high pla-

cebo response (Addington et al. 2002). A small RCT

comparing sertraline and imipramine in postpsycho-

tic depression revealed comparable efficacy, but

more rapid onset with sertraline (Kirli and Caliskan

1998). However, very few studies have examined the

effects of antidepressants in patients treated with

SGA, making it difficult to evaluate the current

utility of adjunctive antidepressant agents. When

prescribed, antidepressants are used in the same

doses that are used for treatment of major depressive

disorder (APA 2004). There are, however, potential

pharmacokinetic interactions with certain antipsy-

chotic medications; for example, the SSRIs (such as

fluoxetine, paroxetine, and fluvoxamine) are inhibi-

tors of cytochrome P450 enzymes and thereby

increase antipsychotic plasma levels. Similarly, the

blood levels of some antidepressants may be elevated

by the concomitant administration of antipsychotic

medications.

Aggressive behaviour

In observational and retrospective cohort studies

violent behaviour occurred more frequently (nearly

4-fold) in those with schizophrenia than in persons

with no form of psychiatric disease (Angermeyer

2000). In a meta-analysis, younger age, higher

rehospitalisation rate, comorbid antisocial personal-

ity disorder and former involvement in criminal acts

were predictors of violence in schizophrenic people

living in the community (Bonta et al. 1998). Other

risk factors for violence in schizophrenia include

prior arrests, substance abuse, the presence of

hallucinations, delusions or bizarre behaviours, the

presence of neurological impairment, and being

male, poor, unskilled, uneducated and unmarried

(APA 1997). Often relatives, e.g., parents, are

frequent targets of violence when it occurs (Anger-

meyer 2000). Identifying risk factors for violence

and violent ideation is part of a standard psychiatric

evaluation. When a patient is found to pose a serious

threat to other people, the psychiatrist must exercise

his or her own best judgement, in accord with the

legal requirements of the jurisdiction, to protect

those people from foreseeable harm (APA 1997).

Effective management of aggression and assaultive-

ness in patients with schizophrenia can often be

achieved through behavioural treatment, limit-set-

ting and ‘talking down’ techniques (APA 2004).

Antipsychotic medications are the mainstay of man-

agement; anticonvulsants (e.g., Citrome et al. 2004),

lithium and high-dose b-blockers (e.g., propranolol,

pindolol) (e.g., Caspi et al. 2001) have also been

reported to provide some utility (Level C) (APA

2004). SGAs, e.g., risperidone (e.g., Chengappa et

al. 2000; Aleman and Kahn 2001) and especially

clozapine (e.g., Chengappa et al. 2002) reveal

efficacy in the treatment of hostility and aggressive

behaviour (Level C). The emergency management

has been already discussed in a previous section (see

Section 3.2. Treating agitation).

Substance use disorders

Substance abuse in individuals with schizophrenia is

very common and has risen to the most prevalent

comorbid psychiatric condition associated with schi-

zophrenia (Cuffel et al. 1993). Estimated life-time

prevalence rates for substance abuse in schizophre-

nia range from approximately 15 to 65% (Kovanas-

zay et al. 1997; Regier et al. 1990; Wobrock et al.

2004). Prevalence rates vary with different screening

instruments, evaluation settings (inpatient or out-

patient) and depend on different social and cultural

factors. In more recent studies the 6-month and

lifetime prevalence of substance abuse or depen-

dence among people with schizophrenia was found

to be ca. 27 and 60%, respectively (Fowler et al.

1998), whereas comorbid non-alcohol lifetime sub-

stance misuse was reported in 16% of the overall

population, and cannabis abuse was reported in 60%

of male schizophrenics under the age of 36 years

(Duke et al. 2001). Comorbid substance use dis-

order (SUD) has been associated with more frequent

and longer periods of hospitalisation and other

negative outcomes, including higher relapse rate,

even in first-episode patients, higher non-compli-

cance, elevated EPS-rate in general and during

antipsychotic treatment, unemployment, homeless-

ness, violence, incarceration, suicide and HIV infec-

tion (e.g., Mueser et al. 1990; Olivera et al. 1990;

Soyka et al. 1993; Linszen et al. 1994; Blanchard et

al. 2000; Mueser et al. 2000; Goldham et al. 2002;

Hunt et al. 2002; Lacro et al. 2002). Besides the

legal substances tobacco and alcohol, cannabis

seems to be the most illicit drug abused in schizo-

phrenics and has been discussed as an important risk

factor for developing schizophrenia (Bersani et al.

2002; Caspari and Wobrock 2004). The presence of

170 P. Falkai et al.

substance abuse or dependence is often missed in

assessments, especially if such a patient is seen

during an acute psychotic episode. Because self-

report may often be unreliable, corroborating evi-

dence from all sources including laboratory tests

(liver function) and drug screening (urine and

blood) should be sought. The effects of abused

substances on schizophrenic symptoms vary, making

the differentiation of substance abuse-related symp-

toms from those related to functional psychosis

difficult (APA 2004). The abused substances may

lead to increased hallucinations, paranoid symptoms

or anxiousness in patients with pre-existing schizo-

phrenic psychoses (e.g., Dixon et al. 1991). Anti-

psychotics may not necessarily neutralise these

psychomimetic effects (e.g., Gawin 1986). In some

cases it may be extremely difficult to distinguish

between schizophrenia and drug-induced psychosis.

The key issue in providing treatment for this

population is developing a dual disorder approach

that integrates treatment of substance abuse and

schizophrenia (Ridgely and Jerrel 1996; Drake and

Mueser 2000; Wobrock et al. 2004). Many pro-

grammes are now providing this integration through

interdisciplinary teams with expertise in the treat-

ment of schizophrenia and substance abuse. This

form of treatment features assertive outreach, case

management, family interventions, housing, rehabi-

litation and pharmacotherapy. It also includes a

stagewise motivational approach for patients who

do not recognise the need for treatment of substance

use disorders, and behavioural interventions for

those who are trying to attain or maintain absti-

nence. Studies show that combined treatment pro-

grammes with motivational elements, psycho-

education and cognitive-behavioural approaches

avoiding direct confrontation can be effective in

reducing substance abuse and in decreasing fre-

quency and severity of psychotic decompensations

(e.g., Hellerstein et al. 1995; Addington and El-

Guebaly 1998; Drake and Mueser 2000; Barrow-

clough et al. 2001; Clark 2001; Baker et al. 2002).

Although prospective randomised controlled stu-

dies for schizophrenic patients with comorbid sub-

stance use are lacking, recommendations for

antipsychotic treatment are mostly derived from

studies in schizophrenia where drug abuse was an

exclusion criteria. In older reviews, haloperidol at a

dose of 5�/10 mg/day (maximal 20 mg/day), or

alternatively flupentixol 5�/20 mg/day, were recom-

mended as drugs of choice for antipsychotic treat-

ment (Soyka 1996; Wilkins 1997). Antipsychotics

with extensive anticholinergic side effects should be

avoided since anticholinergic effects of used sub-

stances could be potentiated. Additional positive

symptoms caused by substances usually remit

quickly with abstinence; hence, schizophrenic pa-

tients with concomitant substance use do not require

higher doses of antipsychotics than schizophrenic

subjects without comorbid substance use (Richard et

al. 1985; Siris 1990; Wilkins 1997). In schizophrenic

patients using cocaine, a possible intensification of

cocaine-induced hyperthermia must be taken into

account during antipsychotic treatment (Kosten and

Kleber 1988). Prolonged absorption, e.g., of canna-

bis from visceral fat tissue, may be responsible for

prolonged psychomimetic effects and pharmacoki-

netic interactions, especially in the initial stages of

therapy. Alcohol, for example, has been reported to

reduce the serum levels of antipsychotic agents (e.g.,

fluphenazine) (Soni et al. 1991).

Open trials with FGAs , in particular flupentixol ,

and case series suggest an effectiveness in the

reduction of comorbid substance use. In schizophre-

nic patients with comorbid alcohol dependence,

alcohol consumption and craving was markedly

reduced by flupentixol decanoate (10�/60 mg i.m.

every 2 weeks), while psychopathology was hardly

affected (Soyka and Sand 1995; Soyka et al. 2003).

In addition to an improvement in mental state,

schizophrenic patients with comorbid substance

dependence (alcohol, cocaine, marijuana, benzodia-

zepines and amphetamines) receiving flupentixol

decanoate displayed a reduction in substance intake

(Schilkrut et al. 1988; Levin et al. 1998). In contrast

to these findings, treatment with flupentixol decan-

oate (compared to placebo) was associated with

more relapses in alcoholics without comorbid schi-

zophrenia (Wiesbeck et al. 2001). In a cross-

sectional study, treatment with SGAs compared

with conventional antipsychotic agents was asso-

ciated with a reduction in alcohol consumption,

although no differences were seen in psychopathol-

ogy or extrapyramidal side effects (Scheller-Gilkey et

al. 2003).

Of SGAs the largest body of evidence has been

gathered for clozapine in the treatment of schizo-

phrenic patients with comorbid substance use. In

case records, case series, retrospective analysis of

treatment-resistant patients and cross-sectional stu-

dies, reduced substance consumption (particularly

alcohol and cocaine) associated with clozapine

treatment has been reported. Reduced substance

use was associated with improvement in (particularly

negative) schizophrenic symptoms in most case

series (Albanese et al. 1994; Yovell & Opler, 1994;

Marcus and Snyder 1995; Tsuang et al. 1999) and

retrospective or cross-sectional studies (Lee et al.

1998; Drake and Muser 2000; Zimmet et al. 2000),

but not in all retrospective analysis (Buckley et al.

1994a,b). Controlled studies for maintenance ther-

apy and relapse prevention in patients with this dual

WFSBP Guidelines for Biological Treatment of Schizophrenia 171

diagnosis are not available. In clozapine-treated

patients with treatment-resistant schizophrenia, no

difference was found in the comparison of patients

with and without comorbid substance use regarding

readmission rate in a prospective observational trial

(Kelly et al. 2003). In cocaine addicts, administra-

tion of clozapine can increase serum concentrations

of cocaine and induce cardiovascular side effects,

therefore close monitoring during the dose titration

phase of clozapine is indicated (Farren et al. 2000).

For patients showing persistent psychotic symptoms

or developing side effects to classical antipsychotics,

treatment with clozapine at doses of 50�/600 mg/day

is recommended (Soyka 1996; Wilkins 1997). In

addition, retrospective, cross-sectional and prospec-

tive studies (switching from haloperidol to cloza-

pine) revealed evidence for the reduction of cigarette

smoking in schizophrenic patients (McEvoy et al.

1995; Procyshyn et al. 2001, 2002). For risperidone ,

case records and an open prospective study with

small sample size revealed evidence for reduction of

substance use, craving, symptomatology and relapse

in comparison with typical antipsychotics (haloper-

idol, fluphenazine, chlorpromazine) (Gupta and

Basu 2001; Smelson et al. 2002). In a prospective

pilot study with small sample size comparing olan-

zapine with haloperidol, reduction in cocaine con-

sumption parallel to an improvement in psycho-

pathological condition and adverse drug effects was

observed (Tsuang et al. 2002). Open prospective

clinical trials showed that switching antipsychotic

treatment from conventional antipsychotic drugs to

olanzapine could decrease substance use and im-

prove schizophrenic symptoms (Noordsy and

O’Keefe 1999; Littrell et al. 2001). Substantial

weight gain was noted as an adverse effect. One

open prospective study demonstrated that olanza-

pine at doses of 10�/25 mg/day is as effective on

psychiatric symptoms in treatment-resistant schizo-

phrenic patients without comorbid substance use as

in those with additional substance use (Conley et al.

1998). An open prospective trial and case reports

suggested that quetiapine may be successful in

reducing alcohol and cocaine craving, relapse and

symptom improvement in patients with comorbid

substance use (Brown et al. 2002; Weisman 2003).

In a retrospective study comparing the efficacy of

clozapine and riperidone on alcohol and cannabis use

in schizophrenic patients, treatment with clozapine

led to more abstinence (Green et al. 2003).

Since depressive symptoms are frequent in pa-

tients with schizophrenia and comorbid substance

use, studies with tricyclic antidepressants (TCAs)

were undertaken. Besides the antidepressant effect

a decrease of substance consumption and craving

could be detected in cocaine-dependent schizophre-

nic patients. This observation was made either in an

open study (Ziedonis et al. 1992) or in a placebo-

controlled study administering desipramine (Wilkins

et al. 1997) or imipramine concurrently to neuro-

leptic treatment (Siris et al. 1993). Because of their

anticholinergic effects, treatment with TCAs should

not be started until drug-detoxification is completed.

Other adverse effects of treatment with TCAs may

be the precipitation of hypertensive crises during

concomitant use of substances with adrenergic

stimulation and the exacerbation of psychotic symp-

toms (Siris 1990).

For anticraving agents , e.g., acamprosate (modu-

lating NMDA receptors) and naltrexone (opioid

receptor antagonist), there are no controlled studies

available for patients with schizophrenia and drug

dependencies. Naltrexone has been reported in

research abstracts to be potentially helpful (Sernyak

et al. 1998; Maxwell and Shinderman 2000). Never-

theless, because of their excellent tolerability these

medications may be administered to alcohol-depen-

dent schizophrenic patients (Noordsy and Green

2003). Disulfiram may itself induce psychosis, pre-

sumably by blocking dopamine b-hydroxlyase (Ew-

ing et al. 1977; Major et al. 1978; Noordsy and

Green 2003). Therefore, and due to the possible

acceleration of antipsychotic drug metabolism, dis-

cussion is controversial regarding the use of dis-

ulfiram in patients with schizophrenia and comorbid

alcohol dependence, although therapeutic success in

doses 250�/500 mg/day has been reported in case

records and chart reviews (Kofoed et al. 1986;

Mueser et al. 2003).

Since the frequency of EPS is increased in

schizophrenic patients with comorbid substance

use disorder, concomitant or early administration

of anticholinergic agents, for instance biperiden,

could be considered (Soyka 1996). The higher risk

for EPS in this population could argue for SGAs as

the first-line treatment option in schizophrenic

patients with comorbid substance use. The benza-

mide antipsychotic tiapride could become the treat-

ment of choice for dyskinesias, as tiapride may have

anticraving properties and is used in the treatment of

alcohol withdrawal syndromes (Soyka 1996).

Based on very limited evidence but supported by

theoretical neuropharmacological considerations

(e.g., lower EPS, ameliorating brain reward dysfunc-

tion, reducing impulsivity) and growing clinical

practice, SGAs may be recommended as first-line

treatment in patients with schizophrenia and comor-

bid substance use disorder (Krystal et al. 1999;

Noordsy and Green 2003; Potvin et al. 2003) (Level

D). It may be favourable for patients with this dual

diagnosis to lower the threshold for recommending a

trial with clozapine than in patients without comor-

172 P. Falkai et al.

bid substance use disorder (Level D). Because of

frequent non-adherence, extended use of depot

antipsychotics such as haloperidol or flupentixolde-

canoate may be a favourable therapeutic option for

this patient group (Soyka 1996) (Level D).

Somatic comorbidity

Epidemiological surveys demonstrate that patients

with schizophrenia suffer to a higher extent from a

variety of somatic comorbidities, including cardio-

vascular disease, respiratory disease, diabetes and

infectious diseases. Somatic comorbidities are deter-

mined by multiple factors, including associations

with schizophrenia itself, lifestyle (e.g., smoking,

substance use, obesity, lack of exercise), environ-

ment and medications (e.g., Jeste et al. 1996).

Therefore treatment selection and clinical manage-

ment of patients with schizophrenia must consider

the patient’s past medical history and general

medical status when determining the treatment

plan (APA 2004). Because patients with psychosis

and mental retardation are at increased risk for

extrapyramidal side effects and tardive dyskinesia,

second-generation antipsychotics, and particularly

those with minimal risk of EPS (e.g., quetiapine),

are recommended (APA 2004). For patients with

preexisting osteopenia or osteoporosis , an antipsycho-

tic with minimal effects on prolactin should be

prescribed if possible. In addition, for women with

breast cancer , antipsychotics with prolactin-elevating

effects should be avoided or prescribed only after

consultation with the patient’s oncologist. In such

instances, aripiprazole, which partially suppresses

prolactin release, may be specifically indicated (APA

2004). For obese patients special treatment consid-

erations may be useful (see the section on management

of side effects in Part 2 of these guidelines). Treatment

selection should always weigh the expected benefits

of antipsychotic treatment against its potential to

exacerbate or contribute to the development of

specific medical conditions. Patients with prolonged

QT syndrome, bradycardia, certain electrolyte dis-

turbances, heart failure or recent myocardial infarc-

tion and patients who are taking drugs that prolong

the QT interval should not be treated with an

antipsychotic that could further prolong the QT

interval or increase the risk of the arrhythmia

torsades de pointes. These antipsychotics include

thioridazine, pimozide, droperidol and ziprasidone

(Marder et al. 2004). Medications with low affinity

for a-adrenergic receptors should be used for pa-

tients who are vulnerable to orthostatic hypotension,

including elderly patients, patients with peripheral

vascular disease or compromised cardiovascular

status, and other severely debilitated patients (APA

2004). For patients with acute angle-closure glau-

coma, severe constipation, history of a paralytic

ileus, urinary retention, prostate hypertrophy or

delirium/dementia, antipsychotics with little or no

antagonism for cholinergic receptors should be

prescribed. Studies demonstrated evidence that

patients with severe dementia may be at increased

risk of stroke when treated with risperidone or

olanzapine compared to placebo. Nevertheless this

elevated risk may also exist for FGAs or other SGAs,

but has not been systematically investigated. Cloza-

pine should not be used in patients with haemato-

logical diseases, especially when presenting with

neutropenia (B/1500/mm3) or low white blood cell

(WBC) count (B/3000/mm3) or a history of such

sensitivities to prior medications (e.g., chlorprothix-

ene, mianserine). Patients with hepatic disease may

have impaired metabolism of antipsychotic medica-

tions and are at risk for toxicity. It is important to

note that all antipsychotic medications lower the

seizure threshold, with increased risks for chlorpro-

mazine, clozapine and zotepine.

Elderly patients

Among middle-aged and elderly persons with schi-

zophrenia, approximately 20% have late-onset schi-

zophrenia (onset after age 40) or very-late-onset

schizophrenia-like psychosis (onset after age 60)

(APA 2004). The approach to the treatment of older

persons with schizophrenia is similar to that for

younger patients and involves combining pharma-

cotherapies with psychosocial interventions (NICE

2002; APA 2004). Several age-related physiological

changes, e.g., reduced cardiac output, reduced

glomerular filtration rate, possible reduction in

hepatic metabolism, and increased fat content alter

absorption, distribution, metabolism and excretion

of medications. Compared to younger patients,

geriatric patients show greater variability of response

and greater sensitivity to medications. While an

exaggerated response is more common in the elderly,

some patients manifest diminished idiosyncratic,

and even paradoxical effects of medications. In

general, recommended starting doses in older pa-

tients are one-quarter to one-half of the usual adult

starting dose (APA 2004). Elderly patients tend to

be more sensitive to the therapeutic and toxic effects

of antipsychotic medications, partly because of age-

related decreases in dopamine and acetylcholine

neurotransmission in the brain. This sensitivity is

especially higher in older patients with structural

brain abnormalities. Side effects of antipsychotic

medications that occur more frequently in older

subjects include sedation, orthostatic hypotension,

anticholinergic reactions, extrapyramidal symptoms

WFSBP Guidelines for Biological Treatment of Schizophrenia 173

(akathisia and parkinsonism but not acute dystonia),

and tardive dyskinesia (e.g., Jeste 2000). Because

nearly all the commonly prescribed antipsychotic

medications are equally efficacious for older patients

with schizophrenia, selection of an antipsychotic

medication should be based primarily on the side

effect profile (for concomitant somatic comorbidity see

previous section in this guideline) (APA 2004). In six

double-blind controlled studies that included schi-

zophrenic patients with a mean age of more than 45

years, the mean rate of relapse for the groups whose

antipsychotic medication was withdrawn was 39.9%

over an average 6-month follow-up period, while the

relapse rate for the groups whose antipsychotic

medication was maintained was 11.4% (Jeste et al.

1993). This result may suggest to continue with

medication. Especially in elderly people, unneces-

sary polypharmacy has to be avoided, and antic-

holinergic side effects of antipsychotic drugs become

more prominent and important. This should be

taken into consideration when prescribing an anti-

psychotic or adjunctive agent.

Gender differences

In the clinical presentation and course of schizo-

phrenia, numerous gender differences are obvious

(e.g., Goldstein and Tsuang 1990; Hafner 2000).

The mean age of onset in women is later (3�/4 years)

than in men and the peak of illness onset is two-

tailed, with a second peak after menopause (Hafner

2000). Probably due to the later onset, a better

premorbid history and potential protective effects of

oestrogens, women present a favourable course with

less negative and more affective symptoms, and are

more likely to show response to neuroleptics than

men with schizophrenia (Seeman 1986; Goldstein

and Tsuang 1990; Castle et al. 1993, Hafner 2000).

Other than biologically mediated factors including

family and societal expectations, lower pre- and

postadmission levels of substance abuse and better

support from family members may also positively

affect outcome. In addition, traditional socialisation

practises may allow greater dependence on the

family and greater acceptance of family treatment

among female schizophrenic patients. Also, it has

been observed that, even after body weight is

considered, women with first-episode schizophrenia

responded to lower doses of antipsychotic medica-

tion than men (Szymanski et al. 1995; Leung and

Chue 2000), although there is evidence that post-

menopausal women may require higher doses

(Leung and Chue 2000). Women tended to develop

more EPS and higher prolactin levels during anti-

psychotic treatment than men (Leung and Chue

2000). Due to hormonal changes, women may need

reduced dosage of neuroleptics in the middle of the

menstrual cycle and increase of dosage during

menstruation (Castle et al. 1995). Women with

schizophrenia have a higher risk of physical and

sexual abuse than healthy women and men, there-

fore careful exploration and special help may be

required (Goodman et al. 1997).

In summary, there is limited evidence that women

may be treated with lower doses of antispychotic

medications than men (Level D). Additional infor-

mation on contraception, protection against physical

and sexual violence and potential gender-specific

risks of antipsychotics (e.g., osteoporosis, breast

cancer) should be given (Level D).

Pregnancy and lactation

Antipsychotic treatment of the pregnant or lactating

women with schizophrenia has to balance the risks of

various psychotropic medications to the foetus, new-

born and breast-fed infant, against untreated psy-

chotic symptoms and inadequate prenatal care (APA

2004). Controlled studies of antipsychotic drugs

during pregnancy are not available, for obvious

ethical reasons. Therefore knowledge of the risks of

these agents arises from animal studies and from

uncontrolled exposures in humans. Two periods of

high risk to the foetus or newborn are identifiable�/

the first trimester with highest teratogenic risk and

the time of birth with the highest withdrawal risk

(APA 2004). Psychotropic drug exposure in the first

trimester is only under full control of the physician

and patient in planned pregnancies. Concerning

FGAs , there is relatively little evidence of harmful

effects, especially with high-potency agents, related

to their former widespread use (Cohen and Rosen-

baum 1998; American Academy of Pediatrics 2000;

Gold 2000). Reports of the use of psychotropic

medications during pregnancy had focused mainly

on chlorpromazine and haloperidol. Less informa-

tion is available regarding foetal exposure to SGAs .

In a review, spontaneous abortion, stillbirth, pre-

maturity, foetal abnormalities, unwanted perinatal

reactions, gestational complications, including new-

onset diabetes and preeclampsia, were reported

under treatment with olanzapine (Ernst and Gold-

berg 2002). There were reports of elective termina-

tions, new-onset or worsening gestational diabetes,

floppy infant syndrome, neonatal seizures, gastro-

oesophageal reflux disease and one intrauterine

foetal death under treatment with clozapine (Gentile

2004). Elective terminations (MacKay et al. 1998)

and one case of corpus callosum agenesis has been

observed with risperidone treatment (Gentile 2004).

Two case reports treating a pregnant schizophrenic

patient with quetiapine revealed no complications

174 P. Falkai et al.

(Tenyi et al. 2002; Taylor et al. 2003). Case reports

are lacking for aripiprazole and ziprasidone, while

possible teratogenic effects have been described in

animal studies (Gentile 2004). In a sytematic review

it was concluded that olanzapine and clozapine do

not increase the teratogenic risk compared to out-

comes in the general healthy population; knowledge

about quetiapine, risperidone, aripiprazole and zi-

prasidone is limited or lacking (Gentile 2004).

Treatment with olanzapine and clozapine may be

associated with a greater risk of hyperglycemia

(Gentile 2004). In addition, pregnant women with

schizophrenia taking SGAs were found frequently to

be obese and have an inadequate intake of folic acid

(related to an increased risk of neural tube defects),

even suggesting indirect effects rather than a direct

medication effect (Koren et al. 2002). Further, a

number of studies demonstrated that pregnant

women with schizophrenia receive relatively poor

prenatal care, have more obstetric complications,

and their offspring are more likely to have low birth

weight and stillbirth (APA 2004). Contributing

factors may be, e.g., low socioeconomic status and

high rates of smoking or substance use disorders.

Compared with antipsychotic medications, mood

stabilisers and benzodiazepines are much more

closely associated with foetal malformations and

behavioural effects, the teratogenic effect of sodium

valproate is especially well known (Ernst and Gold-

berg 2002; American Academy of Pediatrics 2000;

Gold 2000). Thus, their risk/benefit ratio is differ-

ent, and the need for their continuation during

pregnancy and breast-feeding requires strong clinical

justification.

Based on the reported findings and considerations

it is recommended to insist on early involvement of

an obstetrician who can help reduce the risks of the

pregnancy and with whom the risks and benefits of

pharmacological treatment options can be discussed.

When possible, non-pharmacological management

of schizophrenia should be undertaken during preg-

nancy (RANZCP 2003), although there nevertheless

may be more risk to the mother�/infant dyad from

psychosis than benefit from stopping medication

(Working Group for the Canadian Psychiatric Asso-

ciation 1998). If antipsychotic treatment is neces-

sary, the minimum effective dose should be used and

special attention paid to lowering the dose during the

month before delivery or terminating 5�/10 days

before anticipated delivery (APA 1997; Working

Group for the Canadian Psychiatric Association

1998). Because there is still more experience with

FGAs in pregnancy, no advantage of SGAs in this

condition has been demonstrated (Gentile 2004).

Some high-potency agents appear to be safe (e.g.,

haloperidol), and they should be used preferentially

but as briefly as possible, and the dose should be low

enough to avoid EPS and therefore the necessity of

antiparkinsonian medications (APA 1997; DGPPN

1998; Working Group for the Canadian Psychiatric

Association 1998). Antiparkinson medication should

especially be avoided in the first trimester (APA

1997), and treatment with phenothiazines should

not be conducted (DGPPN 1998). After delivery,

resumption of the full dose of FGA or SGA has to be

considered, because there may be an increased risk

of psychosis postpartum (APA 1997; Working

Group for the Canadian Psychiatric Association

1998).

The efficacy of ECT during pregnancy in schizo-

phrenia is assumed to be similar to that in non-

pregnant women, while randomised evidence is

lacking (UK ECT Group 2002). The rate of

complications based on over 300 case reports and

summarised in reviews (Miller 1994; UK ECT

Group 2002) tended to be low (around 1%), and

included four cases of premature labour, five cases of

miscarriage, five cases of congenital abnormalities

and three cases of still birth or neonatal death.

Therefore ECT is recommended as a possible

treatment option in several guidelines if pharmaco-

logical antipsychotic treatment is not appropiate

(APA 1997; DGPPN 1998).

Because antipsychotics can accumulate in breast

milk, especially observed with clozapine medication,

and based on only sparse literature reports, breast

feeding could not be recommended for women taking

antipsychotic drugs (Gentile 2004). There is con-

sensus among different guidelines to avoid breast

feeding during psychotropic treatment of the mother

(APA 1997; DGPPN 1998; Working Group for the

Canadian Psychiatric Association 1998).

Early intervention in the initial prodromal

phase

Subthreshold psychotic features combined with the

onset of disability, especially if there is a family

history, indicate a very high risk to develop psychosis

(McGorry et al. 2003). Therefore persons with high

risk in their family should be actively engaged in

assessment and regular monitoring of mental state

and safety. This should be carried out in a home-,

primary care- or office-based setting, if possible, to

reduce stigma (McGorry et al. 2003). Concurrent

syndromes such as depression and substance abuse,

and problem areas such as interpersonal vocational

and family stress, should be appropriately managed.

Information about the level of risk should be care-

fully provided, conveying a sense of therapeutic

optimism. It should be emphasised that current

problems can be alleviated, progression to psychosis

WFSBP Guidelines for Biological Treatment of Schizophrenia 175

is not inevitable, and if psychosis does occur then

effective and well-tolerated treatments are readily

available. Engagement at this early stage will help to

reduce any subsequent delay in accessing treatment

for first-episode psychosis. The use of antipsychotic

medication during the prodrome is the subject of

research. At present it should be reserved for

patients who display clear psychotic symptoms

(Level D) (McGorry et al. 2003).

Acknowledgements

We would like to thank Jacqueline Klesing, Munich,

for general and editorial assistance in preparing these

guidelines

References

Adityanjee, Aderibigbe YA, Mathews T. 1999. Epidemiology of

neuroleptic malignant syndrome. Clin Neuropharmacol

22:151�/158.

Addington J, El-Guebaly N. 1998. Group treatment for sub-

stance-abuse in schizophrenia. Can J Psychiatry 43:843�/845.

Addington D, Addington J, Patten S, Remington G, Moamai J,

Labelle A, Beauclair L. 2002. Double-blind, placebo-con-

trolled comparison of the efficacy of sertraline as treatment

for a major depressive episode in patients with remitted

schizophrenia. J Clin Psychopharmacol 22:20�/25.

Aguilar EJ, Keshavan MS, Martinez-Quiles MD, Hernandez J,

Gomez-Beneyto M, Schooler NR. 1994. Predictors of acute

dystonia in first-episode psychotic patients. Am J Psychiatry

151:1819�/1821.

Aizenberg D, Zemishlany Z, Etrog PD, Weizman A. 1995. Sexual

dysfunction in male schizophrenic patients. J Clin Psychiatry

56:137�/141.

Akhondzadeh S, Nejatisafa AA, Amini H, Mohammadi MR,

Larijani B, Kashani L, Raisi F, Kamalipour A. 2003. Adjunc-

tive estrogen treatment in women with chronic schizophrenia:

A double-blind, randomized, and placebo-controlled trial. Prog

Neuropsychopharmacol Biol Psychiatry 27(6):1007�/1012.

Albanese MJ, Khantzian EJ, Murphy SL, Green AI. 1994.

Decreased substance use in chronically psychotic patients

treated with clozapine. Am J Psychiatry 151:780�/781 (letter).

Albers LJ, Ozdemir V. 2004. Pharmacogenomic-guided rational

therapeutic drug monitoring: conceptual framework and appli-

cation platforms for atypical antipychotics. Curr Med Chem

11(3):297�/312.

Al Khatib SM, LaPointe NM, Kramer JM, Califf RM. 2003.

What clinicians should know about the QT interval. J Am Med

Assoc 289:2120�/2127.

Allan ER, Alpert M, Sison CE, Citrome L, Laury G, Berman I.

1996. Adjunctive nadolol in the treatment of acutely aggressive

schizophrenic patients. J Clin Psychiatry 57:455�/459.

Aleman A, Kahn RS. 2001. Effects ot the atypical antipsychotic

risperidone on hostility and aggression in schizophrenia: A

meta-analysis of controlled trials. Eur Neuropsychopharmacol

11:289�/293.

Allison DB, Mentore JL, Heo M, Chandler LP, Cappelleri JC,

Infante MC, Weiden PJ. 1999. Antipsychotic-induced weight

gain: a comprehensive research synthesis. Am J Psychiatry

156:1686�/1696.

Allison DB, Mackell JA, McDonnell DD. 2003. The impact of

weight gain on quality of life among persons with schizophre-

nia. Psychiatr Serv 54:565�/567.

Altamura AC, Mauri MC, Mantero M, Brunetti M. 1987.

Clonazepam/haloperidol combination therapy in schizophre-

nia: A double-blind study. Acta Pychiatr Scand 76:702�/706.

Altamura AC, Sassella F, Santini A, Montresor C, Fumagalli S,

Mundo E. 2003. Intramuscular preparations of antipsychotics:

Uses and relevance in clinical practice. Drugs 63(5):493�/512.

Alvarez E, Baron J, Puigdemont JSD, Masip C, Perez-Sola V.

1997. Ten years’ experience with clozapine in treatment-

resistant schizophrenic patients: factors indicating the thera-

peutic response. Eur Psychiatry 12(suppl 5):343�/346.

American Academy of Pediatrics. 2000. Use of psychoactive

medication during pregnancy and possible effects on the fetus

and newborn. Pediatrics 105:880�/887.

American Diabetes Association; American Psychiatric Associa-

tion; American Association of Clinical Endocrinologists; North

American Association for the Study of Obesity. 2004. Con-

sensus development conference on antipsychotic drugs and

obesity and diabetes. J Clin Psychiatry 65(2):267�/272.

American Psychiatric Association. 1997. Practice guideline for the

treatment of patients with schizophrenia. Am J Psychiatry

154(Suppl 4):1�/63.

American Psychiatric Association. 2001. The practice of electro-

convulsive therapy: Recommendations for treatment, training,

and privileging: A Task Force Report of the American

Psychiatric Association. Washington, DC: American Psychia-

tric Association.

American Psychiatric Association. 2004. Practice guideline for

the treatment of patients with schizophrenia. 2nd ed. Am J

Psychiatry 161(Suppl 2):1�/114.

Andreasen NC, Olsen S. 1982. Negative vs positive schizophre-

nia. Definition and validation. Arch Gen Psychiatry 39(7):

789�/794.

Angermeyer MC. 2000. Schizophrenia and violence. Acta Psy-

chiatr Scand 102(Suppl 407):63�/67.

Anil Yagcioglu AE, Kivircik Akdede BB, Turgut TI, Tumuklu M,

Yazici MK, Alptekin K, Ertugrul A, Jayathilake K, Gogus A,

Tunca Z, Meltzer HY. 2005. A double-blind controlled study of

adjunctive treatment with risperidone in schizophrenic patients

partially responsive to clozapine: Efficacy and safety. J Clin

Psychiatry 66(1):63�/72.

Arango C, Kirkpatrick B, Buchanan RW. 2000. Fluoxetine as an

adjunct to conventional antipsychotic treatment of schizophre-

nia patients with residual symptoms. J Nerv Ment Dis 188:50�/

53.

Arango C, Bobes J. 2004. Managing acute exacerbations of

schizophrenia: Focus on quetiapine. Curr Med Res Opin

20(5):619�/626.

Arato M, O’Connor R, Meltzer HY. 2002. A 1-year, double-blind,

placebo-controlled trial of ziprasidone 40, 80 and 160 mg/day

in chronic schizophrenia: The Ziprasidone Extended Use in

Schizophrenia (ZEUS) study. Int Clin Psychopharmacol

17:207�/215.

Arndt S, Tyrrell G, Flaum M, Andreasen NC. 1992. Comorbidity

of substance abuse in schizophrenia: the role of pre-morbid

adjustment. Psychol Med 22:279�/388.

Arvanitis LA, Miller BG (The Seroquel Trial 13 Study Group).

1997. Multiple fixed doses of ‘Seroquel’ (quetiapine) in

patients with acute exacerbation of schizophrenia: A compar-

ison with haloperidol and placebo. Biol Psychiatry 42:233�/

246.

Azorin JM, Spiegel R, Remington G, Vanelle JM, Pere JJ, Giguere

M, Bourdeix I. 2001. A double-blind comparative study of

clozapine and risperidone in the management of severe chronic

schizophrenia. Am J Psychiatry 158:1305�/1313.

Baker AA, Bird G, Lavin NI, Thorpe JG. 1960. ECT in

schizophrenia. J Ment Sci 106:1506�/1511.

176 P. Falkai et al.

Baker A, Lewin T, Reichler H, Clancy R, Carr V, Garrett R, Sly

K, Devir H, Terry M. 2002. Motivational interviewing among

psychiatric in-patients with substance use disorders. Acta

Psychiatr Scand 106:233�/240.

Baldwin D, Birtwistle J. 1997. Schizophrenia, antipsychotic drugs

and sexual function. Prim Care Psychiatry 3:117�/122.

Barak Y, Mirecki I, Knobler HY, Natan Z, Aizenberg D. 2004.

Suicidality and second generation antipsychotics in schizophre-

nia patients: A case-controlled retrospective study during a 5-

year period. Psychopharmacology (Berlin) 175(2):215�/219.

Bark N, Revheim N, Huq F, Khalderov V, Ganz ZW, Medalia A.

2003. The impact of cognitive remediation on psychiatric

symptoms of schizophrenia. Schizophr Res 63:229�/235.

Barnas C, Stuppack CH, Miller C, Haring C, Sperner-Unterwe-

ger B, Fleischhacker WW. 1992. Zotepine in the treatment of

schizophrenic patients with prevailingly negative symptoms. A

double-blind trial vs. haloperidol. Int Clin Psychopharmacol

7:23�/27.

Barrowclough C, Haddock G, Tarrier N, Lewis SW, Moring J,

O’Brien R, Schofield N, McGovern J. 2001. Randomized

controlled trial of motivational interviewing, cognitive behavior

therapy, and family intervention for patients with comorbid

schizophrenia and substance use disorders. Am J Psychiatry

158:1706�/1713.

Battaglia J, Moss S, Rush J, Kang J, Mendoza R, Leedom L,

Dubin W, McGlynn C, Goodman L. 1997. Haloperidol,

lorazepam, or both for psychotic agitation? A multicenter,

prospective, double-blind, emergency department study. Am J

Emerg Med 15(4):335�/340.

Beasley CM Jr, Tollefson G, Tran P, Satterlee W, Sanger T,

Hamilton S. 1996a. Olanzapine versus placebo and haloper-

idol: Acute phase results of the North American double-blind

olanzapine trial. Neuropsychopharmacology 14:111�/123.

Beasley CM Jr, Sanger T, Satterlee W, Tollefson G, Tran P,

Hamilton S. 1996b. Olanzapine versus placebo: results of a

double-blind, fixed-dose olanzapine trial. Psychopharmacology

(Berlin) 124:159�/167.

Beasley CM Jr, Hamilton SH, Crawford AM, Dellva MA,

Tollefson GD, Tran PV, Blin O, Beuzen JN. 1997. Olanzapine

versus haloperidol: Acute phase results of the international

double-blind olanzapine trial. Eur Neuropsychopharmacol

7:125�/137.

Beautrais A. 2001. Suicides and serious suicide attempts: Two

populations or one? Psychol Med 31:837�/845.

Beiser M, Erickson D, Fleming JA, Iacono WG. 1993. Establish-

ing the onset of psychotic illness. Am J Psychiatry 150:1349�/

1354.

Berk M, Ichim C, Brook S. 2001. Efficacy of mirtazapine add on

therapy to haloperidol in the treatment of the negative

symptoms of schizophrenia: A double-blind randomized pla-

cebo-controlled study. Int Clin Psychopharmacol 16(2):87�/92.

Berman I, Sapers BL, Chang HH, Losonczy MF, Schmildler J,

Green AI. 1995. Treatment of obsessive-compulsive symptoms

in schizophrenic patients with clomipramine. J Clin Psycho-

pharmacol 15:206�/210.

Bersani G, Orlandi V, Kotzalidis GD, Pancheri P. 2002. Cannabis

and schizophrenia: Impact on onset, course, psychopathology

and outcomes. Eur Arch Psychiatry Clin Neurosci 252:86�/92.

Bilder RM, Goldman RS, Volavka J, Czobor P, Hoptman M,

Sheitman B, Lindenmayer JP, Citrome L, McEvoy J, Kunz M,

Chakos M, Cooper TB, Horowitz TL, Lieberman JA. 2002.

Neurocognitive effects of clozapine, olanzapine, risperidone,

and haloperidol in patients with chronic schizophrenia or

schizoaffective disorder. Am J Psychiatry 159:1018�/1028.

Bieniek SA, Ownby RL, Penalver A, Dominguez RA. 1998. A

double-blind study of lorazepam versus the combination of

haloperidol and lorazepam in managing agitation. Pharma-

cotherapy 18(1):57�/62.

Birchwood M, Smith J, McMillan JF. 1989. Predicting relapse in

schizophrenia: The development and implementation of an

early signs monitoring system using patients and families as

observers, a preliminary investigation. Psychol Med 19:649�/

656.

Blin O, Azorin JM, Bouhours P. 1996. Antipsychotic and

anxiolytic properties of risperidone, haloperidol, and metho-

trimeprazine in schizophrenic patients. J Clin Psychopharmacol

16:38�/44.

Blumer D. 1997. Catatonia and the neuroleptics: Psychobiologic

significance of remote and recent findings. Comp Psychiatry

38(4):193�/201.

Bobes J, Gibert J, Ciudad A, Alvarez E, Canas F, Carrasco JL,

Gascon J, Gomez JC, Gutierrez M. 2003. Safety and effective-

ness of olanzapine versus conventional antipsychotics in the

acute treatment of first-episode schizophrenic inpatients. Prog

Neuropsychopharmacol Biol Psychiatry 27(3):473�/481.

Bollini P, Pampallona S, Orza MJ, Adams ME, Chalmers TC.

1994. Antipsychotic drugs: is more worse? A meta�/analysis of

the published randomized control trials. Psychol Med 24:307�/

316.

Bondolfi G, Dufour H, Patris M, May JP, Billeter U, Eap CB,

Baumann P. 1998. Risperidone vs. clozapine in treatment-

resistant chronic schizophrenia: a randomized double-blind

study. The Risperidone Study Group. Am J Psychiatry

155:499�/504.

Bonta J, Law M, Hanson K. 1998. The prediction of criminal and

violent recidivism among mentally disordered offenders: A

meta-analysis. Psychol Bull 123:123�/142.

Borison RL, Arvanitis LA, Miller BG (US Seroquel Study

Group). 1996. ICI 204,636, an atypical antipsychotic: efficacy

and safety in a multicenter, placebo-controlled trial in patients

with schizophrenia. J Clin Psychopharmacol 16:158�/169.

Bottlender R, Strauß A, Moller HJ. 2000. Impact of duration of

symptoms prior to first hospitalisation on acute outcome in 998

schizophrenic patients. Schizophr Res 44:145�/150.

Bottlender R, Sato T, Jager M, Groll C, Strauss A, Moller HJ.

2002. The impact of duration of untreated psychosis and

premorbid functioning on outcome of first inpatient treatment

in schizophrenic and schizoaffective patients. Eur Arch Psy-

chiatry Clin Neurosci 252:226�/231.

Bottlender R, Sato T, Jager M, Wegener U, Wittmann J, Strauss

A, Moller HJ. 2003. The impact of the duration of untreated

psychosis prior to first psychiatric admission on the 15-year

outcome in schizophrenia. Schizophr Res 62:37�/44.

Boyer P, Lecrubier Y, Puech AJ, Dewailly, Aubin F. 1995.

Treatment of negative symptoms in schizophrenia with ami-

sulpride. Br J Psychiatry 166:68�/72.

Bradford DW, Perkins DO, Lieberman JA. 2003. Pharmacological

management of first-episode schizophrenia and related non-

affective psychoses. Drugs 63(21):2265�/2283.

Braude WM, Barnes TRE, Gore SM. 1983. Clinical character-

istics of akathisia: A systematic investigation of acute psychia-

tric inpatient admissions. Br J Psychiatry 143:139�/150.

Breier A, Buchanan RW, Kirkpatrick B, Davis OR, Irish D,

Summerfelt A, Carpenter WT. 1994. Effects of clozapine on

positive and negative symptoms in outpatients with schizo-

phrenia. Am J Psychiatry 151:20�/26.

Breier A, Hamilton SH. 1999. Comparative efficacy of olanzapine

and haloperidol for patients with treatment-resistant schizo-

phrenia. Biol Psychiatry 45:403�/411.

Breier AF, Malhotra AK, Su TP, Pinals DA, Elman I, Adler CM,

Lafargue RT, Clifton A, Pickar D. 1999. Clozapine and

risperidone in chronic schizophrenia: effects on symptoms,

WFSBP Guidelines for Biological Treatment of Schizophrenia 177

parkinsonian side effects, and neuroendocrine response. Am J

Psychiatry 156:294�/298.

Brenner HD, Dencker SJ, Goldstein MJ, Hubbard JW, Keegan

DL, Kruger G, Kulhanek F, Liberman RP, Malm U, Midha

KK. 1990. Defining treatment refractoriness in schizophrenia.

Schizophr Bull 16:551�/561.

Brown S, Inskip H, Barraclough B. 2000. Causes of the excess

mortality of schizophrenia. Br J Psychiatry 177:212�/217.

Brown ES, Nejtek VA, Perantie DC. 2002. Neuroleptics and

quetiapine in psychiatric illnesses with comorbid stimulant

abuse. Abstract/Poster, CINP; June 23�/27 2002, Montreal,

Canada.

Buchanan RW, Holstein C, Breier A. 1994. The comparative

efficacy and long-term effect of clozapine treatment on

neuropsychological test performance. Biol Psychiatry 36(11):

717�/725.

Buchanan RW. 1995. Clozapine: efficacy and safety. Schizophr

Bull 21(4):579�/591.

Buchanan RW, Kirkpatrick B, Bryant N, Ball P, Breier A. 1996.

Fluoxetine augmentation of clozapine treatment in patients

with schizophrenia. Am J Psychiatry 153:1625�/1627.

Buchanan RW, Breier A, Kirkpatrick B, Ball P, Carpenter WT Jr.

1998. Positive and negative symptom response to clozapine in

schizophrenic patients with and without the deficit syndrome.

Am J Psychiatry 155:751�/760.

Buchanan RW, Summerfelt A, Tek C, Gold J. 2003. An open-

labeled trial of adjunctive donepezil for cognitive impairments

in patients with schizophrenia. Schizophr Res 59:29�/33.

Buckley P, Thompson PA, Way L, Meltzer HY. 1994a. Substance

use among patients with treatment-resistant schizophrenia:

Characteristics and implication for clozapine therapy. Am J

Psychiatry 151:385�/389.

Buckley P, Thompson PA, Way L, Meltzer HY. 1994b. Substance

use and clozapine treatment. J Clin Psychiatry 55(Suppl

B):114�/116.

Buckley PF, Noffsinger SG, Smith DA, Hrouda DR, Knoll JL.

2003. Treatment of the psychotic patient who is violent.

Psychiatr Clin North Am 26:231�/272.

Burns MJ. 2001. The pharmacology and toxicology of atypical

antipsychotic agents. Clin Toxicol 39(1):1�/14.

Burt T, Lisanby SH, Sackeim HA. 2002. Neuropsychiatric

applications of transcranial magnetic stimulation: A meta

analysis. Int J Neuropsychopharmacol 5:73�/103.

Bush G, Fink M, Petrides G, Dowling F, Francis A. 1996.

Catatonia. II: Treatment with lorazepam and electroconvulsive

therapy. Acta Psychiatr Scand 93:137�/143.

Bushe C, Holt R. 2004. Prevalence of diabetes and impaired

glucose tolerance in patients with schizophrenia. Br J Psychia-

try 184(Suppl 47):s67�/71.

Bushe C, Leonard B. 2004. Association between atypical anti-

psychotic agents and type 2 diabetes: Review of prospective

clinical data. Br J Psychiatry 184(Suppl 47):s87�/93.

Caroff SN, Mann SC, Campbell EC. 2000. Atypical antipsycho-

tics and neuroleptic malignant syndrome. Psychiatr Ann

30:314�/321.

Carpenter WT Jr, Heinrichs DW, Alphs LD. 1985. Treatment of

negative symptoms. Schizophr Bull 11:440�/452.

Carriere P, Bonhomme D, Lemperiere T. 2000. Amisulpride has a

superior benefit/risk profile to haloperidol in schizophrenia:

Results of a multicentre double-blind study (the Amisulpride

Study Group). Eur Psychiatry 15:321�/329.

Casey DE, Daniel DG, Wassef AA, Tracy LA, Wozniak P,

Sommerville KW. 2003a. Reply effect of divalproex combined

with olanzapine or risperidone in patients with an acute

exacerbation of schizophrenia. Neuropsychopharmacology

28:2052�/2053.

Casey DE, Carson WH, Saha AR, Liebeskind A, Ali MW, Jody D,

Ingenito GG; Aripiprazole Study Group. 2003b. Switching

patients to aripiprazole from other antipsychotic agents: a

multicenter randomized study. Psychopharmacology (Berlin)

166(4):391�/399.

Caspari D, Wobrock T. 2004. Cannabis psychosis �/ from a

distinct category to comorbidity. Sucht 50(5):320�/326.

Caspi N, Modai I, Barak P, Waisbourd A, Zbarsky H, Hirsch-

mann S, Ritsner M. 2001. Pindolol augmentation in aggressive

schizophrenic patients: A double-blind crossover randomized

study. Int Clin Psychopharmacol 16:111�/115.

Cassens G, Inglis AK, Appelbaum PS, Gutheil TG. 1990.

Neuroleptic effects on neuropsychological function in chronic

schizophrenic patients. Schizophr Bull 16:477�/499.

Castle DJ, Wessely S, Murray RM. 1993. Sex and schizophrenia:

Effects of diagnostic stringency, and associations with and

premorbid variables. Br J Psychiatry 162:658�/664.

Castle DJ, Abel K, Takei N, Murray RM. 1995. Gender

differences in schizophrenia: Hormonal effect or subtypes?

Schizophr Bull 21(1):1�/12.

Cavallaro C, Cordoba C, Smeraldi E. 1995. A pilot, open study

on the treatment of refractory schizophrenia with risperidone

and clozapine. Hum Psychopharmacol 10:231�/234.

Centre for Outcomes Research and Effectiveness of the British

Psychological Society. Systematic Review Fluphenazine (Phar-

macology Group of the National Institute for Clinical Excel-

lence). In: National Institute for Clinical Excellence.

Schizophrenia: core interventions in the treatment and man-

agement of schizophrenia in primary and secondary care. 2002.

Centre for Outcomes Research and Effectiveness of the British

Psychological Society. Systematic Review Flupenthixol (Phar-

macology Group of the National Institute for Clinical Excel-

lence). In: National Institute for Clinical Excellence.

Schizophrenia: core interventions in the treatment and man-

agement of schizophrenia in primary and secondary care. 2002.

Centre for Outcomes Research and Effectiveness of the British

Psychological Society. Systematic Review Perphenazine (Phar-

macology Group of the National Institute for Clinical Excel-

lence). In: National Institute for Clinical Excellence.

Schizophrenia: core interventions in the treatment and man-

agement of schizophrenia in primary and secondary care. 2002.

Ceskova E, Svestka J. 1993. Double-blind comparison of risper-

idone and haloperidol in schizophrenic and schizoaffective

psychoses. Pharmacopsychiatry 26:121�/124.

Chakos M, Lieberman J, Hoffman E, Bradford D, Sheitman B.

2001. Effectiveness of second-generation antipsychotics in

patients with treatment-resistant schizophrenia: a review and

meta-analysis of randomized trials. Am J Psychiatry 158:518�/

526.

Chanpattana W, Chakrabhand ML, Sackeim HA, Kitaroonchai

W, Kongsakon R, Techakasem P, Buppanharun W, Tuntirung-

see Y, Kirdcharoen N. 1999. Continuation ECT in treatment-

resistant schizophrenia: A controlled study. J ECT 15:178�/

192.

Chanpattana W, Chakrabhand ML. 2001. Combined ECT and

neuroleptic therapy in treatment-refractory schizophrenia:

Prediction of outcome. Psychiatr Res 105:107�/115.

Cheer SM, Wagstaff AJ. 2004. Quetiapine. A review of its use in

the management of schizophrenia. CNS Drugs 18(3):173�/

199.

Cheine M, Ahonen J, Wahlbeck K. 2004. Beta-blocker supple-

mentation of standard drug treatment for schizophrenia

(Cochrane Review). In: The Cochrane Library, Issue 2.

Chichester, UK: John Wiley & Sons Ltd.

Chengappa KN, Vasile J, Levine J, Ulrich R, Baker R, Gopalani A,

Schooler N. 2002. Clozapine: Its impact on aggressive behavior

178 P. Falkai et al.

among patients in a state psychiatric hospital. Schizophr Res

53:1�/6.

Chengappa KN, Levine J, Ulrich R, Parepally H, Brar JS, Atzert

R, Brienzo R, Gopalani A. 2000. Impact of risperidone on

seclusion and restraint at a state psychiatric hospital. Can J

Psychiatry 45(9):827�/832.

Chiu E, Burrows G, Stevenson J. 1976. Double-blind comparison

of clozapine with chlorpromazine in acute schizophrenic illness.

Aust NZ J Psychiatry 10:343�/347.

Chiu NY, Yang YK, Chen PS, Chang CC, Lee IH, Lee JR. 2003.

Olanzapine in Chinese treatment-resistant patients with schizo-

phrenia: An open-label, prospective trial. Psychiatry Clin

Neurosci 57:478�/484.

Chouinard G, Jones B, Remington G, Bloom D, Addington D,

MacEwan GW, Labelle A, Beauclair L, Arnott W. 1993. A

Canadian multicenter placebo-controlled study of fixed doses

of risperidone and haloperidol in the treatment of chronic

schizophrenic patients. J Clin Psychopharmacol 13:25�/40.

Citrome L, Volavka J. 2002. Optimal dosing of atypical anti-

psychotics in adults: A review of the current evidence. Harv

Rev Psychiatr 10(5):280�/291.

Citrome L, Casey DE, Daniel DG, Wozniak P, Kochan LD, Tracy

KA. 2004. Adjunctive divalproex and hostility among patients

with schizophrenia receiving olanzapine or risperidone. Psy-

chiatr Serv 55(3):290�/294.

Clark RE. 2001. Family support and substance use outcome for

persons with mental illness and substance use disorders.

Schizophr Bull 27:93�/101.

Claus A, Bollen J, De Cuyper H, Eneman M, Malfroid M,

Peuskens J, Heylen S. 1992. Risperidone versus haloperidol in

the treatment of chronic schizophrenic inpatients: A multi-

centre double-blind comparative study. Acta Psychiatr Scand

85:295�/305.

Cohen LS, Rosenbaum JF. 1998. Psychotropic drug use during

pregnancy: Weighing the risks. J Clin Psychiatry 59(Suppl

2):18�/28.

Colonna L, Saleem P, Dondey-Nouvel L, Rein W, Amisulpride

Study Group. 2000. Long-term safety and efficacy of amisul-

pride in sub-chronic or chronic schizophrenia. Int J Clin

Psychopharmacol 15(1):13�/22.

Conley RR, Buchanan RW. 1997. Evaluation of treatment-

resistant schizophrenia. Schizophr Bull 23:663�/674.

Conley RR, Carpenter WT, Tamminga CA. 1997. Time to

clozapine response in a standardized trial. Am J Psychiatry

154:1243�/1247.

Conley RR, Kelly DL, Gale EA. 1998. Olanzapine response in

treatment-refractory schizophrenic patients with a history of

substance abuse. Schizophr Res 33:95�/101.

Conley RR, Mahmoud R. 2001. A randomized double-blind

study of risperidone and olanzapine in the treatment of

schizophrenia or schizoaffective disorder. Am J Psychiatry

158:765�/774.

Cooper SJ, Tweed J, Raniwalla J, Butler A, Welch C. 2000a. A

placebo-controlled comparison of zotepine versus chlorproma-

zine in patients with acute exacerbation of schizophrenia. Acta

Psychiatr Scand 101:218�/225.

Cooper SJ, Butler A, Tweed J, Welch C, Raniwalla J. 2000b.

Zotepine in the prevention of recurrence: a randomised,

double-blind, placebo-controlled study for chronic schizophre-

nia. Psychopharmacology (Berlin) 150:237�/243.

Copolov DL, Link CG, Kowalcyk B. 2000. A multicentre,

double-blind, randomized comparison of quetiapine (ICI

204,636, ‘Seroquel’) and haloperidol in schizophrenia. Psychol

Med 30:95�/105.

Cormac I, Jones C, Campbell C. 2004. Cognitive behaviour

therapy for schizophrenia. (Cochrane Review). In: The Co-

chrane Library, Issue 2. Chichester, UK: John Wiley & Sons

Ltd.

Correll CU, Leucht S, Kane JM. 2004. Lower risk for tardive

dyskinesia associated with second-generation antipsychotics: A

systematic review of 1-year studies. Am J Psychiatry 161:414�/

425.

Coryell W, Miller DD, Perry PJ. 1998. Haloperidol plasma levels

and dose optimization. Am J Psychiatry 155:48�/53.

Creese I, Burt DR, Snyder SH. 1976. Dopamine receptor binding

predicts clinical and pharmacological potencies of antischizo-

phrenic drugs. Science 192:481�/483.

Crow TJ. 1985. The two syndrome concept: Origins and current

status. Schizophr Bull 11:471�/486.

Csernansky JG, Riney SJ, Lombarozo L, Overall JE, Hollister LE.

1988. Double-blind comparison of alprazolam, diazepam and

placebo for the treatment of negative schizophrenic symptoms.

Arch Gen Psychiatry 45:655�/659.

Csernansky JG, Mahmoud R, Brenner R. 2002. A comparison of

risperidone and haloperidol for the prevention of relapse in

patients with schizophrenia. New Engl J Med 346:16�/22.

Currier GW, Simpson GM. 2001. Risperidone liquid concentrate

and oral lorazepam versus intramuscular haloperidol and

intramuscular lorazepam for treatment of psychotic agitation.

J Clin Psychiatry 62(3):153�/157.

Cutler AJ. 2003. Sexual dysfunction and antipsychotic treatment.

Psychoneuroendocrinology 28:69�/82.

Dahl SG. 1990. Pharmacokinetics of antipsychotic drugs in man.

Acta Psychiatr Scand (Suppl) 358:37�/40.

Daniel DG, Goldberg TE, Weinberger DR, Kleinman JE, Pickar

D, Lubick LJ, Williams TS. 1996. Different side effect profiles

of risperidone and clozapine in 20 outpatients with schizo-

phrenia or schizoaffective disorder: a pilot study. Am J

Psychiatry 153:417�/419.

Daniel DG, Zimbroff DL, Potkin SG, Reeves KR, Harrigan EP,

Lakshminarayanan M (Ziprasidone Study Group). 1999.

Ziprasidone 80 mg/day and 160 mg/day in the acute exacerba-

tion of schizophrenia and schizoaffective disorder: A 6-week

placebo-controlled trial. Neuropsychopharmacology 20:491�/

505.

Daniel DG, Potkin SG, Reeves KR, Swift RH, Harrigan EP. 2001.

Intramuscular (IM) ziprasidone 20 mg is effective in reducing

acute agitation associated with psychosis: A double-blind,

randomized trial. Psychopharmacology (Berlin) 155:128�/341.

Danion JM, Rein W, Fleurot O, Amisulpride Study Group. 1999.

Improvement of schizophrenic patients with primary negative

symptoms treated with amisulpride. Am J Psychiatry 156:610�/

616.

Datto CJ. 2000. Side effects of electroconvulsive therapy. Depress

Anxiety 12:130�/134.

Davidson L, McGlashan TH. 1997. The varied outcomes of

schizophrenia. Can J Psychiatry 42:34�/43.

Davis JM. 1975. Overview: Maintenance therapy in psychiatry: I.

Schizophrenia. Am J Psychiatry 132:1237�/1245.

Davis JM, Barter JT, Kane JM. 1989. Antipsychotic drugs. In:

Kaplan HI, Sadock BJ, editors. Comprehensive textbook of

psychiatry. 5th ed. Baltimore, MD: Williams & Wilkins. pp

1591�/1626.

Davis JM, Chen N, Glick ID. 2003. A meta-analysis of the efficacy

of second-generation antipsychotics. Arch Gen Psychiatry 60:

553�/564.

Davis JM, Chen N. 2004. Dose response and dose equivalence of

antipsychotics. J Clin Psychopharmacol 24(2):192�/208.

De Nayer A, Windhager E, Irmansyah X, et al. 2003. Efficacy and

tolerability of quetiapine in patients with schizophrenia

switched from other antipsychotics. Int J Psychiatry Clin Pract

7:59�/70.

WFSBP Guidelines for Biological Treatment of Schizophrenia 179

Delcker A, Schoon ML, Oczkowski B, Gaertner HJ. 1990.

Amisulpride versus haloperidol in treatment of schizophrenic

patients*/Results of a double-blind study. Pharmacopsychiatry

23:125�/130.

Devanand DP, Dwork AJ, Hutchinson ER, Bolwig TG, Sackheim

HA. 1994. Does ECT alter brain structure? Am J Psychiatry

151:957�/970.

Devinsky O, Honigfeld G, Patin J. 1991. Clozapine-related

seizures. Neurology 41:369�/371.

Dickson RA, Glazer WM. 1999. Neuroleptic-induced hyperpro-

lactinemia. Schizophr Res 35(Suppl):S75�/86.

Dieterle DM, Muller Spahn F, Ackenheil M. 1991. [Effectiveness

and tolerance of zotepine in a double-blind comparison with

perazine in schizophrenic patients]. Fortschr Neurol Psychiat

59(Suppl 1):18�/22.

Dinan TG. 2004. Stress and the genesis of diabetes mellitus in

schizophrenia. Br J Psychiatry 184(Suppl 47):s72�/75.

Dixon L, Haas G, Weiden PJ, Sweeney J, Frances AJ. 1991. Drug

abuse in schizophrenic patients: Clinical correlates and reasons

for use. Am J Psychiatry 14:224�/230.

Dixon LB, Lehman AF, Levine J. 1995. Conventional antipsy-

chotic medications for schizophrenia. Schizophr Bull 21(4):

567�/577.

Dodwell D, Goldberg D. 1989. A study of factors associated with

response to electroconvulsive therapy in patients with schizo-

phrenic symptoms. Br J Psychiatry 154:635�/639.

Donahue AB. 2000. Electroconvulsive therapy and memory loss:

a personal journey. J ECT 16:133�/143.

Dorevitch A, Katz N, Zemishlany Z, Aizenberg D, Weizman A.

1999. Intramuscular flunitrazepam versus intramuscular halo-

peridol in the emergency treatment of aggressive psychotic

behavior. Am J Psychiatry 156(1):142�/144.

Dose M, Hellweg R, Yassouridis A, Theison M, Emrich HM.

1998. Combined treatment of schizophrenic psychoses with

haloperidol and valproate. Pharmacopsychiatry 31:122�/125.

Dossenbach MRK, Beuzen JN, Avnon M, Belmaker RH, Elizur

A, Merk M, Munitz H, Schneidman M, Shoshani D, Kratky P,

Grundy SL, Tollefson GD. 2000. The effectiveness of olanza-

pine in treatment-refractory schizophrenia when patients are

nonresponsive to or unable to tolerate clozapine. Clin Ther

22:1021�/1034.

Drake RE, Mueser KT. 2000. Psychosocial approaches to dual

diagnosis. Schizophr Bull 26:105�/118.

Drury V, Birchwood M, Cochrane R. 2000. Cognitive therapy

and recovery from acute psychosis: a controlled trial. Br J

Psychiatry 177:8�/14.

Dufresne RL, Valentino D, Kass DJ. 1993. Thioridazine improves

affective symptoms in schizophrenic patients. Psychopharma-

col Bull 29:249�/255.

Duke PJ, Pantelis C, McPhilipps MA, Barnes TRE. 2001.

Comorbid non-alcohol substance misuse among people with

schizophrenia. Br J Psychiatry 179:509�/513.

Dursun SM, McIntosh D, Milliken H. 1999. Clozapine plus

lamotrigine in treatment-resistant schizophrenia. Arch Gen

Psychiatry 56:950.

Dursun SM, Deakin JFW. 2001. Augmenting antipsychotic

treatment with lamotrigine or topiramate in patients with

treatment-resistant schizophrenia: A naturalistic case-series

outcome study. J Psychopharmacol 15(4):297�/301.

Ebrecht M, Jody D, Kane J, Carson W, Kujawa M, Stringfellow J,

Marcus R, Sanchez R, Meltzer H. 2004. Aripiprazole versus

perphenazine in treatment-resistant schizophrenia. Poster pre-

sented at 12th AEP Congress, Geneva, Switzerland, 14�/18

April.

Emsley RA. 1999. Risperidone in the treatment of first-episode

psychotic patients: A double-blind multicenter study. Risper-

idone Working Group. Schizophr Bull 25:721�/729.

Emsley RA, Raniwalla J, Bailey PJ, Jones AM. 2000. A compar-

ison of the effects of quetiapine (‘seroquel’) and haloperidol in

schizophrenic patients with a history of and a demonstrated,

partial response to conventional antipsychotic treatment.

PRIZE Study Group. Int Clin Psychopharmacol 15:121�/131.

Emsley R, Oosthuizen P, van Rensburg SJ. 2003. Clinical

potential of omega-3 fatty acids in the treatment of schizo-

phrenia. CNS Drugs 17(15):1081�/1091.

Ernst LC, Goldberg JF. 2002. The reproductive safety profile of

mood stabilizers, atypical antipsychotics, and broad-spectrum

psychotropics. J Clin Psychiatry 63(Suppl 4):42�/55.

Essock SM, Hargreaves WA, Dohm FA, Goethe J, Carver L,

Hipshman L. 1996a. Clozapine eligibility among state hospital

patients. Schizophr Bull 22:15�/25.

Essock SM, Hargreaves WA, Covell NH, Goethe J. 1996b.

Clozapine’s effectiveness for patients in state hospitals: results

from a randomized trial. Psychopharmacol Bull 32:683�/697.

Evins AE, Fitzgerald SM, Wine L, Rosselli R, Goff DC. 2000.

Placebo-controlled trial of glycine added to clozapine in

schizophrenia. Am J Psychiatry 157:826�/828.

Ewing JA, Meuller RA, Rouse BA, et al. 1977. Low levels of

dopamine-beta-hydroxylase and psychosis. Am J Psychiatry

134:927�/928.

Fabre LF Jr, Arvanitis L, Pultz J, Jones VM, Malick JB, Slotnick

VB. 1995. ICI 204,636, a novel, atypical antipsychotic: early

indication of safety and efficacy in patients with chronic and

subchronic schizophrenia. Clin Ther 17:366�/378.

Fagerlund B, Mackeprang T, Gade A, Glenthoj BY. 2004. Effects

of low-dose risperidone and low-dose zuclopenthixol on

cognitive functions in first-episode drug-naive schizophrenic

patients. CNS Spectr 9(5):364�/374.

Falloon IRH. 2003. Family interventions for mental disorders:

Efficacy and effectiveness. World J Psychiatry 2:20�/28.

Farde L, Nordstrom AL, Wiesel FA, Pauli S, Halldin C, Sedvall

G. 1992. Positron emission tomographic analysis of central D1

and D2 dopamine receptor occupancy in patients treated with

classical neuroleptics and clozapine. Relation to extrapyramidal

side effects. Arch Gen Psychiatry 49:538�/544.

Farren CK, Hameedi FA, Rosen MA, Woods S, Jatlow P, Kosten

TR. 2000. Significant interaction between clozapine and

cocaine in cocaine addicts. Drug Alcohol Depend 59:153�/163.

Fenton WS, McGlashan TH. 1987. Prognostic scale for chronic

schizophrenia. Schizophr Bull 13:277�/286.

Fenton WS. 2000. Prevalence of spontaneous dyskinesia in

schizophrenia. J Clin Psychiatry 61(Suppl 4):10�/14.

Fenton M, Coutinho ESF, Campbell C. 2004. Zuclopenthixol

acetate in the treatment of acute schizophrenia and similar

serious mental illnesses (Cochrane Review). In: The Cochrane

Library, Issue 2. Chichester, UK: John Wiley & Sons Ltd.

Fink M, Sackeim HA. 1996. Convulsive therapy in schizophrenia?

Schizophr Bull 22:27�/39.

Fischer-Cornelssen KA, Ferner UJ. 1976. An example of Eur-

opean multicenter trials: Multispectral analysis of clozapine.

Psychopharmacol Bull 12:34�/39.

Fleischhacker WW, Barnas C, Stuppack CH, Unterweger B,

Miller C, Hinterhuber H. 1989. Zotepine vs. haloperidol in

paranoid schizophrenia: A double-blind trial. Psychopharmacol

Bull 25:97�/100.

Fleischhacker WW. 2002. Second generation antipsychotics.

Psychopharmacology 162:90�/91.

Flynn SW, MacEwan GW, Altman S, Kopala LC, Fredrikson

DH, Smith GN, Honer WG. 1998. An open comparison of

clozapine and risperidone in treatment-resistant schizophrenia.

Pharmacopsychiatry 31:25�/29.

Folstein M, Folstein S, McHugh PR. 1973. Clinical predictors of

im-provement after electroconvulsive therapy of patients with

180 P. Falkai et al.

schizophrenia, neurotic reactions, and affective disorders. Biol

Psychiatry 7:147�/152.

Fortier P, Trudel G, Mottard JP, Piche L. 2000. The influence of

schizophrenia and standard or atypical neuroleptics on sexual

and sociosexual functioning: a review. Sexuality and Disability

18(2):85�/104.

Foster S, Kessel J, Berman ME, Simpson GM. 1997. Efficacy of

lorazepam and haloperidol for rapid tranquilization in a

psychiatric emergency room setting. Int Clin Psychopharmacol

12(3):175�/179.

Fowler IL, Carr VJ, Carter NT, Lewin TJ. 1998. Patterns of

current and lifetime substance use in schizophrenia. Schizophr

Bull 24:443�/455.

Freeman HL. 1997. Amisulpride compared with standard neu-

roleptics in acute exacerbations of schizophrenia: Three effi-

cacy studies. Int Clin Psychopharmacol 12(Suppl 2):S11�/17.

French K, Eastwood D. 2003. Response of catatonic schizophre-

nia to amisulpride: A case report. Can J Psychiatry 48(8):570.

Freudenreich O, Goff DC. 2002. Antipsychotic combination

therapy in schizophrenia. A review of efficacy and risks of

current combinations. Acta Psychiatr Scand 106:323�/330.

Friedman JI, Adler DN, Howanitz E, Harvey PD, Brenner G,

Temporini H, White L, Parrella M, Davis KL. 2002. A double

blind placebo controlled trial of donepezil adjunctive treatment

to risperidone for the cognitive impairment of schizophrenia.

Biol Psychiatry 51:349�/357.

Gaebel W, Frick U, Kopcke W, Linden M, Muller P, Muller-

Spahn F, Pietzcker A, Tegeler J. 1993. Early neuroleptic

intervention in schizophrenia: Are prodromal symptoms valid

predictors of relapse? Br J Psychiatry 163(Suppl 21):8�/12.

Gaebel W, Baumann AE. 2003. Interventions to reduce the stigma

associated with severe mental illness: Experiences from the

Open the doors Program in Germany. Can J Psychiatry 48:

657�/662.

Gaszner P, Makkos Z. 2004. Clozapine maintenance therapy in

schizophrenia. Prog Neuropsychopharmacol Biol Psychiatry

28(3):465�/469.

Gawin FH. 1986. Neuroleptic reduction of cocaine-induced

paranoia but not euphoria? Psychopharmacology 90:142�/143.

Geddes J, Freemantle N, Harrison P, Bebbington P. 2000.

Atypical antipsychotics in the treatment of schizophrenia:

Systematic overview and meta-regression analysis. Br Med J

321(7273):1371�/1376.

Gelenberg AJ, Doller JC. 1979. Clozapine versus chlorpromazine

for the treatment of schizophrenia: Preliminary results from a

double-blind study. J Clin Psychiatry 40:238�/240.

Glassman AH, Bigger JT Jr. 2001. Antipsychotic drugs: prolonged

QTc interval, torsade de pointes, and sudden death. Am J

Psychiatry 158:1774�/1782.

Glazer WM. 2000. Review of incidence studies of tardive

dyskinesia associated with typical antipsychotics. J Clin Psy-

chiatry 61(Suppl 4):15�/20.

Goetz CG, Klawans HL. 1981. Drug-induced extrapyramidal

disorders: A neuropsychiatric interface. J Clin Psychopharma-

col 1:297�/303.

Goff DC, Midha KK, Sarid-Segal O, Hubbard JW, Amico E.

1995. A placebo-controlled trial of fluoxetine added to neuro-

leptic in patients with schizophrenia. Psychopharmacology

(Berlin) 117:417�/423.

Goff DC, Posever T, Herz L, Simmons J, Kletti N, Lapierre K,

Wilner KD, Law CG, Ko GN. 1998. An exploratory haloper-

idol-controlled dose-finding study of ziprasidone in hospita-

lized patients with schizophrenia or schizoaffective disorder. J

Clin Psychopharmacol 18:296�/304.

Goff DC, Henderson DC, Evins AE, Amico E. 1999. A placebo-

controlled crossover trial of d-cycloserine added to clozapine in

patients with schizophrenia. Biol Psychiatry 45:512�/514.

Gold LH. 2000. Use of psychotropic medication during preg-

nancy: Risk management guidelines. Psychiatr Ann 30:421�/

432.

Goldstein JM, Tsaung MT. 1990. Gender and schizophrenia: an

introduction and synthesis of findings. Schizophr Bull 16:179�/

183.

Goodman LA, Rosenberg SD, Mueser KT, Drake RE. 1997.

Physical and sexual assault history in women with serious

mental illness: Prevalence, correlates, treatment, and future

research directions. Schizophr Bull 23(4):685�/696.

Grebb JA. 1995. Medication induced movement disorders. In:

Kaplan HI, Sadock BJ, editors. Comprehensive textbook of

psychiatry. New York: Williams & Wilkins.

Green MF. 1996. What are the functional consequences of

neurocognitive deficits in schizophrenia? Am J Psychiatry

153:321�/330.

Green MF, Marshall BD Jr, Wirshing WC, Ames D, Marder SR,

McGurk S, Kern RS, Mintz J. 1997. Does risperidone improve

verbal working memory in treatment-resistant schizophrenia?

Am J Psychiatry 154:799�/804.

Green MF, Marder SR, Glynn SM, McGurk SR, Wirshing WC,

Wirshing DA, Liberman RP, Mintz J. 2002. The neurocogni-

tive effects of low-dose haloperidol: A two-year comparison

with risperidone. Biol Psychiatry 51:972�/978.

Green AI, Burgess ES, Dawson R, Zimmet SV, Strous RD. 2003.

Alcohol and cannabis use in schizophrenia: Effects of clozapine

vs. risperidone. Schizophr Res 60(1):81�/85.

Growe GA, Crayton JW, Klass DB, Evans H, Strizich M. 1979.

Lithium in chronic schizophrenia. Am J Psychiatry 136:454�/

455.

Guest JF, Cookson RF. 1999. Costs of schizophrenia to UK

Society. An incidence-based cost-of-illness model for the first 5

years following diagnosis. Pharmaeconomics 15:597�/610.

Guirguis E, Voineskos G, Gray J, Schlieman E. 1977. Clozapine

(Leponex) versus chlorpromazine (Largactil) in acute schizo-

phrenia (a double-blind controlled study). Curr Ther Res

21:707�/719.

Gupta S, Sonnenberg SJ, Frank B. 1998. Olanzapine augmenta-

tion of clozapine. Ann Clin Psychiatry 10:113�/115.

Gupta N, Basu D. 2001. Does risperidon reduce concomitant

substance abuse in cases of schizophrenia? Can J Psychiatry

46:862�/863.

Haddad PM, Anderson IM. 2002. Antipsychotic-related QTc

prolongation, torsade de pointes and sudden death. Drugs

62:1649�/1671.

Haddad PM. 2004. Antipsychotics and diabetes: Review of non-

prospective data. Br J Psychiatry 184(Suppl 47):80�/86.

Hafner H. 2000. Gender differences in schizophrenia. In: Frank

B, editor. Gender and its effects on psychopathology. Washing-

ton, DC: American Psychiatric Press. pp 187�/228.

Hafner H, an der Heiden W. 2003. Course and outcome of

schizophrenia. In: Hirsch SR, Weinberger DR, editors. Schizo-

phrenia. Massachusetts, Oxford, Victoria, Berlin: Blackwell

Science. pp 101�/141.

Hafner H, Maurer K, Loffler W, Riecher-Rossler A. 1993. The

influence of age and sex on the onset and early course of

schizophrenia. Br J Psychiatry 162:80�/86.

Hamilton SH, Revicki DA, Genduso LA, Beasley CM Jr. 1998.

Olanzapine versus placebo and haloperidol: Quality of life and

efficacy results of the North American double-blind trial.

Neuropsychopharmacology 18:41�/49.

Hanlon TE, Ota KY, Burland AA. 1970. Comparative effects of

fluphenazine, fluphenazine-chlordiazepoxide, and fluophena-

zine-imipramine. Dis Nerv Syst 31:171�/177.

Harada T, Otsuki S, Fujiwara Y. 1991. Effectiveness of zotepine in

therapy-refractory psychoses. An open, multicenter study in

WFSBP Guidelines for Biological Treatment of Schizophrenia 181

eight psychiatric clinics. Fortschr Neurol Psychiatr 59(Suppl

1):41�/44.

Harvey PD, Keefe RS. 2001. Studies of cognitive change in

patients with schizophrenia following novel antipsychotic

treatment. Am J Psychiatry 158(2):176�/184.

Harvey PD, Siu CO, Romano S. 2004. Randomized, controlled,

double-blind, multicenter comparison of the cognitive effects of

ziprasidone versus olanzapine in acutely ill inpatients with

schizophrenia or schizoaffective disorder. Psychopharmacology

(Berlin) 172(3):324�/332.

Hawkins JM, Archer KJ, Strakowski SM, Keck PE. 1995. Somatic

treatment of catatonia. Int J Psychiatry Med 25(4):345�/369.

Hegarty JD, Baldessarini RJ, Tohen M, Waternaux C, Oepen G.

1994. One hundred years of schizophrenia: A meta-analysis of

the outcome literature. Am J Psychiatry 151:1409�/1416.

Heinrich K, Klieser E, Lehmann E, Kinzler E, Hruschka H. 1994.

Risperidone versus clozapine in the treatment of schizophrenic

patients with acute symptoms: A double-blind, randomized

trial. Prog Neuropsychopharmacol Biol Psychiatry 18:129�/

137.

Helgason L. 1990. Twenty years’ follow-up of first psychiatric

presentation for schizophrenia: What could have been pre-

vented? Acta Psychiatr Scand 81:231�/235.

Hellerstein DJ, Rosenthal RN, Miner CR. 1995. A prospective

study of integrated outpatient treatment for substance- abusing

schizophrenic patients. Am J Addict 4:33�/42.

Henderson DC, Goff DC. 1996. Risperidone as an adjunct to

clozapine therapy in chronic schizophrenics. J Clin Psychiatry

57:395�/397.

Henderson DC, Nasrallah RA, Goff DC. 1998. Switching from

clozapine to olanzapine in treatment-refractory schizophrenia:

Safety, clinical efficacy, and predictors of response. J Clin

Psychiatry 59:585�/588.

Hesslinger B, Normann C, Langosch JM, Klose P, Berger M,

Walden J. 1999. Effects of carbamazepine and valproate on

haloperidol plasma levels and on psychopathologic outcome in

schizophrenic patients. J Clin Psychopharmacol 19:310�/315.

Hesslinger B, Walden J, Normann C. 2001. Acute and long-term

treatment of catatonia with risperidone. Pharmacopsychiatry

34(1):25�/26.

Hirose S, Ashby CR, Mills MJ. 2001. Effectiveness of ECT

combined with risperidone against aggression in schizophrenia.

J ECT 17(1):22�/26.

Hirsch SR, Kissling W, Bauml J, Power A, O’Connor R. 2002. A

28-week comparison of ziprasidone and haloperidol in out-

patients with stable schizophrenia. J Clin Psychiatry 63:516�/

523.

Hoff AL, Sakuma M, Wieneke M, Horon R, Kushner M, DeLisi

LE. 1999. Longitudinal neuropsychological follow-up study of

patients with first-episode schizophrenia. Am J Psychiatry

156:1336�/1341.

Hoffman RE, Hawkins KA, Gueorguieva R, Boutros NN, Rachid

F, Carroll K, Krystal JH. 2003. Transcranial magnetic stimula-

tion of left temporoparietal cortex and medication-resistant

auditory hallucinations. Arch Gen Psychiatry 60:49�/56.

Hoffman RE, Boutros NN, Hu S, Berman RM, Krystal JH,

Charney DS. 2000. Transcranial magnetic stimulation and

auditory hallucinations in schizophrenia. Lancet 355:1073�/

1075.

Holden JM, Itil TM, Keskiner A, Fink M. 1968. Thioridazine and

chlordiazepoxide, alone and combined, in the treatment of

chronic schizophrenia. Compr Psychiatry 9:633�/643.

Hong CJ, Chen JY, Chiu HJ, Sim CB. 1997. A double-blind

comparative study of clozapine versus chlorpromazine on

Chinese patients with treatment-refractory schizophrenia. Int

Clin Psychopharmacol 12:123�/130.

Honigfeld G, Patin J, Singer J. 1984. Clozapine antipsychotic

activity in treatment-resistant schizophrenics. Adv Ther 1:77�/

97.

Hori M, Suzuki T, Sasaki M, Shiraishi H, Koizumi J. 1992.

Convulsive seizures in schizophrenic patients induced by

zotepine administration. Jpn J Psychiatry Neurol 46:161�/167.

Hoyberg OJ, Fensbo C, Remvig J, Lingjaerde O, Sloth-Nielsen M,

Salvesen I. 1993. Risperidone versus perphenazine in the

treatment of chronic schizophrenic patients with acute exacer-

bations. Acta Psychiatr Scand 88:395�/402.

Hunt GE, Bergen J, Bashir M. 2002. Medication compliance and

comorbid substance abuse in schizophrenia: Impact on com-

munity survival 4 years after a relapse. Schizophr Res 54:253�/

264.

Huq ZU, RIS-GBR-32 Investigators. 2004. A trial of low doses of

risperidone in the treatment of patients with first-episode

schizophrenia, schizophreniform disorder, or schizoaffective

disorder. J Clin Psychopharmacol 24(2):220�/224.

Huttunen M, Piepponen T, Rantanen H, Larmo I, Nyholm R,

Raitasuo V. 1995. Risperidone versus zuclopenthixol in the

treatment of acute schizophrenic episodes: A double-blind

parallel group trial. Acta Psychiatr Scand 91:271�/277.

Hwang TJ, Lin SK, Lin HN. 2001. Efficacy and safety of zotepine

for the treatment of Taiwanese schizophrenicpatients: A

double-blind comparison with haloperidol. J Formos Med

Assoc 100:811�/816.

Hwang TJ, Lee SM, Sun HJ, Lin HN, Tsai SJ, Lee YC, Chen YS.

2003. Amisulpride versus risperidone in the treatment of

schizophrenic patients: A double-blind pilot study in Taiwan.

J Formos Med Assoc 102:30�/36.

Ishigooka J, Inada T, Miura S. 2001. Olanzapine versus haloper-

idol in the treatment of patients with chronic schizophrenia:

Results of the Japan multicenter, double-blind olanzapine trial.

Psychiatry Clin Neurosci 55:403�/414.

Jablensky A, Sartorius N, Ernberg G, Anker M, Korten A, Cooper

JE, Day R, Bertelsen A. 1992. Schizophrenia: Manifestations,

incidence, and course in different cultures. A World Health

Organization ten-country study. Psychol Med Monogr Suppl

20:1�/97.

Janicak P, Davis J, Preskorn S, Ayd F. 1993. Principles and

practice of psychopharmacotherapy. Baltimore. MD: Williams

& Wilkins.

Javitt DC, Silipo G, Cienfuegos A, Shelley AM, Bark N, Park M,

Lindenmayer JP, Suckow R, Zukin SR. 2001. Adjunctive high-

dose glycine in the treatment of schizophrenia. Int J Neurop-

sychopharmacol 4:385�/391.

Jeste DV, Lacro JP, Gilbert PL, Kline J, Kline N. 1993. Treatment

of late-life schizophrenia with neuroleptics. Schizophr Bull

19(4):817�/830.

Jeste DV, Gladsjo JA, Lindamer LA, Lacro JP. 1996. Medical

comorbidity in schizophrenia. Schizophr Bull 22:413�/430.

Jeste DV. 2000. Tardive dyskinesia in older patients. J Clin

Psychiatry 61(Suppl 4):27�/32.

Johnstone EC, Crow TJ, Johnson AL, MacMillan JF. 1986. The

Northwick Park Study of first episodes of schizophrenia, I:

Presentation of the illness and problems relating to admission.

Br J Psychiatry 148:115�/120.

Johnstone EC, Conelly J, Frith CD, Lambert MT, Owens DG.

1996. The nature of transient and partial psychoses: Findings

from the Northwick Park Study Functional Psychosis Study.

Psychol Med 26:361�/369.

Josiassen RC, Joseph A, Kohegyi E, Stokes S, Dadvand M, Paing

WW, Shaughnessy RA. 2005. Clozapine augmented with

risperidone in the treatment of schizophrenia: A randomized,

double-blind, placebo-controlled trial. Am J Psychiatry

162(1):130�/136.

182 P. Falkai et al.

Joy CB, Adams CE, Lawrie SM. 2004. Haloperidol versus

placebo for schizophrenia (Cochrane Review). In: The Co-

chrane Library, Issue 2. Chichester, UK: John Wiley & Sons

Ltd.

Kalinowsky LB, Worthing HJ. 1943. Results with electroconvul-

sive therapy in 200 cases of schizophrenia. Psychiatr Q 17:144�/

153.

Kalinowsky LB. 1943. Electric convulsive therapy, with emphasis

on importance of adequate treatment. Arch Neurol Psychiatry

50:652�/660.

Kane J, Honigfeld G, Singer J, Meltzer H. 1988. Clozapine for the

treatment-resistant schizophrenic: A double-blind comparison

with chlorpromazine. Arch Gen Psychiatry 45:789�/796.

Kane JM, Marder SR. 1993. Psychopharmacologic treatment of

schizophrenia. Schizophr Bull 19(2):287�/302.

Kane J. 1994. The use of higher-dose antipsychotic medication.

Br J Psychiatry 164:431�/432.

Kane JM, Marder SR, Schooler NR, Wirshing WC, Umbricht D,

Baker RW, Wirshing DA, Safferman A, Ganguli R, McMeni-

man M, Borenstein M. 2001. Clozapine and haloperidol in

moderately refractory schizophrenia: A 6-month randomized

and double-blind comparison. Arch Gen Psychiatry 58:965�/

972.

Kane JM, Carson WH, Saha AR, McQuade RD, Ingenito GG,

Zimbroff DL, Ali MW. 2002. Efficacy and safety of aripiprazole

and haloperidol versus placebo in patients with schizophrenia

and schizoaffective disorder. J Clin Psychiatry 63:763�/771.

Kane JM, Leucht S, Carpenter D, Docherty JP, editors. 2003.

The Expert Consensus Guideline Series. Optimizing Pharma-

cologic Treatment of Psychotic Disorders. J Clin Psychiatry

63(Suppl 12):1�/100.

Karagianis JL, LeDrew KK, Walker DJ. 2003. Switching treat-

ment-resistant patients with schizophrenia or schizoaffective

disorder to olanzapine: A one-year open-label study with five-

year follow-up. Curr Med Res Opin 19:473�/480.

Kasckow JW, Mohamed S, Thallasinos A, Carroll B, Zisook S,

Jeste DV. 2001. Citalopram augmentation of antipsychotic

treatment in older schizophrenia patients. Int J Geriatr Psy-

chiatry 16:1163�/1167.

Kasper S, Lerman MN, McQuade RD, Saha A, Carson WH, Ali

M, Archibald D, Ingenito G, Marcus R, Pigott T. 2003.

Efficacy and safety of aripiprazole vs. haloperidol for long-

term maintenance treatment following acute relapse of schizo-

phrenia. Int J Neuropsychopharmacol 6:325�/337.

Kaye NS. 2003. Ziprasidone augmentation of clozapine in 11

patients. J Clin Psychiatry 64:215�/216.

Keck P Jr, Buffenstein A, Ferguson J, Feighner J, Jaffe W,

Harrigan EP, Morrissey MR. 1998. Ziprasidone 40 and 120

mg/day in the acute exacerbation of schizophrenia and schi-

zoaffective disorder: A 4-week placebo-controlled trial. Psy-

chopharmacology (Berlin) 140:173�/184.

Keck PE Jr, Strakowski SM, McElroy SL. 2000. The efficacy of

atypical antipsychotics in the treatment of depressive symp-

toms, hostility, and suicidality in patients with schizophrenia. J

Clin Psychiatry 61:4�/9.

Kee KS, Kern RS, Marshall BD, Green MF. 1998. Risperidone

versus haloperidol for perception of emotion in treatment-

resistant schizophrenia: prelimimary findings. Schizophr Res

31:159�/165.

Keefe RS, Silva SG, Perkins DO, Lieberman JA. 1999. The effects

of atypical antipsychotic drugs on neurocognitive impairment

in schizophrenia: A review and meta-analysis. Schizophr Bull

25(2):201�/222.

Keefe RS, Seidman LJ, Christensen BK, Hamer RM, Sharma T,

Sitskoorn MM, Lewine RR, Yurgelun-Todd DA, Gur RC,

Tohen M, Tollefson GD, Sanger TM, Lieberman JA. 2004.

Comparative effect of atypical and conventional antipsychotic

drugs on neurocognition in first-episode psychosis: A rando-

mized, double-blind trial of olanzapine versus low doses of

haloperidol. Am J Psychiatry 161(6):985�/995.

Kelly DL, Gale EA, Conley RR. 2003. Clozapine treatment in

patients with prior substance abuse. Can J Psychiatry 48:111�/

114.

Killian JG, Kerr K, Lawrence C, Celermajer DS. 1999. Myocar-

ditis and cardiomyopathy associated with clozapine. Lancet

354:1841�/1845.

Kinon B, Kane J, Johns C, Perovich R, Ismi M, Koreen A, Weiden

P. 1993. Treatment of neuroleptic-resistant schizophrenic

relapse. Psychopharmacol Bull 29:309�/314.

Kinon BJ, Lieberman JA. 1996. Mechanisms of action of atypical

antipsychotic drugs: A critical analysis. Psychopharmacology

(Berlin) 124:2�/34.

Kinon BJ, Ahl J, Stauffer VL, Hill AL, Buckley PF. 2004. Dose

response and atypical antipsychotics in schizophrenia. CNS

Drugs 18(9):597�/616.

Kirli S, Caliskan M. 1998. A comparative study of sertraline

versus imipramine in postpsychotic depressive disorder of

schizophrenia. Schizophr Res 33:103�/111.

Kissling W, editor. 1991. Guidelines for neuroleptic relapse

prevention in schizophrenia. Berlin: Springer Verlag.

Klein E, Kolsky Y, Puyerovsky M, Koren D, Chistyakov A,

Feinsod M. 1999. Right prefrontal slow repetitive transcranial

magnetic stimulation in schizophrenia: A double-blind sham-

controlled pilot study. Biol Psychiatry 46:1451�/1454.

Kleinberg DL, Davis JM, de Coster R, Van Baelen B, Brecher M.

1999. Prolactin levels and adverse events in patients treated

with risperidone. J Clin Psychopharmacol 19:57�/61.

Klieser E, Lehmann E. 1987. Experimental comparison of the

effectivity of individually adapted and standardized dosages of

haloperidol. Pharmacopsychiatry 18:122�/126.

Klieser E, Strauss WH, Lemmer W. 1994. The tolerability and

efficacy of the atypical neuroleptic remoxipride compared with

clozapine and haloperidol in acute schizophrenia. Acta Psy-

chiatr Scand 89(Suppl 380):68�/73.

Klieser E, Lehmann E, Kinzler E, Wurthmann C, Heinrich KJ.

1995. Randomized, double-blind, controlled trial of risperi-

done versus clozapine in patients with chronic schizophrenia.

Clin Psychopharmacol 15(1 Suppl 1):45�/51S.

Klimke A, Klieser E, Lehmann E, Miele L. 1993. Initial

improvement as a criterion for drug choice in acute schizo-

phrenia. Pharmacopsychiatry 26:25�/29.

Kofoed L, Kania J, Walsh T, Atkinson RM. 1986. Outpatient

treatment of patients with substance abuse and coexisting

psychiatric disorders. Am J Psychiatry 143:867�/872.

Konrad C, Schormair C, Ophaus P, Knickelbein U, Eikelmann B.

2000. Risperidon und Clozapin in der Behandlung thera-

pieresistenter Schizophrenien. Krankenhauspsychiatrie 11:

89�/93.

Kopala LC, Good KP, Honer WG. 1997. Extrapyramidal signs

and clinical symptoms in first-episode schizophrenia: Response

to low-dose risperidone. J Clin Psychopharmacol 17(4):308�/

313.

Kopala LC, Caudle C. 1998. Acute and longer-term effects of

risperidone in a case of first-episode catatonic schizophrenia. J

Psychopharmacol 12(3):314�/317.

Kosten TR, Kleber HD. 1988. Rapid death during cocain abuse:

Variant of the neuroleptic malignant syndrome? Am J Drug

Alcohol Abuse 14:335�/346.

Kotler M, Strous RD, Reznik I, Shwartz S, Weizman A, Spivak B.

2004. Sulpiride augmentation of olanzapine in the manage-

ment of treatment-resistant chronic schizophrenia: evidence for

improvement of mood symptomatology. Int Clin Psychophar-

macol 19:23�/26.

WFSBP Guidelines for Biological Treatment of Schizophrenia 183

Kovasznay B, Fleischer J, Tanenberg-Karant M, Jandorf L, Miller

AD, Bromet E. 1997. Substance use disorder and the early

course of illness in schizophrenia and affective psychosis.

Schizophr Bull 23:195�/201.

Knapp M. 1997. Costs of schizophrenia. Br J Psychiatry 171:

509�/518.

Konig P, Glatter-Gotz U. 1990. Combined electroconvulsive and

neuroleptic therapy in schizophrenia refractory to neuroleptics.

Schizophr Res 3:351�/354.

Kolivakis TT, Beauclair L, Margolese HC, Chouinard G. 2004.

Long-term lamotrigine adjunctive to antipsychotic monother-

apy in schizophrenia: Further evidence. Can J Psychiatry

49(4):280.

Koreen AR, Siris SG, Chakos M, Alvir J, Mayerhoff D, Lieber-

man J. 1993. Depression in first-episode schizophrenia. Am J

Psychiatry 150:1643�/1648.

Koren G, Cohn T, Chitayat D, Kapur B, Remington G, Reid

DM, Zipursky RB. 2002. Use of atypical antipsychotics during

pregnancy and the risk of neural tube defects in infants. Am J

Psychiatry 159:136�/137.

Koro CE, Fedder DO, L’Italien GJ, Weiss SS, Magder LS,

Kreyenbuhl J, Revicki DA, Buchanan RW. 2002a. Assessment

of independent effect of olanzapine and risperidone on risk of

diabetes among patients with schizophrenia: Population based

nested case-control study. Br Med J 325(7358):243.

Koro CE, Fedder DO, L’Italien GJ, Weiss S, Magder LS,

Kreyenbuhl J, Revicki D, Buchanan RW. 2002b. An assessment

of the independent effects of olanzapine and risperidone

exposure on the risk of hyperlipidemia in schizophrenic

patients. Arch Gen Psychiatry 59(11):1021�/1026.

Kossen M, Selten JP, Kahn RS. 2001. Elevated clozapine plasma

level with lamotrigine. Am J Psychiatry 158(11):1930.

Krakowski M, Czobor P, Volavka J. 1997. Effect of neuroleptic

treatment on depressive symptoms in acute schizophrenic

episodes. Psychiatry Res 71:19�/26.

Kremer I, Vass A, Gorelik I, Bar G, Blanaru M, Javitt DC,

Heresco-Levy U. 2004. Placebo-controlled trial of lamotrigine

added to conventional and atypical antipsychotics in schizo-

phrenia. Biol Psychiatry 56:441�/446.

Krystal JH, D’Souza DC, Madonick S, Petrakis IL. 1999. Toward

a rational pharmacotherapy of comorbid substance abuse in

schizophrenic patients. Schizophr Res 35:S35�/49.

Kumra S, Frazier JA, Jacobsen LK, McKenna K, Gordon CT,

Lenane MC, Hamburger SD, Smith AK, Albus KE, Alagh-

band-Rad J, Rapoport JL. 1996. Childhood-onset schizophre-

nia: a double-blind clozapine-haloperidol comparison. Arch

Gen Psychiatry 53:1090�/1097.

Kupchik M, Spivak B, Mester R, Reznik I, Gonen N, Weizman A,

Kotler M. 2000. Combined electroconvulsive-clozapine ther-

apy. Clin Neuropharmacol 23(1):14�/16.

Lacro JP, Dunn LB, Dolder CR, Leckband SG, Jeste DV. 2002.

Prevalence of and risk factors for medication nonadherence in

patients with schizophrenia: A comprehensive review of recent

literature. J Clin Psychiatry 63:892�/909.

La Grenade L, Graham D, Trontell A. 2001. Myocarditis and

cardiomyopathy associated with clozapine use in the United

States. New Engl J Med 345:224�/225.

Landmark J, Joseph L, Merskey H. 1987. Characteristics of

schizophrenic patients and the outcome of fluphenazine and of

electroconvulsive treatments. Can J Psychiatry 32:425�/428.

Lausberg H, Hellweg R. 1998. Catatonic dilemma. Therapy with

lorazepam and clozapine. Nervenarzt 69(9):818�/822.

Lee ML, Dickson RA, Campbell M, Oliphant J, Gretton H, Dalby

JT. 1998. Clozapine and substance abuse in patients with

schizophrenia. Can J Psychiatry 43:855�/856 (letter).

Lee MS, Kim YK, Lee SK, Suh KY. 1998. A double-blind study

of adjunctive sertraline in haloperidol-stabilized patients with

chronic schizophrenia. J Clin Psychopharmacol 18:399�/403.

Lee JW, Schwartz DL, Hallmayer J. 2000. Catatonia in a

psychiatric intensive care facility: Incidence and response to

benzodiazepines. Ann Clin Psychiatry 12:89�/96.

Lehmann AF, Steinwachs DM, PORT Co-investigators. 1998.

Translating research into practice: The Schizophrenia Patient

Outcomes Research Team (PORT) treatment recommenda-

tions. Schizophr Bull 24:1�/10.

Lehman AF, Lieberman JA, Dixon LB, McGlashan TH, Miller

AL, Perkins DO, Kreyenbuhl J. 2004. Practice guideline for the

treatment of patients with schizophrenia, 2nd ed. Am J

Psychiatry 161:1�/56.

Lerner V, Chudakova B, Kravets S, Polyakova I. 2000. Combined

use of risperidone and olanzapine in the treatment of patients

with resistant schizophrenia: A preliminary case series report.

Clin Neuropharmacol 23:284�/286.

Leucht S, Hartung B. 2004a. Benperidol for schizophrenia

(Cochrane Review). In: The Cochrane Library, Issue 2.

Chichester, UK: John Wiley & Sons Ltd.

Leucht S, Hartung B. 2004b. Perazine for schizophrenia (Co-

chrane Review). In: The Cochrane Library, Issue 2. Chiche-

ster, UK: John Wiley & Sons Ltd.

Leucht S, Pitschel-Walz G, Abraham D, Kissling W. 1999.

Efficacy and extrapyramidal side-effects of the new antipsycho-

tics olanzapine, quetiapine, risperidone, and sertindole com-

pared to conventional antipsychotics and placebo: A meta-

analysis of randomized controlled trials. Schizophr Res 35:51�/

68.

Leucht S, Pitschel-Walz G, Engel RR, Kissling W. 2002.

Amisulpride, an unusual ‘atypical’ antipsychotic: A meta-

analysis of randomized controlled trials. Am J Psychiatry

159(2):180�/190.

Leucht S, Wahlbeck K, Hamann J, Kissling W. 2003a. New

generation antipsychotics versus low-potency conventional

antipsychotics: A systematic review and meta-analysis. Lancet

361(9369):1581�/1589.

Leucht S, Barnes TR, Kissling W, Engel RR, Correll C, Kane JM.

2003b. Relapse prevention in schizophrenia with new-genera-

tion antipsychotics: A systematic review and exploratory meta-

analysis of randomized, controlled trials. Am J Psychiatry

160:1209�/1222.

Leucht S, Kissling W, McGrath J. 2004. Lithium for schizo-

phrenia revisited: A systematic review and meta-analysis of

randomized controlled trials. J Clin Psychiatry 65:177�/186.

Leung A, Chue P. 2000. Sex differences in schizophrenia: A

review of the literature. Acta Psychiatr Scand (Suppl) 401:3�/

38.

Levin RF, Evans SM, Coomaraswammy S, Collins ED, Regent N,

Kleber HD. 1998. Flupentixol treatment for cocaine abusers

with schizophrenia: A pilot study. Am J Drug Alcohol Abuse

24:343�/360.

Levinson DF, Umapathy C, Musthaq M. 1999. Treatment of

schizoaffective disorder and schizophrenia with mood symp-

toms. Am J Psychiatry 156:1138�/1148.

Lewis S, Tarrier N, Haddock G, Bentall R, Kinderman P,

Kingdon D, Siddle R, Drake R, Everitt J, Leadley K, Benn

A, Grazebrook K, Haley C, Akhtar S, Davies L, Palmer S,

Faragher B, Dunn G. 2002. Randomised controlled trial of

cognitive-behavioral therapy in early schizophrenia: Acute-

phase outcomes. Br J Psychiatry (Suppl) 43:S91�/97.

Lieberman JA, Safferman AZ, Pollack S, Szymanski S, Johns S,

Howard A, Kronig M, Bookstein P, Kane JM. 1994. Clinical

effects of clozapine in chronic schizophrenia: Response to

treatment and predictors of outcome. Am J Psychiatry

151:1744�/1752.

184 P. Falkai et al.

Lieberman JA, Koreen AR, Chakos M, Sheitman B, Woerner M,

Alvir JM, Bilder R. 1996. Factors influencing treatment

response and outcome of first-episode schizophrenia: Implica-

tions for understanding the pathophysiology of schizophrenia. J

Clin Psychiatry 57(Suppl 9):5�/9.

Lieberman JA, Phillips M, Gu H, Stroup S, Zhang P, Kong L, Ji

Z, Koch G, Hamer RM. 2003a. Atypical and conventional

antipsychotic drugs in treatment-naive first-episode schizo-

phrenia: A 52-week randomized trial of clozapine vs chlorpro-

mazine. Neuropsychopharmacology 28:995�/1003.

Lieberman JA, Tollefson G, Tohen M, Green AI, Gur RE, Kahn

R, McEvoy J, Perkins D, Sharma T, Zipursky R, Wei H, Hamer

RM. 2003b. Comparative efficacy and safety of atypical and

conventional antipsychotic drugs in first-episode psychosis: A

randomized, double-blind trial of olanzapine versus haloper-

idol. Am J Psychiatry 160:1396�/1404.

Lindenmayer JP, Grochowski S, Mabugat L. 1994. Clozapine

effects on positive and negative symptoms: a six-month trial in

treatment-refractory schizophrenics. J Clin Psychopharmacol

14:201�/204.

Lindenmayer JP, Iskander A, Park M, Apergi FS, Czobor P, Smith

R, Allen D. 1998. Clinical and neurocognitive effects of

clozapine and risperidone in treatment-refractory schizophre-

nic patients: A prospective study. J Clin Psychiatry 59:521�/

527.

Lindenmayer JP, Volavka J, Lieberman J, Sheitman B, Citrome L,

Chakos M, Czobor P, Parker B, Iskander A. 2001. Olanzapine

for schizophrenia refractory to typical and atypical antipsycho-

tics: An open-label, prospective trial. J Clin Psychopharmacol

21:448�/453.

Lindenmayer JP, Czobor P, Volavka J, Lieberman JA, Citrome L,

Sheitman B, Chakos M, McEvoy JP. 2002. Olanzapine in

refractory schizophrenia after failure of typical or atypical

antipsychotic treatment: An open-label switch study. J Clin

Psychiatry 63:931�/935.

Lingjaerde O, Engstrand E, Ellingsen P, Stylo DA, Robak OH.

1979. Antipsychotic effect of diazepam when given in additino

to neuroleptics in chronic psychotic patients: A double-blind

clinical trial. Curr Ther Res 26:505�/512.

Linszen DH, Dingemans PM, Lenior ME. 1994. Cannabis abuse

and course of recent-onset schizophrenic disorders. Arch Gen

Psychiatry 51:273�/279.

Littrell KH, Petty RG, Hilligoss NM, Peabody CD, Johnson CG.

2001. Olanzapine treatment for patients with schizophrenia

and substance abuse. J Subst Abuse Treat 21:217�/221.

Loebel AD, Lieberman JA, Alvir JM, Mayerhoff DI, Geisler SH,

Szymanski SR. 1992. Duration of psychosis and outcome in

first episode schizophrenia. Am J Psychiatry 149:1183�/1188.

Loo H, Poirier-Littre MF, Theron M, Rein W, Fleurot O. 1997.

Amisulpride versus placebo in the medium-term treatment of

the negative symptoms of schizophrenia. Br J Psychiatry

170:18�/22.

Louza MR, Marques AP, Elkis H, Bassitt D, Diegoli M, Gattaz

WF. 2004. Conjugated estrogens as adjuvant therapy in the

treatment of acute schizophrenia: A double-blind study.

Schizophr Res 66(2�/3):97�/100.

Luchins D. 1987. Carbamazepine in violent non-epileptic schizo-

phrenics. Psychopharmacol Bull 20:569�/571.

Luchins DJ, Klass D, Hanrahan P, Malan R, Harris J. 1998.

Alteration in the recommended dosing schedule for risper-

idone. Am J Psychiatry 155(3):365�/366.

MacKay FJ, Wilton GL, Pearce SN, Freemantle SN, Mann RD.

1998. The safety of risperidone: A post-marketing study on

7684 patients. Hum Psychopharmacol Clin Exp 13:413�/418.

Major LF, Lerner P, Ballenger JC, Brown GL, Goodwin FK,

Lovenberg W. 1978. Dopamine-b-hydroxylase in the cere-

brospinal fluid: Relationship to disulfiram-induced psychosis.

Biol Psychiatry 14:337�/343.

Marcus P, Snyder R. 1995. Reduction of comorbid substance

abuse with clozapine. Am J Psychiatry 152:959 (letter).

Marder SR, Meibach RC. 1994. Risperidone in the treatment of

schizophrenia. Am J Psychiatry 151:825�/835.

Marder SR, Essock SM, Miller AL, Buchanan RW, Davis JM,

Kane JM, Lieberman J, Schooler NR. 2002. The Mount Sinai

conference on the pharmacotherapy of schizophrenia. Schi-

zophr Bull 28(1):5�/16.

Marder SR, McQuade RD, Stock E, Kaplita S, Marcus R,

Safferman AZ, Saha A, Ali M, Iwamoto T. 2003. Aripiprazole

in the treatment of schizophrenia: Safety and tolerability in

short-term, placebo-controlled trials. Schizophr Res 61:123�/

136.

Marder SR, Essock SM, Miller AL, Buchanan RW, Casey DE,

Davis JM, Kane JM, Lieberman JA, Schooler NR, Covell N,

Stroup S, Weissman EM, Wirshing DA, Hall CS, Pogach L, Pi-

Sunyer X, Bigger JT Jr, Friedman A, Kleinberg D, Yevich SJ,

Davis B, Shon S. 2004. Physical health monitoring of patients

with schizophrenia. Am J Psychiatry 161(8):1334�/1349.

Marenco S, Weinberger DR. 2000. The neurodevelopmental

hypothesis of schizophrenia: Following a trail of evidence

from cradle to grave. Dev Psychopathol 12:501�/527.

Marques LO, Lima MS, Soares BGO. 2004. Trifluoperazine for

schizophrenia (Cochrane Review). In: The Cochrane Library,

Issue 2. Chichester, UK: John Wiley & Sons Ltd.

Maxwell S, Shinderman MS. 2000. Use of naltrexone in the

treatment of alcohol use disorders in patients with concomitant

major mental illness. J Addict Dis 19:61�/69.

McCarthy RH, Terkelsen KG. 1995. Risperidone augmentation

of clozapine. Phamacopsychiatry 28:61�/63.

McEvoy J, Freudenreich O, McGee M, VanderZwaag C, Levin E,

Rose J. 1995. Clozapine decreases smoking in patients with

chronic schizophrenia. Biol Psychiatry 37(8):550�/552.

McGlashan TH, Fenton WS. 1993. Subtype progression and

pathophysiologic deterioriation in early schizophrenia. Schi-

zophr Bull 19:71�/84.

McGlashan TH, Johannessen JO. 1996. Early detection and

intervention with schizophrenia: Rationale. Schizophr Bull

22:201�/222.

McGorry P, Killackey E, Elkins K, Lambert M, Lambert T for

the RANZCP Clinical Practice Guideline Team for the

treatment of schizophrenia. 2003. Summary Australian and

New Zealand clinical practice guideline for the treatment of

schizophrenia. Australasian Psychiatry 11(2):136�/147.

Meltzer HY, Okayli G. 1995. Reduction of suicidality during

clozapine treatment of neuroleptic-resistant schizophrenia:

Impact on risk-benefit assessment. Am J Psychiatry 152:183�/

190.

Meltzer HY, McGurk SR. 1999. The effects of clozapine,

risperidone, and olanzapine on cognitive function in schizo-

phrenia. Schizophr Bull 25(2):233�/255.

Meltzer HY. 2002. Suicidality in schizophrenia: A review of the

evidence for risk factors and treatment options. Curr Psychiatry

Rep 4:279�/283.

Meltzer HY, Alphs L, Green AI, Altamura AC, Anand R, Bertoldi

A, Bourgeois M, Chouinard G, Islam MZ, Kane J, Krishnan R,

Lindenmayer JP, Potkin S. 2003. Clozapine treatment for

suicidality in schizophrenia: International Suicide Prevention

Trial (InterSePT). Arch Gen Psychiatry 60:82�/91.

Merlo MC, Hofer H, Gekle W, Berger G, Ventura J, Panhuber I,

Latour G, Marder SR. 2002. Risperidone, 2 mg/day vs. 4 mg/

day, in first-episode, acutely psychotic patients: Treatment

efficacy and effects on fine motor functioning. J Clin Psychiatry

63:885�/891.

WFSBP Guidelines for Biological Treatment of Schizophrenia 185

Meyer-Lindenberg A, Gruppe H, Bauer U, Lis S, Krieger S,

Gallhofer B. 1997. Improvement of cognitive function in

schizophrenic patients receiving clozapine or zotepine: Results

from a double-blind study. Pharmacopsychiatry 30:35�/42.

Miller LJ. 1994. Use of electroconvulsive therapy during preg-

nancy. Hosp Com Psychiatry 45(5):444�/450.

Miller AL, Chiles JA, Chiles JK, Crismon ML, Rush AJ, Shon SP.

1999. The Texas Medication Algorithm Project (TMAP)

schizophrenia algorithms. J Clin Psychiatry 60(10):649�/657.

Miller DD. 2000. Review and management of clozapine side

effects. J Clin Psychiatry 61(Suppl 8):14�/17.

Moller HJ. 1996. Treatment of schizophrenia: State of the art. Eur

Arch Psychiatry Clin Neurosci 246:229�/234.

Moller HJ. 2000a. Definition, psychopharmacological basis and

clinical evaluation of novel/atypical neuroleptics: Methodologi-

cal issues and clinical consequences. World J Biol Psychiatry

1:75�/91.

Moller HJ. 2000b. State of the art of drug treatment of

schizophrenia and the future position of the novel/atypical

antipsychotics. World J Biol Psychiatry 1:204�/214.

Moller HJ, Bottlender R, Groß A, Hoff P, Wittmann J, Wegner U,

Strauss A. 2002. The Kraepelinian dichotomy: Preliminary

results of a 15-year follow-up study on functional psychoses:

Focus on negative symptoms. Schizophr Res 56:87�/94.

Moller HJ. 2003. Management of the negative symptoms of

schizophrenia. New treatment options. CNS Drugs 17(11):

793�/823.

Moller HJ. 2004a. Novel antipsychotics in the long-term treat-

ment of schizophrenia. World J Biol Pschiatry 5:9�/19.

Moller HJ. 2004b. Non-neuroleptic approaches to treating

negative symptoms in schizophrenia. Eur Arch Psychiatry

Clin Neurosci 254:108�/116.

Moller HJ. 2005a. Antidepressive effects of traditional and second

generation antipsychotics: A review of the clinical data. Eur

Arch Psychiatry Clin Neurosci 255:83�/96.

Moller HJ. 2005b. Occurrence and treatment of depressive

comorbidity/cosyndromality in schizophrenic psychoses. World

J Biol Psychiatry (in press).

Moller HJ, Muller H, Borison RL, Schooler NR, Chouinard G.

1995. A path-analytical approach to differentiate between

direct and indirect drug effects on negative symptoms in

schizophrenic patients. A re-evaluation of the North American

risperidone study. Eur Arch Psychiatry Clin Neurosci 245:45�/

49.

Moller HJ, Boyer P, Fleurot O, Rein W. 1997. Improvement of

acute exacerbations of schizophrenia with amisulpride: A

comparison with haloperidol. Psychopharmacology 132:396�/

401.

Moller HJ, Riedel M, Muller N, Fischer W, Kohnen R. 2004.

Zotepine versus placebo in the treatment of schizophrenic

patients with stable primary negative symptoms: A randomized

double-blind multicenter trial. Pharmacopsychiatry 37:270�/

278.

Morera AL, Barreiro P, Cano-Munoz JL. 1999. Risperidone and

clozapine combination for the treatment of refractory schizo-

phrenia. Acta Psychiatr Scand 99:305�/306.

Morgenstern H, Glazer WM. 1993. Identifying risk factors for

tardive dyskinesia among long-term outpatients maintained

with neuroleptic medication. Results of the Yale Tardive

Dyskinesia Study. Arch Gen Psychiatry 50:723�/733.

Morinigo A, Martin J, Gonzalez S, Mateo I. 1989. Treatment of

resistant schizophrenia with valproate and neuroleptic drugs.

Hillside J Clin Psychiatry 11:199�/207.

Morris S, Bagnall A, Cooper SJ, DeSilva P, Fenton M, Gammelin

GO, Leitner M. 2004. Zotepine for schizophrenia (Cochrane

Review). In: The Cochrane Library, Issue 2. Chichester, UK:

John Wiley & Sons Ltd.

Mortimer A, Martin S, Loo H, Peuskens J, SOLIANOL Study

Group. 2004. A double-blind, randomized comparative trial of

amisulpride versus olanzapine for 6 months in the treatment of

schizophrenia. Int Clin Psychopharmacol 19:63�/69.

Mota Neto JIS, Lima MS, Soares BGO. 2003. Amisulpride for

schizophrenia (Cochrane Review). In: The Cochrane Library,

Issue 2. Oxford: Update Software.

Mulholland C, Cooper S. 2000. The symptom of depression and

its management. Adv Psychiatr Treatment 6:169�/177.

Mueser KT, Yarnold PR, Levinson DR, Singh H, Bellack AS, Kee

K, Morrison RL, Yadalam DG. 1990. Prevalence of substance

abuse in schizophrenia: Demographic and clinical correlates.

Schizophr Bull 16:31�/56.

Mueser KT, Noordsy DL, Fox L, Wolfe R. 2003. Disulfiram

treatment for alcoholism in severe mental illness. Am J Addict

12(3):242�/252.

Murray CJ, Lopez AD. 1997. Global mortality, disability and the

contribution of risk factors: Global Burden of Disease Study.

Lancet 17(349):1436�/1442.

Myers DH, Campbell PL, Cocks NM, Flowerdew JA, Muir A.

1981. A trial of propranolol in chronic schizophrenia. Br J

Psychiatry 139:118�/121.

National Institutes of Health Psychopharmacology Service Center

Collaborative Study Group. 1964. Phenothiazine treatment in

acute schizophrenia. Arch Gen Psychiatry 10:246�/261.

National Institute for Clinical Excellence. 2003. Core Interven-

tions in the Treatment of Schizophrenia. NICE, London

(www.nice.org.uk).

National Institute for Clinical Excellence. 2002. Guidance on the

use of newer (atypical) antipsychotic drugs for the treatment of

schizophrenia. Technology Appraisal Guidance No. 43, Lon-

don (www.nice.org.uk).

Noorbala AA, Akhondzadeh S, Davari-Ashtiani R, Amini-

Nooshabadi H. 1999. Piracetam in the treatment of schizo-

phrenia: Implications for the glutamate hypothesis of schizo-

phrenia. J Clin Pharm Ther 24(5):369�/374.

Noordsy DL, O’Keefe C. 1999. Effectiveness of combining

atypical antipsychotics and psychosocial rehabilitation in a

community mental health center setting. J Clin Psychiatry

60(Suppl 19):47�/51.

Noordsy DL, O’Keefe C, Mueser KT, Xie H. 2001. Six-month

outcomes for patients who switched to olanzapine treatment.

Psychiatr Serv 52:501�/507.

Noordsy DL, Green AI. 2003. Pharmacotherapy for schizophre-

nia and co-occuring substance use disorders. Curr Psychiatry

Rep 5:340�/346.

Nordentoft M, Jeppesen P, Abel M, Kassow P, Petersen L,

Thorup A, Krarup G, Hemmingsen R, Jorgensen P. 2002.

OPUS study: Suicidal behaviour, suicidal ideation and hope-

lessness among patients with first-episode psychosis. One-year

follow-up of a randomised controlled trial. Br J Psychiatry

Suppl 43:s98�/106.

Nuechterlein KH, Dawson ME, Ventura J, Gitlin M, Subotnik

KL, Snyder KS, Mintz J, Bartzokis G. 1994. The vulnerability/

stress model of schizophrenic relapse. Acta Psychiatr Scand

Suppl 382:58�/64.

Okuma T, Yamashita I, Takahashi R, Itoh H. 1989. A double-

blind study of adjunctive carbamazepine versus placebo on

excited states of schizophrenic and schizoaffective disorders.

Acta Psychiatr Scand 80(3):250�/259.

Olivera AA, Kiefer MW, Manley NK. 1990. Tardive dyskinesia in

psychiatric patients with substance abuse. Am J Alcohol Abuse

16:57�/66.

Oosthuizen P, Emsley RA, Turner J, Keyter N. 2001. Determin-

ing the optimal dose of haloperidol in first-episode psychosis. J

Psychopharmacol 15(4):251�/255.

186 P. Falkai et al.

Oosthuizen P, Emsley RA, Roberts MC, Turner J, Keyter L,

Keyter N, Torreman M. 2002. Depressive symptoms at base-

line predict fewer negative symptoms at follow-up in patients

with first-episode schizophrenia. Schizophr Res 58:247�/252.

Oosthuizen PP, Emsley RA, Maritz JS, Turner JA, Keyter N.

2003. Incidence of tardive dyskinesia in first-episode psychosis

patients treated with low-dose haloperidol. J Clin Psychiatry

64(9):1075�/1080.

Oosthuizen P, Emsley R, Jadri Turner H, Keyter N. 2004. A

randomized, controlled comparison of the efficacy and toler-

ability of low and high doses of haloperidol in the treatment of

first-episode psychosis. Int J Neuropsychopharmacol 7(2):

125�/131.

Osser DN, Sigadel R. 2001. Short-term inpatient pharmacother-

apy of schizophrenia. Harv Rev Psychiatry 9:89�/104.

Pailliere-Martinot ML, Lecrubier Y, Martinot JL, Aubin F. 1995.

Improvement of some schizophrenic deficit symptoms with low

doses of amisulpride. Am J Psychiatry 152:130�/133.

Pato CN, Wolkowitz OM, Rapaport M, Schulz SC, Pickar D.

1989. Benzodiazepine augmentation of neuroleptic treatment

in patients with schizophrenia. Psychopharmacol Bull 25:263�/

266.

Pelonero AL, Levenson JL, Pandurangi AK. 1998. Neuroleptic

malignant syndrome: a review. Psychiatr Serv 49:1163�/1172.

Perkins D, Lieberman J, Gu H, Tohen M, McEvoy J, Green A,

Zipursky R, Strakowski S, Sharma T, Kahn R, Gur R, Tollefson

G, HGDH Research Group. 2004. Predictors of antipsychotic

treatment response in patients with first-episode schizophrenia,

schizoaffective and schizophreniform disorders. Br J Psychiatry

185:18�/24.

Petit M, Raniwalla J, Tweed J, Leutenegger E, Dollfus S, Kelly F.

1996. A comparison of an atypical and typical antipsychotic,

zotepine versus haloperidol in patients with acute exacerbation

of schizophrenia: A parallel-group double-blind trial. Psycho-

pharmacol Bull 32:81�/87.

Petrides G, Divadeenam KM, Bush G, Francis A. 1997. Syner-

gism of lorazepam and electroconvulsive therapy in the treat-

ment of catatonia. Biol Psychiatry 42:375�/381.

Peuskens J, Risperidone Study Group. 1995. Risperidone in the

treatment of patients with chronic schizophrenia: A multi-

national, multi-centre, double-blind, parallel-group study ver-

sus haloperidol. Br J Psychiatry 166:712�/726.

Peuskens J, Link CG. 1997. A comparison of quetiapine and

chlorpromazine in the treatment of schizophrenia. Acta Psy-

chiatr Scand 96:265�/273.

Peuskens J, Bech P, Moller HJ, Bale R, Fleurot O, Rein W and the

Amisulpride Study Group. 1999. Amisulpride versus risper-

idone in the treatment of acute exacerbations of schizophrenia.

Psychiatr Res 88:107�/117.

Peuskens J, van Baelen B, de Smedt C, Lemmens P. 2000. Effects

of risperidone on affective symptoms in patients with schizo-

phrenia. Int Clin Psychopharmacol 15:343�/349.

Pichot P, Boyer P. 1988. Controlled double-blind multi-centre

trial of low dose amisulpride versus fluphenazine in the

treatment of the negative syndrome of chronic schizophrenia.

Ann Psychiatr 3(3 bis):312�/320.

Pigott TA, Carson WH, Saha AR, Torbeyns AF, Stock EG,

Ingenito GG. 2003. Aripiprazole for the prevention of relapse

in stabilized patients with chronic schizophrenia: A placebo-

controlled 26-week study. J Clin Psychiatry 64:1048�/1056.

Potkin SG, Jin Y, Bunney BG, Costa J, Gulasekaram B. 1999.

Effect of clozapine and adjunctive high-dose glycine in treat-

ment-resistant schizophrenia. Am J Psychiatry 156(1):145�/

147.

Potkin SG, Saha AR, Kujawa MJ, Carson WH, Ali M, Stock E,

Stringfellow J, Ingenito G, Marder SR. 2003. Aripiprazole, an

antipsychotic with a novel mechanism of action, and risper-

idone vs placebo in patients with schizophrenia and schizoaf-

fective disorder. Arch Gen Psychiatry 60:681�/690.

Potvin S, Stip E, Roy JY. 2003. Clozapine, quetiapine and

olanzapine among addicted schizophrenic patients: Towards

testable hypotheses. Int Clin Psychopharmacol 18:121�/132.

Poyurousky M, Bergman J, Weizman A. 1997. Risperidone in the

treatment of catatonia in a schizophrenic patient. Isr J

Psychiatry Relat Sci 34(4):323�/324.

Prior TI, Baker GB. 2003. Interactions between the cytochrome

P450 system and the second-generation antipsychotics. J

Psychiatry Neurosci 28:99�/112.

Procyshyn RM, Ihsan N, Thompson D. 2001. A comparison of

smoking behaviours between patients treated with clozapine

and depot neuroleptics. Int Clin Psychopharmacol 16(5):291�/

294.

Procyshyn RM, Tse G, Sin O, Flynn S. 2002. Concomitant

clozapine reduces smoking in patients treated with risperidone.

Eur Neuropsychopharmacol 12(1):77�/80.

Prudic J, Peyser S, Sackeim HA. 2000. Subjective memory

complaints: A review of patient self-assessment of memory

after electroconvulsive therapy. J ECT 16:121�/132.

Puech A, Fleurot O, Rein W, Amisulpride Study Group. 1998.

Amisulpride, an atypical antipsychotic, in the treatment of

acute episodes of schizophrenia: A dose-ranging study versus

haloperidol. Acta Psychiatr Scand 98:65�/72.

Pugh CR, Steinert J, Priest RG. 1983. Propranolol in schizo-

phrenia: A double-blind, placebo-controlled trial of proprano-

lol as an adjunct to neuroleptic medication. Br J Psychiatry

143:151�/155.

Purdon SE, Jones BD, Stip E, Labelle A, Addington D, David SR,

Breier A, Tollefson GD (Canadian Collaborative Group for Re-

search in Schizophrenia). 2000. Neuropsychological change in

early phase schizophrenia during 12 months of treatment with

olanzapine, risperidone, or haloperidol. Arch Gen Psychiatry

57:249�/258.

Purdon SE, Malla A, Labelle A, Lit W. 2001. Neuropsychological

change in patients with schizophrenia after treatment with

quetiapine or haloperidol. J Psychiatry Neurosci 26:137�/149.

Rabinowitz T, Frankenburg FR, Centorrino F, Kando J. 1996.

The effect of clozapine on saliva flow rate: a pilot study. Biol

Psychiatry 40:1132�/1134.

Rabinowitz J, Hornik T, Davidson M. 2001. Rapid onset of

therapeutic effect of risperidone versus haloperidol in a double-

blind randomized trial. J Clin Psychiatry 62:343�/346.

Raggi MA, Mandrioli R, Sabbioni C, Pucci V. 2004. Atypical

antipsychotics: Pharmacokinetics, therapeutic drug monitoring

and pharmacological interactions. Curr Med Chem 11:279�/

296.

Raskin S, Durst R, Katz G, Zislin J. 2000. Olanzapine and

sulpiride: A preliminary study of combination/augmentation in

patients with treatment-resistant schizophrenia. J Clin Psycho-

pharmacol 20:500�/503.

Raskin S, Katz G, Zislin Z, Knobler HY, Durst R. 2000.

Clozapine and risperidone: Combination/augmentation treat-

ment of refractory schizophrenia: A preliminary observation.

Acta Psychiatr Scand 101:334�/336.

Regier DA, Farmer ME, Rae DS, Locke BZ, Keith SJ, Judd LL,

Goodwin FK. 1990. Comorbidity of mental disorders with

alcohol and other drug abuse: Results form the Epidemiologic

Catchment Area (ECA) Study. J Am Med Assoc 264:2511�/

2518.

Remington G, Kapur S, Zipursky B. 1998. Pharmacotherapy of

first-episode schizophrenia. Br J Psychiatry 172(Suppl 33):66�/

70.

Revicki DA, Genduso LA, Hamilton SH, Ganoczy D, Beasley

CM Jr. 1999. Olanzapine versus haloperidol in the treatment of

schizophrenia and other psychotic disorders: Quality of life and

WFSBP Guidelines for Biological Treatment of Schizophrenia 187

clinical outcomes of a randomized clinical trial. Qual Life Res

8:417�/426.

Rey JM, Walter G. 1997. Half a century of ECT use in young

people. Am J Psychiatry 154(5):595�/602.

Reznik I, Sirota P. 2000. An open study of fluvoxamine

augmentation of neuroleptics in schizophrenia with obsessive

and compulsive symptoms. Clin Neuropharmacol 23:157�/

160.

Rice EH, Sombrotto LB, Markowitz JC, Leon AC. 1994.

Cardiovascular morbidity in high-risk patients during ECT.

Am J Psychiatry 151:1637�/1641.

Risch SC, McGurk S, Horner MD, Nahas Z, Owens SD, Molloy

M, Gilliard C, Christie S, Markowitz JS, DeVane CL, Mintzer

J, George MS. 2001. A double-blind placebo-controlled case

study of the use of donepezil to improve cognition in a

schizoaffective disorder patient: Functional MRI correlates.

Neurocase 7:105�/110.

Robinson DG, Woerner MG, Alvir JM, Geisler S, Koreen A,

Sheitman B, Chakos M, Mayerhoff D, Bilder R, Goldman R,

Lieberman JA. 1999a. Predictors of treatment response from a

first episode of schizophrenia or schizoaffective disorder. Am J

Psychiatry 156:544�/549.

Robinson DG, Woerner MG, Alvir JM, Bilder R, Goldman R,

Geisler S, Koreen A, Sheitman B, Chakos M, Mayerhoff D,

Lieberman JA. 1999b. Predictors of relapse following response

from a first episode of schizophrenia or schizoaffective disorder.

Arch Gen Psychiatry 56:241�/247.

Rodriguez-Perez V, Lopez A, Blanco C, Pena C, Lopez A, Abel A,

Gomez Y, Ferreiro MJ, Rego C, Lopez A, Cudeiro F, Alvarez V,

Prieto R, Ciudad A. 2002. Olanzapine for the treatment of

chronic refractory schizophrenia: A 12-month follow-up nat-

uralistic study. Prog Neuropsychopharmacol Biol Psychiatry

26:1055�/1062.

Rollnik JD, Huber TJ, Mogk H, Siggelkow S, Kropp S, Dengler

R, Emrich HM, Schneider U. 2000. High frequency repetitive

transcranial magnetic stimulation (rTMS) of the dorsolateral

prefrontal cortex in schizophrenic patients. Neuroreport

11:4013�/4015.

Rosebush PI, Hildebrand AM, Furlong BG, Mazurek MF. 1990.

Catatonic syndrome in a general psychiatric inpatient popula-

tion: Frequency, clinical presentation, and response to loraze-

pam. J Clin Psychiatry 51:357�/362.

Rosenheck R, Cramer J, Xu W, Thomas J, Henderson W, Frisman

L, Fye C, Charney D (Department of Veterans Affairs

Cooperative Study Group on Clozapine in Refractory Schizo-

phrenia). 1997. A comparison of clozapine and haloperidol in

hospitalized patients with refractory schizophrenia. New Engl J

Med 337:809�/815.

Rosenheck R, Evans D, Herz L, Cramer J, Xu W, Thomas J,

Henderson W, Charney D. 1999. How long to wait for a

response to clozapine: A comparison of time course of response

to clozapine and conventional antipsychotic medication in

refractory schizophrenia. Schizophr Bull 25:709�/719.

Royal Australian and New Zealand College of Psychiatrists

(RANZCP). 2003. Australian and New Zealand clinical

practice guideline for the treatment of schizophrenia. Draft

only.

Ruther E, Blanke J. 1998. In: Helmchen H, Hippius H, Tolle R,

editors. Therapie mit Neuroleptika �/ Perazin. Stuttgart, New

York: Thieme Verlag. pp 65�/70.

Rummel C, Hamann J, Kissling W, Leucht S. 2003. New

generation antipsychotics for first-episode schizophrenia (Co-

chrane Review). In: The Cochrane Library, Issue 4.

Rupniak NM, Jenner P, Marsden CD. 1986. Acute dystonia

induced by neuroleptic drugs. Psychopharmacology (Berlin)

88:403�/419.

Rupp A, Keith SJ. 1993. The costs of schizophrenia: Assessing the

burden. Psychiatr Clin North Am 16:413�/423.

Rupprecht R, Soyka M, Grohmann R, Ruther E, Moller HJ.

2004. Considerations in the combination of clozapine and

benzodiazepines]. Nervenarzt 75:857�/860.

Ruskin P, Averbukh I, Belmaker RH, Dasberg H. 1979. Benzo-

diazepines in chronic schizophrenia. Biol Psychiatry 14:557�/

558.

Ryan MC, Collins P, Thakore JH. 2003. Impaired fasting glucose

tolerance in first-episode, drug-naive patients with schizophre-

nia. Am J Psychiatry 160(2):284�/289.

Saba G, Dumortier G, Kalalou K, Benadhira R, Degrassat K,

Glikman J, Januel D. 2002. Lamotrigine-clozapine combina-

tion in refractory schizophrenia: Three cases. J Neuropsychia-

try Clin Neurosci 14(1):86.

Sajatovic M, Meltzer HY. 1993. The effect of short-term

electroconvulsive treatment plus neuroleptics in treatment-

resistant schizophrenia and schizoaffective disorder. Convuls

Ther 9:167�/175.

Salokangas RK, Saarijarvi S, Taiminen T, Kallioniemi H, Lehto

H, Niemi H, Tuominen J, Ahola V, Syvalahti E. 1996.

Citalopram as an adjuvant in chronic schizophrenia: A dou-

ble-blind placebo-controlled study. Acta Psychiatr Scand

94:175�/180.

Salzman C, Solomon D, Miyawaki E, Glassman R, Rood L,

Flowers E, Thayer S. 1991. Parenteral lorazepam versus

parenteral haloperidol for the control of psychotic disruptive

behavior. J Clin Psychiatry 52:177�/180.

Sanger TM, Lieberman JA, Tohen M, Grundy S, Beasley C,

Tollefson GD. 1999. Olanzapine versus haloperidol treatment

in first-episode psychosis. Am J Psychiatry 156:79�/87.

Sartorius N, Fleischhacker W, Gjerris A, Kern U, Knapp M,

Leonhard BE, Lieberman A, Lopez-Ibor JJ, van Raay B,

Tornquist E, Twomey E. 2002. The usefulness and use of

second-generation antipsychotic medications �/ An update:.

Curr Opin Psychiatry 15(Suppl 1):S1�/51.

Scheller-Gilkey G, Woolwine BJ, Cooper I, Olumuyiwa G,

Moynes KA, Miller AH. 2003. Relationship of clinical symp-

toms and substance use in schizophrenia patients on conven-

tional versus atypical antipsychoticsAm J Drug Alcohol Abuse

29:553�/566.

Schilkrut R, Cabrera J, Morales E, Herrera L. 1988. Neuroleptics

in the treatment of drug dependence in schizophrenics. A study

with flupenthixol decanoate. Psychopharmacology 96

(Suppl):342 (Abstract No. 33.01.50).

Schmidt LG, Schussler G, Kappes CV, Muller-Oerlinghausen B.

1982. Vergleich einer hoher dosierten Haloperidol-Therapie

mit einer Perazin-Standard-Therapie bei akut schizophrenen

Patienten. Nervenarzt 53(9):530�/536.

Schmidt AW, Lebel LA, Howard HR Jr, Zorn SH. 2001.

Ziprasidone: A novel antipsychotic agent with a unique human

receptor binding profile. Eur J Pharmacol 425:197�/201.

Schutz G, Berk M. 2001. Reboxetine add on therapy to

haloperidol in the treatment of schizophrenia: A preliminary

double-blind randomized placebo-controlled study. Int Clin

Psychopharmacol 16(5):275�/278.

Scottish Intercollegiate Gudelines Network. 1998. Psychosocial

Interventions in the Management of Schizophrenia. Scottish

Intercollegiate Gudelines Network (SIGN), SIGN Publication

Number 30, www.show.scot.nhs.uk/sign/index.html.

Sechter D, Peuskens J, Fleurot O, Rein W, Lecrubier Y. 2002.

Amisulpride vs. risperidone in chronic schizophrenia: Results

of a 6-month double-blind study. Neuropsychopharmacology

27:1071�/1081.

Sedvall G, Farde L. 1995. Chemical brain anatomy in schizo-

phrenia. Lancet 346(8977):743�/749.

188 P. Falkai et al.

Seeman MV. 1986. Current outcome in schizophrenia: Women vs

men. Acta Psychiatr Scand 73:609�/617.

Sharma T. 1999. Cognitive effects of conventional and atypical

antipsychotic in schizophrenia. Br J Psychiatry 174(Suppl

38):44�/51.

Shekelle PG, Woolf SH, Eccles M, Grimshaw J. 1999. Developing

guidelines. Br Med J 318:593�/596.

Shiloh R, Zemishlany Z, Aizenberg D, Radwan M, Schwartz B,

Dorfman-Etrog P, Modai I, Khaikin M, Weizman A. 1997.

Sulpiride augmentation in people with schizophrenia partially

responsive to clozapine. A double-blind, placebo-controlled

study. Br J Psychiatry 171:569�/573.

Shopsin B, Klein H, Aaronsom M, Collora M. 1979. Clozapine,

chlorpromazine, and placebo in newly hospitalized, acutely

schizophrenic patients: A controlled, double-blind comparison.

Arch Gen Psychiatry 36(6):657�/664.

Simhandl C, Meszaros K, Denk E, Thau K, Topitz A. 1996.

Adjunctive carbamazepine or lithium carbonate in therapy-

resistant chronic schizophrenia. Can J Psychiatry 41(5):317.

Silver H, Nassar A. 1992. Fluvoxamine improves negative

symptoms in treated chronic schizophrenia: An add-on dou-

ble-blind, placebo-controlled study. Biol Psychiatry 31:698�/

704.

Silver H, Shmugliakov N. 1998. Augmentation with fluvoxamine

but not maprotiline improves negative symptoms in treated

schizophrenia: Evidence for a specific serotonergic effect from a

double-blind study. J Clin Psychopharmacol 18:208�/211.

Silver H, Barash I, Aharon N, Kaplan A, Poyurovsky M. 2000.

Fluvoxamine augmentation of antipsychotics improves negative

symptoms in psychotic chronic schizophrenic patients: A

placebo-controlled study. Int Clin Psychopharmacol 15:257�/

261.

Silver H, Nassar A, Aharon N, Kaplan A. 2003. The onset and

time course of response of negative symptoms to add-on

fluvoxamine treatment. Int Clin Psychopharmacol 18:87�/92.

Silver H. 2003. Selective serotonin reuptake inhibitor augmenta-

tion in the treatment of negative symptoms of schizophrenia.

Int Clin Psychopharmacol 18:305�/313.

Siris SG. 1990. Pharmacological treatment of substance-abusing

schizophrenic patients. Schizophr Bull 16:111�/122.

Siris SG, Bermanzohn PC, Mason SE, Shuwall MA. 1991.

Antidepressant for substance-abusing schizophrenic patients:

A minireview. Prog Neuropsychopharmacol Biol Psychiat

15:1�/13.

Siris SG, Mason SE, Bermanzohn PC, Shuwall MA, Aseniero

MA. 1993. Dual diagnosis/psychiatric comorbidity of drug

dependence: Epidemiology and treatment. Adjunctive imipra-

mine in substance-abusing dysphoric schizophrenic patients.

Psychopharmacol Bull 29:127�/133.

Siris SG, Bermanzohn PC, Mason SE, Shuwall MA. 1994.

Maintenance imipramine therapy for secondary depression in

schizophrenia: A controlled trial. Arch Gen Psychiatry 51:109�/

115.

Siris SG. 2000. Depression in schizophrenia: perspective in the

era of ‘atypical’ antipsychotic agents. Am J Psychiatry

157:1379�/1389.

Siris S, Pollack S, Bermanzohn P, Stronger R. 2000. Adjunctive

imipramine for a broader group of post-psychotic depressions

in schizophrenia. Schizophr Res 44:187�/192.

Siris SG. 2001. Suicide and schizophrenia. J Psychopharmacol

15(2):127�/135.

Small JG, Kellams JJ, Milstein V, Moore J. 1975. A placebo-

controlled study of lithium combined with neuroleptics in

chronic schizophrenic patients. Am J Psychiatry 132:1315�/

1317.

Small JG, Hirsch SR, Arvanitis LA, Miller BG, Link CG

(Seroquel Study Group). 1997. Quetiapine in patients with

schizophrenia: A high- and low-dose double-blind comparison

with placebo. Arch Gen Psychiatry 54:549�/557.

Smelson DA, Losonczy MF, Davis CW, Kaune M, Williams J,

Ziedonis D. 2002. Risperidone decreases craving and relapses

in individuals with schizophrenia and cocain dependence. Can

J Psychiatry 47:671�/675.

Smith RC, Infante M, Singh A, Khandat A. 2001. The effects of

olanzapine on neurocognitive functioning in medication-refrac-

tory schizophrenia. Int J Neuropsychopharmacol 4:239�/250.

Soares BGO, Fenton M, Chue P. 2004. Sulpiride for schizo-

phrenia (Cochrane Review). In: The Cochrane Library, Issue

2. Chichester, UK: John Wiley & Sons Ltd.

Soni SD, Bamrah JS, Krska J. 1991. Effects of alcohol on serum

fluphenazine levels in stable chronic schizophrenics. Hum

Psychopharmacol 6:301�/306.

Soyka M, Albus M, Kathmann N, Finelli A, Hofstetter S,

Holzbach R, Immler B, Sand P. 1993. Prevalence of alcohol

and drug abuse in schizophrenia inpatients. Eur Arch Psychia-

try Clin Neurosci 242:362�/372.

Soyka M, Sand P. 1995. Successful treatment with flupenthixol

decanoate of a patient with both schizophrenia and alcoholism.

Pharmacopsychiatry 28:64�/65.

Soyka M. 1996. Dual diagnosis in patients with schizophrenia.

Issues in pharmacological treatment. CNS Drugs 5:414�/425.

Soyka M, Aichmuller C, von Bardeleben U, Beneke M, Glaser T,

Hornung-Knobel S, Wegner U on behalf of the study group.

2003. Flupenthixol in relapse prevention in schizophrenics with

comorbid alcoholism: Results from an open clinical study. Eur

Addict Res 9:65�/72.

Speller JC, Barnes TRE, Curson DA, Pantelis C, Alberts JL.

1997. One-year, low-dose neuroleptic study of in-patients with

chronic schizophrenia characterised by persistent negative

symptoms: Amisulpride v. haloperidol. Br J Psychiatry

171:564�/568.

Spina E, De Domenico P, Ruello C, Longobardo N, Gitto C,

Ancione M, Di Rosa AE, Caputi AP. 1994. Adjunctive

fluoxetine in the treatment of negative symptoms in chronic

schizophrenic patients. Int Clin Psychopharmacol 9(4):281�/

285.

Spina E, Avenoso A, Facciola G, Scordo MG, Ancione M, Madia

AG, Ventimiglia A, Perucca E. 2000. Relationship between

plasma concentrations of clozapine and norclozapine and

therapeutic response in patients with schizophrenia resistant

to conventional neuroleptics. Psychopharmacology 148:83�/89.

Spina E, Scordo MG, D’Arrigo C. 2003. Metabolic drug

interactions with new psychotropic agents. Fund Clin Pharma-

col 17:517�/538.

Steinert T. 2002. Prediction of inpatient violence. Acta Psychiatr

Scand (Suppl) 104:133�/141.

Still DJ, Dorson PG, Crismon ML, Pousson C. 1996. Effects of

switching inpatients with treatment-resistant schizophrenia

from clozapine to risperidone. Psychiatr Serv 47:1382�/1384.

Stone CK, Garve DL, Griffith J, Hirschowitz J, Bennett J. 1995.

Further evidence of a dose-response threshold for haloperidol

in psychosis. Am J Psychiatry 152:1210�/1212.

Sultana A, McMonagle T. 2004. Pimozide for schizophrenia or

related psychoses (Cochrane Review). In: The Cochrane

Library, Issue 2. Chichester, UK: John Wiley & Sons Ltd.

Sultana A, Reilly J, Fenton M. 2004. Thioridazine for schizo-

phrenia (Cochrane Review). In: The Cochrane Library, Issue

2. Chichester, UK: John Wiley & Sons Ltd.

Suzuki K, Awata S, Matsuoka H. 2003. Short-term effect of ECT

in middle-aged and elderly patients with intractable catatonic

schizophrenia. J ECT 19:73�/80.

Taylor CG, Flynn SW, Altman S, Ehmann T, MacEwan GW,

Honer WG. 2001. An open trial of risperidone augmentation of

partial response to clozapine. Schizophr Res 48:155�/158.

WFSBP Guidelines for Biological Treatment of Schizophrenia 189

Taylor TM, O’Toole MS, Ohlsen RI, Walters J, Pilowsky LS.

2003. Safety of quetiapine during pregnancy (letter). Am J

Psychiatry 160:588�/589.

Tenyi T, Trixler M, Kereszetes Z. 2002. Quetiapine and preg-

nancy (letter). Am J Psychiatry 159:674.

Tharyan P, Adams CE. 2004. Electroconvulsive therapy for

schizophrenia (Cochrane Review). In: The Cochrane Library,

Issue 2. Chichester, UK: John Wiley & Sons Ltd.

The Scottish First Episode Schizophrenia Study. II. 1987.

Treatment: Pimozide versus flupenthixol. The Scottish Schizo-

phrenia Research Group. Br J Psychiatry 150:334�/338.

Thornley B, Rathbone J, Adams CE, Awad G. 2004. Chlorpro-

mazine versus placebo for schizophrenia (Cochrane Review).

In: The Cochrane Library, Issue 2. Chichester, UK: John Wiley

& Sons Ltd.

Tiihonen J, Hallikainen T, Ryynanen OP, Repo-Tiihonen E,

Kotilainen I, Eronen M, Toivonen P, Wahlbeck K, Putkonen A.

2003. Lamotrigine in treatment-resistant schizophrenia: a

randomized placebo-controled crossover trial. Biol Psychiatry

54:1241�/1248.

Tollefson GD, Beasley CM Jr, Tran PV, Street JS, Krueger JA,

Tamura RN, Graffeo KA, Thieme ME. 1997. Olanzapine

versus haloperidol in the treatment of schizophrenia and

schizoaffective and schizophreniform disorders: Results of an

international collaborative trial. Am J Psychiatry 154:457�/465.

Tollefson GD, Sanger TM. 1997. Negative symptoms: A path

analytic approach to a double-blind, placebo- and haloperidol-

controlled clinical trial with olanzapine. Am J Psychiatry

154:466�/474.

Tollefson GD, Sanger TM, Lu Y, Thieme ME. 1998. Depressive

signs and symptoms in schizophrenia: A prospective blinded

trial of olanzapine and haloperidol. Arch Gen Psychiatry

55:250�/258.

Tollefson GD, Birkett MA, Kiesler GM, Wood AJ. 2001. Double-

blind comparison of olanzapine versus clozapine in schizo-

phrenic patients clinically eligible for treatment with clozapine.

Biol Psychiatry 49:52�/63.

Tondo L, Ghiani C, Albert M. 2001. Pharmacologic interventions

in suicide prevention. J Clin Psychiatry 62:51�/55.

Tran PV, Hamilton SH, Kuntz AJ, Potvin JH, Andersen SW,

Beasley C Jr, Tollefson GD. 1997. Double-blind comparison of

olanzapine versus risperidone in the treatment of schizophrenia

and other psychotic disorders. Clin Psychopharmacol

17(5):407�/418.

Tsai G, Yang P, Chung LC, Lange N, Coyle JT. 1998. D-Serine

added to antipsychotics for the treatment of schizophrenia. Biol

Psychiatry 44:1081�/1089.

Tsai GE, Yang P, Chung LC, Tsai IC, Tsai CW, Coyle JT. 1999.

D-Serine added to clozapine for the treatment of schizophre-

nia. Am J Psychiatry 156:1822�/1825.

Tsuang J, Eckman TE, Shaner A, Marder SR. 1999. Clozapine

for substance-abusing schizophrenic patients. Am J Psychiatry

156:1119�/1120 (letter).

Tsuang J, Marder SR, Han A, Hsieh W. 2002. Olanzapine

treatment for patients with schizophrenia and cocaine abuse.

J Clin Psychiatry 63:1180�/1181 (letter).

Tyson SC, Devane CL, Risch SC. 1995. Pharmacokinetic

interaction between risperidone and clozapine. Am J Psychiatry

152:1401�/1402.

Ungvari GS, Leung HC, Lee TS. 1994. Benzodiazepines and the

psychopathology of catatonia. Pharmacopsychiatry 27(6):242�/

245.

Ungvari GS, Chiu HF, Chow LY, Lau BS, Tang WK. 1999.

Lorazepam for chronic catatonia: a randomized, double-blind,

placebo-controlled cross-over study. Psychopharmacology

(Berlin) 142(4):393�/398.

Valevski A, Loebl T, Keren T, Bodinger L, Weizman A. 2001.

Response of catatonia to risperidone: Two case reports. Clin

Neuropharmacol 24(4):228�/231.

Van Bruggen J, Tijssen J, Dingemans P, Gersons B, Linszen D.

2003. Symptom response and side-effects of olanzapine and

risperidone in young adults with recent onset schizophrenia. Int

Clin Psychopharmacol 18:341�/346.

VanderZwaag C, McGee M, McEvoy JP, Freudenreich O, Wilson

WH, Cooper TB. 1996. Response of patients with treatment-

refractory schizophrenia to clozapine within three serum level

ranges. Am J Psychiatry 153:1579�/1584.

Van Putten T, Marder SR, Mintz J. 1990. A controlled dose

comparison of haloperidol in newly admitted schizophrenic

patients. Arch Gen Psychiatry 47(8):754�/758.

Van Putten T, Aravagiri M, Marder SR, Wirshing WC, Mintz J,

Chabert N. 1991. Plasma fluphenazine levels and clinical

response in newly admitted schizophrenic patients. Psycho-

pharmacol Bull 27(2):91�/96.

Vartiainen H, Tiihonen J, Putkonen A, Koponen H, Virkkunen

M, Hakola P, Lehto H. 1995. Citalopram, a selective serotonin

re-uptake inhibitor, in the treatment of aggression in schizo-

phrenia. Acta Psychiatr Scand 91:348�/351.

Velligan DI, Newcomer J, Pultz J, Csernansky J, Hoff AL,

Mahurin R, Miller AL. 2002. Does cognitive function improve

with quetiapine in comparison to haloperidol? Schizophr Res

53:239�/248.

Volavka J, Cooper TB, Czobor P, Meisner M. 1996. Effect of

varying haloperidol plasma levels on negative symptoms in

schizophrenia and schizoaffective disorder. Psychopharmacol

Bull 32:75�/79.

Volavka J, Cooper TB, Czobor P, Lindenmayer JP, Citrome LL,

Mohr P, Bark N. 2000. High-dose treatment with haloperidol:

The effect of dose reduction. J Clin Psychopharmacol 20:252�/

256.

Volavka J, Czobor P, Sheitman B, Lindenmayer JP, Citrome L,

McEvoy JP, Cooper TB, Chakos M, Lieberman JA. 2000.

Clozapine, olanzapine, risperidone, and haloperidol in the

treatment of patients with chronic schizophrenia and schizoaf-

fective disorder. Am J Psychiatry 159:255�/262.

Waehrens J, Gerlach J. 1980. Antidepressant drugs in anergic

schizophrenia. Acta Psychiatr Scand 61:438�/444.

Wagstaff A, Perry C. 2003. Clozapine: In prevention of suicide in

patients with schizophrenia or schizoaffective disorder. CNS

Drugs 17:273�/280.

Wahlbeck K, Cheine M, Essali A, Adams C. 1999. Evidence of

clozapine’s effectiveness in schizophrenia: A systematic review

and meta-analysis of randomized trials. Am J Psychiatry

156:990�/999.

Wahlbeck K, Cheine M, Tuisku K, Ahokas A, Joffe G, Rimon R.

2000. Risperidone versus clozapine in treatment-resistant

schizophrenia: A randomized pilot study. Prog Neuropsycho-

pharmacol Biol Psychiatry 24:911�/922.

Walter G, Rey JM, Mitchell PB. 1999. Practitioner review:

Electroconvulsive therapy in adolescents. J Child Psychol

Psychiatry 40(3):325�/334.

Waraich PS, Adams CE, Roque M, Hamill KM, Marti J. 2004.

Haloperidol dose for the acute phase of schizophrenia (Co-

chrane Review). In: The Cochrane Library, Issue 2. Chiche-

ster, UK: John Wiley & Sons Ltd.

Warner B, Alphs L, Schaedelin J, Koestler T. 2000. Clozapine and

sudden death (letter). Lancet 355:842.

Wassef A, Dott SG, Harris A, Brown A, O’Boyle M, Meyer WJ

3rd, Rose RM. 2000. Randomized, placebo-controlled pilot

study of divalproex sodium in the treatment of acute exacer-

bations of chronic schizophrenia. J Clin Psychopharmacol

20:357�/361.

190 P. Falkai et al.

Wassef A, Baker J, Kochan LD. 2003b. GABA and schizophrenia:

A review of basis science and clinical studies. J Clin Psycho-

pharmacol 23:601�/640.

Weiden PJ, Simpson GM, Potkin G, O’Sullivan RL. 2003a.

Effectiveness of switching to ziprasidone for stable but sympto-

matic outpatients with schizophrenia. J Clin Psychiatry

64:580�/588.

Weisman RL. 2003. Quetiapine in the successful treatment of

schizophrenia with comorbid alcohol and drug dependence: A

case report. Int J Psychiatry Med 33(1):85�/89.

Weiss EM, Bilder RM, Fleischhacker WW. 2002. The effects of

second-generation antipsychotics on cognitive functioning and

psychosocial outcome in schizophrenia. Psychopharmacology

(Berlin) 162(1):11�/17.

Wetzel H, von Bardeleben U, Holsboer F, Benkert O. 1991.

Zotepine versus perazine in patients with paranoid schizophre-

nia: A double-blind controlled trial of its effectiveness].

Fortschr Neurol Psychiat 59(Suppl 1):23�/29.

Wetzel H, Grunder G, Hillert A, Philipp M, Gattaz WF, Sauer H,

Adler G, Schroder J, Rein W, Benkert O and the Amisulpride

Study Group. 1998. Amisulpride versus flupentixol in schizo-

phrenia with predominantly positive symptomatology �/ dou-

ble-blind study comparing a selective D2-like antagonist to a

mixed D1-/D2-like antagonist. Psychopharmacology 137:223�/

232.

Whitehead C, Moss S, Cardno A, Lewis G. 2004. Antidepressants

for people with both schizophrenia and depression (Cochrane

Review). In: The Cochrane Library, Issue 2. Chichester, UK:

John Wiley & Sons Ltd.

Wiedemann G, Klingberg S. 2003. Psychotherapie produktiver

Symptomatik bei Patienten mit schizophrener Psychose. Ner-

venarzt 74:76�/84.

Wiesbeck GA, Weijers HG, Lesch OM, Glaser T, Toennes PJ,

Boening J. 2001. Flupenthixol decanoate and relapse preven-

tion in alcoholics: Results from a placebo-controlled study.

Alcohol Alcohol 36(4):329�/334.

Wilkins JN. 1997. Pharmacotherapy of schizophrenia patients

with comorbid substance abuse. Schizophr Bull 23:215�/228.

Wirshing DA, Marshall BD Jr, Green MF, Mintz J, Marder SR,

Wirshing WC. 1999. Risperidone in treatment-refractory

schizophrenia. Am J Psychiatry 156:1374�/1379.

Wirshing DA, Boyd JA, Meng LR, Ballon JS, Marder SR,

Wirshing WC. 2002. The effects of novel antipsychotics on

glucose and lipid levels. J Clin Psychiatry 63(10):856�/865.

Wobrock T, Falkai P, Pajonk FG. 2004. Acute treatment of

schizophrenia]. Fortschr Neurol Psychiatr 72:705�/726.

Wolkowitz OM, Pickar D. 1991. Benzodiazepines in the treatment

of schizophrenia: A review and reappraisal. Am J Psychiatry

148:714�/726.

Wolkowitz OM, Turetsky N, Reus VI, Hargreaves WA. 1992.

Benzodiazepine augmentation of neuroleptics in treatment-

resistant schizophrenia. Psychopharmacol Bull 28:291�/295.

Woods SW, Stolar M, Sernyak MJ, Charney DS. 2001. Consis-

tency of atypical antipsychotic superiority to placebo in recent

clinical trials. Biol Psychiatry 49(1):64�/70.

World Health Organization. 2000. WHO Guide to Mental Health

in Primary Care, London (www.royscomed.ac.uk).

Wright P, Birkett M, David SR, Meehan K, Ferchland I, Alaka

KJ, Saunders JC, Krueger J, Bradley P, San L, Bernardo M,

Reinstein M, Breier A. 2001. Double-blind, placebo-controlled

comparison of intramuscular olanzapine and intramuscular

haloperidol in the treatment of acute agitation in schizophrenia.

Am J Psychiatry 158(7):1149�/1151.

Yap HL, Mahendran R, Lim D, Liow PH, Lee A, Phang S, Tiong

A. 2001. Risperidone in the treatment of first episode psycho-

sis. Singapore Med J 42(4):170�/173.

Yorkston NJ, Gruzelier JH, Zaki SA, Hollander D, Pitcher DR,

Sergeant HG. 1977. Propranolol as an adjunct to the treatment

of schizophrenia. Lancet ii:575�/578.

Yovell Y, Opler LA. 1994. Clozapine reverses cocaine craving in a

treatment-resistant mentally ill chemical abuser: A case report

and a hypothesis. J Nerv Ment Dis 182:591�/592 (letter).

Zemlan FP, Hirschowitz J, Sautter F, Garver DL. 1986. Relation-

ship of psychotic symptom clusters in schizophrenia to neuro-

leptic treatment and growth hormone response to

apomorphine. Psychiatry Res 18(3):239�/255.

Zhang XY, Zhou DF, Cao LY, Zhang PY, Wu GY, Shen YC.

2001. Risperidone versus haloperidol in the treatment of acute

exacerbations of chronic inpatients with schizophrenia: A

randomized double-blind study. Int Clin Psychopharmacol

16:325�/330.

Ziedonis DM, Richardson T, Lee E, Petrakis I, Kosten TR. 1992.

Adjunctive desipramine in the treatment of cocaine abusing

schizophrenics. Psychopharmacol Bull 28:309�/314.

Zimmet S, Strous RD, Burgess E, Kohnstamm S, Green AI.

2000. Effects of clozapine on substance use in patients with

schizophrenia and schizoaffective disorder. J Clin Psychophar-

macol 20:94�/98.

Zoccali R, Muscatello MR, Cedro C, Neri P, La Torre D, Spina E,

Di Rosa AE, Meduri M. 2004. The effect of mirtazapine

augmentation of clozapine in the treatment of negative symp-

toms of schizophrenia: A double-blind, placebo-controlled

study. Int Clin Psychopharmacol 19:71�/76.

WFSBP Guidelines for Biological Treatment of Schizophrenia 191

REVIEW

Cognitive neuropsychiatry: Conceptual, methodological andphilosophical perspectives

JAKOB HOHWY1 & RABEN ROSENBERG2

1Department of Philosophy, Aarhus University, Aarhus, Denmark, and 2Centre for Basic Psychiatric Research, University

Psychiatric Hospital in Aarhus, Risskov, Denmark

AbstractCognitive neuropsychiatry attempts to understand psychiatric disorders as disturbances to the normal function of humancognitive organisation, and it attempts to link this functional framework to relevant brain structures and their pathology.This recent scientific discipline is the natural extension of cognitive neuroscience into the domain of psychiatry. We presenttwo examples of recent research in cognitive neuropsychiatry: delusions of control in schizophrenia, and affective disorders.The examples demonstrate how the cognitive approach is a fruitful and necessary supplement to the otherwise successfulbiological psychiatry paradigm, which tend to bypass the cognitive level. Philosophy concerns some of the core conceptsinvolved in psychiatric illness, particularly concerning rationality, thought and action, reality testing, and the self. Wepresent concrete examples that illustrate how philosophical conceptual tools can be particularly important for theconstruction and interpretation of the cognitive models relevant to the understanding of psychiatric illness. We concludethat cognitive neuropsychiatry is a fruitful and necessary supplement to biological psychiatry. Furthermore, cognitiveneuropsychiatry itself may benefit significantly from employing philosophical conceptual tools in the interpretation andconstruction of its cognitive models. The cognitive and philosophical approaches may thus be further steps towards ascientific psychopathology.

Key words: Neuropsychiatry, cognitive neuropsychiatry, philosophy, delusions of control, affective disorder

Introduction

Cognitive neuropsychiatry is a new branch of

psychiatric research. It brings together cognitive

and neurobiological approaches to the study of

mental disorders. It promises to advance our under-

standing of psychiatric illness by finding the right

theoretical and methodological balance between

otherwise successful paradigms, such as biological

psychiatry and cognitive psychology.

We review conceptual and scientific aspects of

cognitive neuropsychiatry and present examples of

research in this field. The examples demonstrate that

the cognitive approach is a useful and necessary

supplement to neuropsychiatry.

Philosophy may be especially important for the

study and understanding of psychiatric illnesses that

often touch on philosophical notions concerning

rationality, thought and action, reality testing, and

the self. We present some concrete examples that

illustrate the relevance of philosophical analysis to

psychiatric research.

We conclude that there is reason to believe that

cognitive neuropsychiatry will be a fruitful and

necessary supplement to biological psychiatry, and

that philosophical analysis will be a useful aid to

cognitive neuropsychiatry. Both the cognitive and

the philosophical approaches may therefore consti-

tute further steps towards a scientific psychopathol-

ogy.

Neuropsychiatry and cognitive science

Neuropsychiatry is claimed to be a discipline for the

future (Sachdev 2002) and several recent papers

have focused on the integration of psychiatry,

neurology and neuroscience (Cowan et al. 2000;

Martin 2002; Yudofsky et al. 2002), thereby broad-

ening the concept of biological psychiatry with

its primary focus on neurobiology. To influential

Correspondence: Raben Rosenberg, Centre for Basic Psychiatric Research, University Psychiatric Hospital in Aarhus, Skovagervej 2,

DK-8240 Risskov, Denmark. Tel: �/45 7789 3511. Fax: �/45 7789 3549. E-mail: [email protected]

The World Journal of Biological Psychiatry, 2005; 6(3): 192�/197

(Received 28 April 2003; accepted 21 October 2004)

ISSN 1562-2975 print/ISSN 1814-1412 online # 2005 Taylor & Francis

DOI: 10.1080/15622970510029867

scientists, such as Kandel, modern neuroscience

theory and methodology represent a new framework

for psychiatry (Kandel 1998) that hopefully will lead

to better understanding and treatment of psychiatric

disorders, based on dysfunction of specific genes,

molecules, neuronal organelles and specific neuronal

systems.

Cognitive science is a discipline that attempts to

understand the mind in terms of information pro-

cessing, that is, representational structures in the

brain and computational procedures that operate

on them. With the development over the last 30

years of sophisticated neuroimaging and neurophy-

siological methods for studying the living brain,

neuroscience has become a core component of

cognitive science. Now the dominating science of

the mind is the amalgamation of neuroscience and

cognitive science in cognitive neuroscience �/ optimis-

tically called the ‘biology of the mind’ (Gazzaniga

et al. 1998).

Cognitive neuropsychiatry

Cognitive neuropsychiatry (CNP, hereafter) is the

extension of the methods of cognitive neuroscience

and cognitive neuropsychology to the study of

psychiatric disorders. It was advanced as a scientific

discipline in its own right in the beginning of the

1990s, and got its own journal (Cognitive Neuropsy-

chiatry) in 1996. Instrumental in promoting CNP

were Hadyn Ellis, Anthony David, Peter Halligan

and John Marshall (see e.g. David 1993). The state

of the art in terms of scientific results has been

elegantly reviewed (Halligan et al. 2001).

CNP attempts to understand psychiatric disorders

as disturbances to the normal function of human

cognitive organisation and it attempts to link this

functional framework to relevant brain structures

and their pathology. The focus is on symptoms

rather than syndromes or diagnoses because phe-

nomenologically based diagnoses rarely lend them-

selves to a differentiated functional analysis or

correlation with specific brain structures. Thus,

insight into normal psychological processes is essen-

tial for a scientific analysis of psychiatric symptoma-

tology. CNP has a somewhat broader approach to

the analysis of psychiatric disorders than neuropsy-

chiatry because it focuses on a cognitive and often

more philosophical analysis of human action and

behaviour (see below). CNP and philosophical

analysis could therefore be a fruitful and necessary

supplement to neuropsychiatry, something that al-

lows us to better understand how and why biological

malfunctions and biological damage give rise to

mental disorders.

Examples of research in cognitive

neuropsychiatry

Current research in CNP includes studies of the

various positive symptoms found in schizophrenia

such as auditory hallucinations, persecutory delu-

sions and passivity experiences, delusions of mis-

identification; as well as autism, neuropsychiatric

symptoms and signs associated with motor control,

and emotional and affective disorders. We now

briefly review two sets of studies in order to

demonstrate how taking a cognitive approach en-

ables us to better understand how psychiatric

disorders arise from malfunctions and damage to

brain activity.

Delusions of control

Delusions of control is a first-rank Schneiderian

symptom of schizophrenia. Patients believe that

actions they perform are really performed by other

agents. It is classed with other passivity experiences,

such as thought withdrawal and thought insertion. It

is possible to begin to explain this delusion in terms

of the general functioning of our sensorimotor

system and our awareness of it (Frith et al. 2000).

The account begins with cognition, or informa-

tion-processing at the most basic level: when we

issue a motor command, an efference copy of it is

also issued. This copy serves to predict how the state

of the system will change given the performance of

the movement, as well as what the sensory feedback

will be given the new state (Wolpert et al. 2000). The

first prediction is used in comparisons with the

desired state and allows rapid error correction,

before re-afferent feedback is available (Blakemore

et al. 2002). The second prediction probably serves

to attenuate activity in the inferior parietal cortex

associated with the sensory consequences of our own

movement (Blackwood et al. 2000; Blakemore et al.

1998). It may be that the feeling of being in control

of our movements is associated with activity in this

area: when there is activity it is not our movement,

when activity is attenuated via the prediction, we can

tell it is our movement (Frith et al. 2000).

If there is no awareness of the predicted limb

position, then the subject will be aware of the goal, of

the intention to move and of the movement having

occurred, but not of the exact specification of the

movement. The subject may thus be unaware of

having initiated that movement.

Delusions of control may therefore arise when

there is an abnormality in the production of forward

models of movement such that the patient feels that,

though he or she intended the movement, it was

initiated by some external force. This account

predicts that patients should be relatively poor at

Cognitive neuropsychiatry 193

rapid error control, in the absence of visual feedback,

and this has been experimentally confirmed (Frith et

al. 1988). Spence et al. (1997) scanned delusional

schizophrenic subjects while they performed joystick

movements. During the scan they experienced delu-

sions of control. As compared to normal controls

and other groups of schizophrenic subjects, this

group showed overactivity in the inferior parietal

cortex, consistent with the claim that activity in this

area is normally attenuated in expectation of the

sensory consequences of one’s own movements.

This overactivity was associated with relative under-

activity in the dorsolateral prefrontal cortex that may

be an area involved in modulating activity in the

inferior parietal cortex.

In a recent study (Blakemore et al. 2003), delu-

sions of control were induced in normal subjects.

Hypnotised subjects were scanned in three condi-

tions: while actively raising their arm, while passively

having their arm raised, and while actively raising

their arm but under the hypnotic suggestion that it

was being passively raised. In the last condition they

believed they were not in control, even though they

did actually raise their arm, and this was associated

with increased activity in the inferior parietal cortex.

This empirically tested cognitive approach to

delusions of control yields a conceptual framework

for beginning to understand what the observed brain

activity means, and how it may give rise to delusions

of control rather than other disorders. Without this

cognitive account, the brain activity (e.g., reduction

of activity in inferior parietal cortex) in itself would

be difficult to make sense of.

Affective disorder

A major task for psychiatric nosology has been to

classify and to define subgroups of affective dis-

orders. As evidenced by current classification sys-

tems, psychopathological description of symptoms

evaluated cross-sectionally or during lifetime course

have been the most influential approaches. Although

primarily defined by mood changes, affective dis-

orders are characterized by significant impairment of

cognitive functions. Numerous neuropsychological

studies of unipolar (Ravnkilde et al. 2002) and

bipolar disorders (Chowdhury et al. 2003) have

revealed abnormal functioning of a variety of cogni-

tive functions, including attention, memory, psycho-

motor speed and executive functions, indicating that

cognitive deficits are not just epiphenomena but

essential aspects of the disorders. Neuroimaging

techniques are extensively used to characterize the

brain mechanisms underlying the clinically impor-

tant cognitive dysfunctions in depression (Videbech

1997; Videbech 2000). The classical symptoms of

guilt feeling, self-reproaches, lack of initiative, and

concentration difficulties have been linked to

changes in normal neuronal activation patterns

(Davidson et al. 2002). Important pathways include

the orbitofrontal cortex, dorsolateral prefrontal

cortex, the anterior cingulate, hippocampus and

amygdala (Davidson et al. 2002; Videbech et al.

2001), and dysfunction in different parts of this

circuitry are indicated by several studies.

Furthermore, precipitation of depressive episodes

by life events are interpreted in terms of stress

related to cognitive coping strategies and neuroen-

docrinological reactions (Gould et al. 1999; Nestler

et al. 2002). Finally long term deviations of cognitive

functions in depressive patients have been correlated

to atrophy of hippocampus (Lee et al. 2002;

MacQueen et al. 2003).

The clinical efficacy of antidepressant drugs is

suggested to involve neurogenesis in hippocampus

(Kempermann 2002) and by changing serotonergic

and noradrenergic neurotransmission, antidepres-

sant drugs may influence stress responses, cognitive

functions and mood. Cognitive therapy is efficacious

in treating mild to moderate depressive episodes

(Scott 2001) and increasing evidence has been

obtained for the importance of the treatment mod-

ality in relapse prevention (Jarrett et al. 2001), also

on the potential benefits of combining cognitive

therapy with medication (Arean and Cook 2002).

In a recent PET study, cognitive therapy as well as

paroxetine were suggested to induce clinical recovery

by modulating the functioning of specific sites in

limbic and cortical regions (Goldapple et al. 2004).

In conclusion, analysis of cognitive functioning in

affective disorders has transpired as a fruitful

approach that might lead to a comprehensive theory

integrating clinical phenomenology, neuropsychol-

ogy, neuroimaging, neuropathology and neurophar-

macology.

Philosophical perspectives on cognitive

neuropsychiatry

Psychiatric disorders hold particular interest for

philosophy because they concern core philosophical

concepts such as belief, perception, experience,

reality, self, other, emotion, and consciousness.

Specifically, the construction and interpretation of

cognitive models in CNP may benefit from philoso-

phical analyses of the involved concepts*/at least

to the extent these analyses are apt for application

in ongoing empirical research. We now present

some concrete examples of philosophical analysis

of psychiatric disorder. When there are relatively

well-developed cognitive models, as in the case of

delusions of control, philosophical analysis may be

194 J. Hohwy & R. Rosenberg

needed to address outstanding questions. And

philosophical analysis will also be needed in cases,

such as affective disorder, where the cognitive

approach is in its beginning phases.

Delusions and rationality

There may often be an experiential factor in the

formation of delusional beliefs (for an overview, see

Davies et al. 2001). For example, in the case of

delusions of control there may be the unusual and

distressing experience that one’s intended bodily

movements are initiated by another agent. Some

cognitive process involving reasoning must then be

active in the production of the irrational delusional

belief that the movement is in fact initiated by

someone else. It is an outstanding question how

this process goes (Davies et al. 2000). We need to

explain why the alien agent hypothesis presents itself

(rather than the more probable hypothesis that

something has gone wrong in the patient’s brain);

why the subject begins to believe the hypothesis; and

why this belief is tenaciously maintained in spite of

what everyone else says.

Clearly, our explanations in this area depend to a

large degree on which concept of rationality we

chose to operate with, and on our understanding of

the normal processes for belief revision. Philoso-

phers operate with notions of ideal and realistic

rationality, that is, the rules of rationality that people

should (but in general cannot) satisfy to be com-

pletely rational, versus the rules of rationality that

people actually employ, partly as a consequence of

the fact that they are not ideally rational (Harman

1986).

Our conceptualisation of pathological belief for-

mation will be influenced by our conception of

rationality. Do delusional beliefs arise as rational or

as irrational responses to the supposed unusual

experiences? Is the experience of such a character

that the alien agent hypothesis is in fact rational,

or is it necessary to posit a further deficit that

results in a preference for irrational, bizarre hypoth-

eses? If some delusions arise as rational responses to

unusual experiences, then the prediction is that

healthy subjects that have such experiences will

develop delusional states. The mentioned study by

Blakemore et al. (2003) may bear out this prediction

(at least if the subjects’ reported states of uninten-

tionality are comparable to delusions of control,

and unless hypnosis induces the requisite kind of

irrationality).

It may also be necessary to distinguish more

departments of rationality than the merely rule-

based or procedural. Sometimes, it is the content of

a propositional attitude like a belief or a desire that

appears irrational and pathological, never mind the

procedural rules (whether ideal or realistic) that

were used to adopt it (Lewis 1986); consider for

example the bare desire for a saucer of mud

(Anscombe 1957). Further conditions on rationality

may also be necessary, for example, disturbances to

the sense of egocentricity or self (that I am the agent

or thinker behind these actions and thoughts) may

play a role in the production of irrational belief and

the ability to reason coherently (as in delusions of

control and thought insertion) (see Frith (1992), for

discussion; see Gold et al. (2000)).

These philosophical analyses suggest the possibility

of a range of cognitive models to account for

delusion formation. Delusions of various kinds may

arise as the results of disturbances to cognitive

systems concerning (a) content formation, and/or

(b) procedural rules of belief formation, and/or (c)

other cognitive systems, such as those involved in

egocentricity (see also Gerrans 2001).

Reality testing

It seems plausible that reality testing is part of what

makes a knowledge gathering mechanism reliable.

Reality testing is required when initial beliefs are

formed in sub-optimal conditions (bad lighting,

background noise, etc), or when they conflict with

strongly held prior beliefs. When it comes to sensory

modalities such as vision, touch and hearing, we

have a range of procedures for reality testing that can

override poor initial belief formation. We can turn on

the light, feel with our hands, move closer to the

source of sound, and so on. If these ordinary

procedures for reality testing converge on the same

result, then we come to believe the content in

question, and the beliefs would be normally be

justified.

This commonsense epistemological notion of reality

testing may connect with the hypothesis that some

delusions arise as responses to unusual experiences,

caused by biological deficit or damage. These

experiences may occur in sensory modalities where

we do not have such ready procedures for reality

testing, for example, experiences of sensorimotor

feedback (also of speech production), including

awareness of the output of discrepancy monitors as

mentioned in the above account of delusions of

control; awareness of introspection of thought,

and awareness of emotional and affective states. If

there are no or very limited strategies for reality

testing, then it may be impossible to override and

revise the beliefs formed on the basis of those

unusual experiences. Every time the subject tries to

reality test, the same result obtains, and there is no

alternative route to reality test experiences in these

Cognitive neuropsychiatry 195

modalities. This would go some way towards ex-

plaining the tenacity with which patients hold on

to delusional beliefs that incorporate an unusual

experience. On this view, there is no specific reason-

ing deficit or bias involved in delusion formation. If

it was possible to give people unusual experiences

in these sensory modalities, then the hypothesis

would be supported by finding that normal subjects

with no reasoning deficits or biases form delusional

beliefs (Blakemore et al. 2003) may be an example;

see Hohwy and Rosenberg (2005) for more detailed

discussion).

Decision, reasoning and affect

The exploration of the cognitive aspects of affective

disorders (see above) is still in its beginning phases,

and intriguing philosophical questions arise. It is

plausible that affective and emotional states play a

role in decision processes (Bechara et al. 1994). This

suggests that affective disorders are related to

realistic, rather than ideal rationality, as discussed

above. It may be that defective processing of

affective and emotional states play a role in produ-

cing the cognitive deficits seen in affective disorders.

On the other hand, it may also be that the cognitive

deficits cause failures in the way affective states

are brought to bear on decision making and belief

formation. Deciding on these issues again partly

depend on how we analyse and organise philosophi-

cal notions of rationality, decision, belief and emo-

tion.

A scientific psychopathology?

The move away from Freudian psychoanalysis to a

biologically based psychiatry in the last century has

made psychiatry a more scientific enterprise, even

though most new drugs have appeared due to

serendipity and not as a consequence of insight

into disease processes. Thus, a majority of current

theories of mental disorders have been developed

secondary to the pharmacodynamics of drugs, e.g.,

theories of neurotransmitter deficiencies for schizo-

phrenia and depression. The cognitive approach

taken by cognitive neuropsychiatry may help gain a

more direct focus on psychiatric disorders because it

goes from symptoms to brain activity via cognitive

models of normal psychological functioning.

Furthermore, philosophical analysis may be a fruit-

ful supplement to cognitive neuropsychiatry. Cogni-

tive neuropsychiatry and philosophy may thus

constitute further steps towards a scientific psycho-

pathology.

Statement of interest

The authors have no conflict of interest with any

commercial or other associations in connection with

the submitted article.

References

Anscombe GEM. 1957. Intention. Oxford: Blackwells.

Arean PA, Cook BL. 2002. Psychotherapy and combined

psychotherapy/pharmacotherapy for late life depression. Biol

Psychiatry 52:293�/303.

Austin DW, Richards JC. 2001. The catastrophic misinterpreta-

tion model of panic disorder. Behav Res Ther 39:1277�/1291.

Bandelow B, Zohar J, Hollander E, Kasper S, Moller HJ. 2002.

World Federation of Societies of Biological Psychiatry

(WFSBP) guidelines for the pharmacological treatment of

anxiety, obsessive-compulsive and posttraumatic stress disor-

ders. World J Biol Psychiatry 3:171�/199.

Bechara A, Damasio AR, Damasio H, Anderson SW. 1994.

Insensitivity to future consequences following damage to

human prefrontal cortex. Cognition 50:7�/15.

Blackwood NJ, Howard RJ, ffytche DH, Simmons A, Bentall RP,

Murray RM. 2000. Imaging attentional and attributional bias:

An fMRI approach to the paranoid delusion. Psychol Med

30:873�/883.

Blackwood NJ, Howard RJ, Bentall RP, Murray RM. 2001.

Cognitive models of persecutory delusions. Am J Psychiatry

158:527�/539.

Blakemore SJ, Wolpert DM, Frith CD. 1998. Central cancellation

of self-produced tickle sensation. Nat Neurosci 1:635�/640.

Blakemore SJ, Wolpert DM, Frith CD. 2002. Abnormalities in the

awareness of action. Trends Cogn Sci 6:237�/242.

Blakemore SJ, Oakley DA, Frith CD. 2003. Delusions of alien

control in the normal brain. Neuropsychologia 41:1058�/1067.

Chowdhury R, Ferrier IN, Thompson JM. 2003. Cognitive

dysfunction in bipolar disorder. Curr Opin Psychiatry 16:7�/12.

Churchland PS, Sejnowski TJ. 1999. Perspectives on cognitive

neuroscience. In: Gazzaniga MS, editor. Cognitive neuro-

science: A reader. Malden, MA: Blackwell.

Clark D. 1986. A cognitive approach to panic. Behav Res Ther

24:461�/470.

Cowan WM, Harter DH, Kandel ER. 2000. The emergence of

modern neuroscience: some implications for neurology and

psychiatry. Annu Rev Neurosci 23:343�/391.

David AS. 1993. Cognitive neuropsychiatry? Psychol Med 23:1�/

5.

Davidson RJ, Pizzagalli D, Nitschke JB, Putnam K. 2002.

Depression: Perspectives from affective neuroscience. Annu

Rev Psychol 53:545�/574.

Davies M, Coltheart M. 2000. Introduction: Pathologies of belief.

Mind Lang 15:1�/46.

Davies M, Coltheart M, Langdon R, Breen N. 2001. Monothe-

matic delusions: Toward a two-factor account. Phil Psychiatr

Psychol 8:133�/158.

Elliott R. 1998. The neuropsychological profile in unipolar

depression. Trends Cogn Sci 11:447�/454.

Frith CD. 1992. The cognitive neuropsychology of schizophrenia.

Hillsdale, NJ: Lawrence Erlbaum Ass.

Frith CD, Done DJ. 1988. Towards a neuropsychology of

schizophrenia. Br J Psychiatry 153:437�/443.

Frith CD, Blakemore SJ, Wolpert DM. 2000. Abnormalities in

the awareness and control of action. Phil Trans R Soc London

B Biol Sci 355:1771�/1788.

Furmark T, Tillfors M, Marteinsdottir I, Fischer H, Pissiota A,

Langstrom B, Fredrikson M. 2002. Common changes in

cerebral blood flow in patients with social phobia treated with

196 J. Hohwy & R. Rosenberg

citalopram or cognitive-behavioral therapy. Arch Gen Psychia-

try 59:425�/433.

Gazzaniga M, Ivry RB, Mangun GR. 1998. Cognitive neuro-

science: The biology of the mind. New York: WW Norton.

Gerrans P. 2001. Delusions as performances failures. Cogn

Neuropsychiatry 3:161�/173.

Gold I, Hohwy J. 2000. Rationality and schizophrenic delusion.

Mind Lang 15:146�/167.

Goldapple K, Segal Z, Garson C, Lau M, Bieling P, Kennedy S, et

al. 2004. Modulation of cortical-limbic pathways in major

depression: Treatment-specific effects of cognitive behavior

therapy. Arch Gen Psychiatry 61:34�/41.

Gorman JM, Kent JM, Sullivan GM, Coplan JD. 2000. Neuroa-

natomical hypothesis of panic disorder, revised. Am J Psychia-

try 157:493�/505.

Gould E, Tanapat P. 1999. Stress and hippocampal neurogenesis.

Biol Psychiatry 46:1472�/1479.

Green RL, Gazzaniga MS. 2001. Linking neuroimaging and

cognitive neuropsychiatry. Cogn Neuropsychiatry 6:229�/234.

Halligan PW, David AS. 2001. Cognitive neuropsychiatry:

Towards a scientific psychopathology. Nat Rev Neurosci

2:209�/215.

Harman G. 1986. Change in view. Cambridge, MA: MIT Press.

Hohwy J, Rosenberg, R. 2005. Reality testing, unusual experi-

ences and delusions of control. Mind Lang (in press).

Jarrett RB, Kraft D, Doyle J, Foster BM, Eaves GG, Silver PC.

2001. Preventing recurrent depression using cognitive therapy

with and without a continuation phase: A randomized clinical

trial. Arch Gen Psychiatry 58:381�/388.

Kandel ER. 1998. A new intellectual framework for psychiatry.

Am J Psychiatry 155:457�/469.

Kempermann G. 2002. Regulation of adult hippocampal neuro-

genesis: Implications for novel theories of major depression.

Bipolar Disord 4:17�/33.

Lee AL, Ogle WO, Sapolsky RM. 2002. Stress and depression:

possible links to neuron death in the hippocampus. Bipolar

Disord 4:117�/128.

Lewis D. 1986. On the plurality of worlds. Oxford: Oxford

University Press.

Libet B, Gleason CA, Wright EW, Pearl DK. 1983. Time of

conscious intention to act in relation to onset of cerebral

activity (readiness-potential). The unconscious initiation of a

freely voluntary act. Brain 106:623�/642.

MacQueen GM, Campbell S, McEwen BS, MacDonald K,

Amano S, Joffe RT, Nahmias C, Young LT. 2003. Course of

illness, hippocampal function, and hippocampal volume in

major depression. Proc Natl Acad Sci USA 100:1387�/1392.

Martin JB. 2002. The integration of neurology, psychiatry, and

neuroscience in the 21st century. Am J Psychiatry 159:695�/

704.

Nestler EJ, Barrot M, Eisch AJ, Gold SJ, Monteggia LM. 2002.

Neurobiology of depression. Neuron 34:13�/25.

Ravnkilde B, Videbech P, Clemmensen K, Egander A, Rasmussen

NA, Rosenberg R. 2002. Cognitive deficits in major depres-

sion. Scand J Psychol 43:239�/251.

Sachdev PS. 2002. Neuropsychiatry �/ A discipline for the future.

J Psychosom Res 53:625�/627.

Scott J. 2001. Cognitive therapy for depression. Br Med Bull

57:101�/113.

Spence SA, Brooks DJ, Hirsch SR, Liddle PF, Meehan J, Grasby

PM. 1997. A PET study of voluntary movement in schizo-

phrenic patients experiencing passivity phenomena (delusions

of alien control). Brain 120:1997�/2011.

Spence SA, Frith CD. 1999. Towards a functional anatomy of

volition. J Consc Stud 6:11�/29.

van Balkom AJLM, van Oppen P, Vermeulen AWA, Van Dyck R,

Nauta MCE, Vorst HCM. 1994. A meta-analysis on the

treatment of obsessive compulsive disorder: A comparison of

antidepressants, behavior, and cognitive therapy. Clin Psych

Rev 14:359�/381.

van Balkom AJ, Bakker A, Spinhoven P, Blaauw BM, Smeenk S,

Ruesink B. 1997. A meta-analysis of the treatment of panic

disorder with or without agoraphobia: A comparison of

psychopharmacological, cognitive- behavioral, and combina-

tion treatments. J Nerv Ment Dis 185:510�/516.

Videbech P. 1997. MRI findings in patients with affective

disorder. A meta-analysis. Acta Psychiatr Scand 96:157�/168.

Videbech P. 2000. PET measurements of brain glucose metabo-

lism and blood flow in major depressive disorder: A critical

review. Acta Psychiatr Scand 101:11�/20.

Videbech P, Ravnkilde B, Pedersen AR, Egander A, Landbo B,

Rasmussen NA, Andersen F, Stodkilde-Jorgensen H, Gjedde

A, Rosenberg R. 2001. The Danish PET/depression project:

PET findings in patients with major depression. Psychol Med

31:1147�/1158.

Wolpert DM, Ghahramani Z. 2000. Computational principles of

movement neuroscience. Nat Neurosci 3(Suppl):1212�/1217.

Yudofsky SC, Hales RE. 2002. Neuropsychiatry and the future of

psychiatry and neurology. Am J Psychiatry 159:1261�/1264.

Cognitive neuropsychiatry 197

ORIGINAL INVESTIGATION

Agitated depression in bipolar II disorder

FRANCO BENAZZI1,2,3

1E. Hecker Outpatient Psychiatry Center, Ravenna, Italy, 2University of California at San Diego, San Diego, CA, USA, and3Department of Psychiatry, National Health Service, Forli, Italy

AbstractStudy aim : To test diagnostic validity or utility of agitated depression (AD) in bipolar II disorder (BP-II).Methods : Three hundred and twenty BP-II major depressive episode (MDE) outpatients interviewed with the StructuredClinical Interview for DSM-IV, Hypomania Interview Guide (HIG), and Family History Screen. AD defined as MDE withpsychomotor agitation. Mixed depression defined as MDE with ]/4 hypomanic symptoms. AD, non-AD, mixed-AD, non-mixed-AD, and mixed-non-AD were compared versus diagnostic validators.Results : AD was present in 35.0%, 75.8% of AD were mixed, while only 14.3% of non-AD were mixed (P�/0.0000). AD(n�/112), versus non-AD (n�/208), had significantly higher age, more females, recurrences, bipolar I family history, andmuch more concurrent hypomanic symptoms. Mixed-AD (n�/85), versus non-mixed-AD (n�/27), was not significantlydifferent, apart from more hypomanic symptoms (by definition), but there were clinically significant differences.Conclusions : Findings may partly support subtyping of AD in BP-II, on the basis of its frequent clustering of hypomanicsymptoms, and its different family history. This subtyping may impact on treatment of BP-II depression, as antidepressantsalone may increase agitation while mood stabilising agents can treat agitation before using antidepressants.

Key words: Agitated depression, bipolar II disorder, depressive mixed state, agitation

Introduction

The diagnostic validity or utility of agitated depres-

sion (AD, a major depressive episode (MDE) with

psychomotor agitation) in the classification of mood

disorders is unclear. In DSM-IV-TR (American

Psychiatric Association 2000) and in ICD-10

(WHO 1992) MDE with psychomotor agitation

can be present in bipolar and unipolar major

depressive disorders, without any subtyping (speci-

fier). Only DSM-IV-TR melancholic features speci-

fier, and ICD-10 ‘somatic syndrome’ of depression,

give more importance to psychomotor change (agi-

tation or retardation). Parker (2000) views psycho-

motor change as the core feature of melancholic

depression, a view shared also by DSM-IV-TR (in

the text). However, no distinction is made between

agitated and retarded depression in both diagnostic

systems.

Kraepelin (1913, 1921, English translation by

Barclay) and Weygandt (1899, English translation

by Marneros 2001) described ‘excited depression’ as

a depressive mixed state (i.e., opposite symptoms of

excitement and inhibition concurrently present in

the same mood episode). It was described as a

depression with the excitement symptoms restless-

ness (psychomotor agitation), increased talkative-

ness, and irritability.

The diagnostic validity of AD has been supported

by some studies and questioned by other studies (for

a review, see Koukopoulos and Koukopoulos 1999;

Swann et al. 1993, 1994; Schatzberg and DeBattista

1999; Parker 2000; Benazzi et al. 2002; Benazzi et

al. 2004). AD status remains uncertain. Recently,

Maj et al. (2003) found, in bipolar-I disorder, that

AD was a more severe variant of bipolar I depres-

sion, its most common manic symptoms being, apart

from psychomotor agitation, irritability, more talka-

tiveness, distractibility, and racing thoughts.

Goodwin and Ghaemi (2003) suggested that the

diagnostic validity of AD would be supported by

comparing diagnostic validators among the following

subtypes: (1) AD plus other concurrent manic

symptoms (mixed-AD), (2) AD without other manic

symptoms (non-mixed-AD), and (3) depression

without psychomotor agitation (non-AD). As anti-

Correspondence: Dr Franco Benazzi, Via Pozzetto 17, 48010 Castiglione di Cervia RA, Italy. Tel: �/39 335 6191 852. Fax: �/39 054 330

069. E-mail: [email protected]; [email protected]

The World Journal of Biological Psychiatry, 2005; 6(3): 198�/205

(Received 20 February 2004; accepted 27 September 2004)

ISSN 1562-2975 print/ISSN 1814-1412 online # 2005 Taylor & Francis

DOI: 10.1080/15622970510029858

depressants used alone (i.e., without mood stabilis-

ing agents) could increase agitation in depression

(Altshuler et al. 1995; Koukopoulos and Kouko-

poulos 1999; Baldessarini 2001), the study of AD is

important for clinical practice.

Study aim was to test the diagnostic validity or

utility (Kendell 2003) of AD in bipolar II disorder

(BP-II), by comparing Goodwin and Ghaemi sug-

gested subtypes versus often used diagnostic valida-

tors (Kreapelin 1921; Robins and Guze 1970;

Kendler 1990; Akiskal 2003; Gheami et al. 2002;

Akiskal and Benazzi 2003).

In previous studies (Benazzi 2002; Benazzi et al.

2004), it was shown, in mixed BP-II and major

depressive disorder (MDD) samples, that AD was

much more common in BP-II and that it was

associated with bipolar diagnostic validators, sup-

porting its inclusion into the bipolar spectrum. The

present study, in a much larger sample, focused only

on AD in BP-II. Previous community studies

(Kessler et al. 1994, 1997; Weissman et al. 1988,

1996; Hirschfeld et al. 2003; Jonas et al. 2003)

underdiagnosed BP-II (prevalence around 0.5%),

according to the American Psychiatric Association

literature review on bipolar disorders (2002), mainly

because of strict diagnostic criteria and use of

structured interviews by lay interviewers (American

Psychiatric Association 2002; Ghaemi et al. 2002).

Kessler et al. (1997) stated that ‘methodological

research is needed to develop more accurate mea-

sures of other bipolar symptom profiles for use in

general population epidemiological studies’. Instead,

recent community studies (Angst 1998; Szadoczky

et al. 1998; Judd and Akiskal 2003) and clinical

sample studies (Cassano et al. 1992; Dunner and

Tay 1993; Hantouche et al. 1998; Manning et al.

1999; Akiskal et al. 2000; Angst et al. 2003; Benazzi

2003a) found BP-II to be more common (5�/8% in

community, 30�/50% compared to MDD in clinical

samples), by using advanced probing methods for

hypomania or different criteria for hypomania.

Because of its frequency, the study of BP-II can

have an important impact on clinical practice.

Methods

More details on study methods can be found in

previous reports (Benazzi 2003b,c).

Study setting

An outpatient psychiatry private practice, is more

representative of the mood disorders (apart from

bipolar-I) usually seen in clinical practice in Italy,

because (1) it is the first or second (after family

doctors) line of treatment of mood disorders, (2) the

most severe and socially disadvantaged cases are

usually seen in academic and national health service

(NHS) tertiary-care centers, (3) mood disorder

patients do not like to be treated for depression in

the NHS for fear of stigma, and (4) most individuals

can see a private psychiatrist (fee-for-service), redu-

cing a possible income bias.

Interviewer

A senior clinical and mood disorder research psy-

chiatrist.

Patients

Three hundred and twenty consecutive BP-II out-

patients, presenting voluntarily (off psychopharma-

cotherapy) for treatment of a major depressive

episode (MDE), were included in the last 5 years.

Substance-related and borderline personality disor-

ders were excluded as these confound the diagnosis

of BP-II (Akiskal and Pinto 1999), and are, however,

rare in the study setting (Benazzi 2000). Clinically

significant general medical illnesses and cognitive

disorders were also excluded.

Interview methods

During the assessment visit (off psychoactive drugs

for at least 2 weeks, apart from a few cases on small

doses of benzodiazepines, in order to avoid inclusion

of antidepressant-induced agitation or its suppres-

sion by mood stabilising agents) the following

instruments were used: (1) Structured Clinical

Interview for DSM-IV Axis I Disorders-Clinician

Version (First et al. 1997) (SCID-CV); the question

on racing thoughts was supplemented by Kouko-

poulos and Koukopoulos’ (1999) definition of

crowded thoughts (i.e., mind continuously full of

non-stop thoughts) following Kraepelin’s descrip-

tion (1913) of the grading of the thought disorders of

hypomania. (2) The Global Assessment of Func-

tioning scale (GAF, in the SCID-CV) for index

MDE severity. (3) the Hypomania Interview Guide

(HIG, item score range 0�/4, Williams et al. 1994) to

assess intra-MDE hypomanic symptoms. (4) The

structured Family History Screen (Weissman et al.

2000) for assessing bipolar disorders (type I and type

II) family history in probands’ first-degree relatives.

Often, family members or close friends supplemen-

ted clinical information during interview, increasing

the validity of BP-II diagnosis and family history

(Akiskal et al. 2000; American Psychiatric Associa-

tion 2000).

Systematic interviews about history of hypo-manic

episodes were always conducted after MDE diag-

nosis, before the assessment of study variables, in

Agitated depression in bipolar II disorder 199

order to avoid a possible interviewer bias. The

SCID-CV is partly semi-structured and based on

clinical evaluation (not on simple yes/no answers to

structured questions). Wording of the sentences can

be changed to improve and check understanding.

This is an important advantage versus fully struc-

tured interviews, because it was shown that semi-

structured interviews reduced the false-negatives

(Dunner and Tay 1993; Simpson et al. 2002;

Benazzi 2003d; Benazzi and Akiskal 2003a). The

skip out instruction of the stem question on history

of past mood change was not followed, in order to

assess all past hypomanic symptoms, especially

overactivity (increased goal-directed activity), in

line with previous reports (Dunner and Tay 1993;

Akiskal et al. 2001; Simpson et al. 2002; Angst et al.

2003; Benazzi 2003d; Benazzi and Akiskal 2003a).

This behavioural change is easier to remember than

mood change (always required for the diagnosis, and

easier to remember when overactivity has been

remembered) (Benazzi and Akiskal 2003a).

Definitions

Psychomotor agitation was defined by a HIG

psychomotor agitation score ]/2. This cutoff was

chosen because it scores observable behaviour (as

required by DSM-IV). Agitated depression (AD)

was defined as an MDE with psychomotor agitation.

Mixed depression was defined as an MDE with ]/4

concurrent hypomanic symptoms. Previous studies

had validated (by clinical, family history, and psy-

chometric validation) a dimensional definition of

mixed depression (versus several categorical defini-

tions) defined as an MDE with ]/3 concurrent

hypomanic symptoms (Akiskal and Benazzi 2003;

Benazzi 2002, 2003e; Benazzi and Akiskal 2003b). It

was found that the higher the number of hypomanic

symptoms the higher the bipolar family history

loading (dose�/response relationship). The bipolar

nature of mixed depression was replicated by Sato et

al (2003). In the present study, the cutoff of

hypomanic symptoms for defining mixed depression

had to be higher because almost all BP-II depres-

sions had at least one hypomanic symptom (98.1%),

and there were only six depressions with two

concurrent hypomanic symptoms. Non-mixed de-

pression was defined as an MDE with B/4 con-

current hypomanic symptoms. Mixed-AD was

defined as an MDE with ]/4 hypomanic symptoms

(including psychomotor agitation). Non-mixed-AD

was defined as an MDE with B/4 concurrent

hypomanic symptoms (including psychomotor agi-

tation). Non-AD was defined as an MDE (mixed or

non-mixed) without psychomotor agitation. Mixed-

non-AD was defined as a mixed MDE without

psychomotor agitation.

Testing

To test the diagnostic validity or utility of AD,

Goodwin and Ghaemi suggested subtypes were

compared versus often used bipolar and diagnostic

validators (Kreapelin 1921; Robins and Guze 1970;

Kendler 1990; Coryell 1999; Gheami et al. 2002;

Akiskal 2003; Akiskal and Benazzi 2003; Angst et al.

2003), such as gender, age at onset, number of

recurrences, frequency of atypical depressions, axis I

comorbidity, bipolar (type I and type II) family

history (the most important validator).

Statistics

For pairwise comparisons of continuous variables t-

test was used, and Fisher’s exact test was used for

categorical variables. Univariate logistic regression

was also used to study associations. STATA Statis-

tical Software, Release 7, was used (Stata Corpora-

tion, College Station, TX, USA, 2001). P values

were two-tailed, and a level was set at 0.05, given the

exploratory nature of the study.

Results

AD was present in 35.0% of BP-II depressed

patients (112/320). AD was mixed in 75.8%, while

only 14.3% of non-AD was mixed (P�/0.0000).

Comparisons between AD and non-AD are pre-

sented in Table I. AD had significantly higher index

age, more females, depression recurrences, bipolar I

disorder family history, and more concurrent hypo-

manic symptoms: four to five such symptoms were

present in 74.9% versus 14.3% in non-AD, and

irritability, more talkativeness, and risky behavior

were more common in AD versus non-AD. Com-

parisons of all MDE symptoms were not significantly

different (apart from the 0.0% versus 1.9% psycho-

motor retardation difference, not clinically signifi-

cant), and were not reported in the table.

Comparisons between mixed-AD and non-mixed-

AD are presented in Table II. Mixed-AD had

significantly more of hypomanic symptoms (by

definition), and distractibility, racing thoughts, irrit-

ability, more talkativeness, and risky behaviour were

significantly more common. All the other variables

were not significantly different, but there were

clinically significant differences on gender, comor-

bidity, atypical depressions, and BP-II family history.

Comparisons of all MDE symptoms were not

significantly different (apart from more diminished

200 F. Benazzi

ability to concentrate and less decreased eating in

mixed-AD), and were not reported in the table.

A definition of mixed depression not always

requiring psychomotor agitation was also studied.

Univariate logistic regression of mixed depression

(n�/116) (dependent variable) versus non-mixed

depression (n�/204) found that, among the diag-

nostic validators gender, age at onset, recurrences,

atypical depression, and bipolar I and bipolar II

family history, mixed depression was significantly

associated with female gender (odds ratio�/1.8, 95%

confidence interval�/1.1�/3.1, P�/0.017), atypical

depression (odds ratio�/1.7, 95% confidence inter-

val�/1.1�/2.8, P�/0.018), and bipolar I family his-

tory (odds ratio�/3.7, 95% confidence interval�/

1.0�/2.9, P�/0.044). Psychomotor agitation was

present in 73.2% of mixed depressions.

In order to test the importance of psychomotor

agitation for the definition of mixed depression,

mixed-AD was compared to mixed-non-AD (Table

III). Mixed-AD had significantly higher index age,

more hypomanic symptoms (not a clinically signifi-

cant difference), more melancholic features, and less

frequent risky and goal-directed activities.

Discussion

Findings are in line with Maj et al. (2003) study of

AD versus non-AD in bipolar-I disorder, relative to

the higher frequency of concurrent hypomanic

symptoms, and to the specific hypomanic symptoms

present in AD. Findings are also in line with Perugi

et al. (2001) study of bipolar-I mixed depression,

relative to the specific manic symptoms. Similar

findings in bipolar-I and BP-II disorders, in different

settings (primary care versus tertiary-care, outpati-

ents versus inpatients), assessment methods, and

diagnostic systems, may partly support the subtyp-

ing of AD. A link between bipolar-I disorder and BP-

II AD was supported by the higher bipolar-I family

history in the BP-II AD sample (and also by the

similar, but less severe, hypomanic symptoms of BP-

II and bipolar-I AD).

AD was not uncommon in this outpatient sample,

even if psychomotor changes are more likely in

inpatient depression (American Psychiatric Associa-

tion 2000; Parker 2000). This finding supports the

clinical utility of subtyping AD, as this is a disorder

that a clinician is likely to see every one in three

Table I. Comparisons between agitated (AD) and non-agitated (non-AD) depressions.

AD

(n�/112)

non-AD

(n�/208)

T, Fisher,

df�/318

P

Variables: mean (SD),% (N )

Index age (years) 43.7(12.8) 40.5(13.6) 2.0 0.041

Age at onset first MDE (years) 21.4(10.3) 23.5(10.7) �/1.6 0.090

Index GAF 49.5(9.9) 51.1(8.6) �/1.5 0.133

N hypomanic symptoms 4.1(1.1) 2.3(1.1) 13.9 0.000

Female gender (%) 78.5(88) 62.9(131) 0.005

�/4 MDEs (%) 88.3(99) 77.4(161) 0.017

Index MDE symptoms �/2 years (%) 39.2(44) 41.3(86) 0.811

Axis I comorbidity (%) 57.1(64) 52.4(109) 0.481

Index psychotic features (%) 10.7(12) 6.7(14) 0.282

Index melancholic features (%) 15.1(17) 10.5(22) 0.282

Index atypical features (%) 58.9(66) 52.8(110) 0.346

Bipolar I disorder family history (%) 12.5(14) 3.3(7) 0.001

Bipolar II disorder family history (%) 46.4(52) 45.1(94) 0.906

Intra-MDE hypomanic symptoms:

1 hypomanic symptom (%) 0.8(1) 19.2(40) 0.000

2 hypomanic symptoms (%) 5.3(6) 32.2(67) 0.000

3 hypomanic symptoms (%) 17.8(20) 30.7(64) 0.016

4 hypomanic symptoms (%) 38.3(43) 11.0(23) 0.000

5 hypomanic symptoms (%) 36.6(41) 3.3(7) 0.000

Distractibility (%) 83.0(93) 73.0(152) 0.053

Racing/crowded thoughts (%) 77.6(87) 73.0(152) 0.419

Irritable mood (%) 69.6(78) 55.7(116) 0.017

More talkativeness (%) 49.1(55) 12.9(27) 0.000

Increased risky activities (%) 24.1(27) 14.9(31) 0.048

Increased goal-directed activity (%) 7.1(8) 7.6(16) 1.000

Reduced need for sleep (%) 2.6(3) 1.4(3) 0.426

Increased self-esteem (%) 0.0 0.0 nc nc

Elevated mood (%) 0.0 0.0 nc nc

t -test for means, Fisher’s exact test for frequencies.

GAF, Global Assessment of Functioning scale; MDE, major depressive episode; nc, not calculable.

Agitated depression in bipolar II disorder 201

outpatient BP-II depressions. As BP-II was found to

be common in clinical (30�/50% relative to MDD)

and community (5�/8%) samples, and as nearly one

in two MDD may become bipolar (I or II) during

long follow-ups (Angst et al. in press; Goldberg et al.

2001), AD subtyping seems to be clinically useful.

AD, versus non-AD, was significantly more likely

to have many (]/4) hypomanic symptoms (76% vs.

14%), suggesting that psychomotor agitation often

clusters with other hypomanic symptoms in BP-II

depression. Also, this finding may partly support the

subtyping of AD. In the study sample, and in the

sample by Maj et al. (2003), AD had a clinical

picture similar to Kraepelin’s ‘excited depression’,

which could be seen as another piece of evidence

partly supporting the subtyping of AD.

AD, versus non-AD, had significant differences on

some diagnostic validators (gender, recurrences,

family history), partly supporting the subtyping of

AD.

The comparison between mixed and non-mixed

AD did not produce significant differences, but there

were several clinically significant differences on

diagnostic validators, partly supporting the subtyp-

ing of AD. As the sample of non-mixed-AD was not

large, there is the possibility of a type II error (i.e.,

missing a real difference). So these findings must be

seen as preliminary and requiring replication in

larger samples.

The comparison between mixed-AD and mixed-

non-AD, carried out to test the importance of

psychomotor agitation for the definition of mixed

depression, showed few significant differences (and

no clinically significant difference). No diagnostic

validators (especially family history) were signifi-

cantly different. Among the hypomanic symptoms,

there were significant differences only on risky and

increased-goal directed activities. The sample of

mixed-non-AD was not large, so type II error could

be possible. Findings are therefore preliminary and

need replication in larger samples.

On the basis of the result of the mixed-AD versus

mixed-non-AD comparison, it seems that psycho-

motor agitation may not always be required to define

mixed depression (as suggested by previous studies

supporting a dimensional definition, Benazzi 2002,

2003e; Akiskal and Benazzi 2003; Sato et al. 2003).

However, psychomotor agitation was present in 73%

Table II. Comparisons between mixed (M-AD) and non-mixed (non-M-AD) agitated depressions.

M-AD

(n�/85)

non-M-AD

(n�/27)

T, Fisher,

df�/110

P

Variables: mean (SD), (%) (N )

Index age (years) 43.2(12.6) 45.4(13.8) �/0.7 0.441

Age at onset first MDE (years) 21.7(11.0) 20.5(8.1) 0.5 0.602

Index GAF 49.2(9.9) 50.3(10.1) �/0.5 0.617

N hypomanic symptoms 4.6(0.8) 2.7(0.5) 11.6 0.000

Female gender (%) 81.1(67) 70.3(19) 0.434

�/4 MDEs (%) 89.4(76) 85.1(23) 0.510

Index MDE symptoms �/2 years (%) 37.6(32) 44.4(12) 0.652

Axis I comorbidity (%) 60.0(51) 48.1(13) 0.372

Index psychotic features (%) 11.7(10) 7.4(2) 0.727

Index melancholic features (%) 15.2(13) 14.8(4) 1.000

Index atypical features (%) 63.5(54) 44.4(12) 0.115

Bipolar I disorder family history (%) 12.9(11) 11.1(3) 1.000

Bipolar II disorder family history (%) 50.5(43) 33.3(9) 0.128

Intra-MDE Hypomanic symptoms:

1 hypomanic symptom (%) 0.0(0) 3.7(1) 0.241

2 hypomanic symptoms (%) 0.0(0) 22.2(6) 0.000

3 hypomanic symptoms (%) 0.0(0) 74.0(20) 0.000

4 hypomanic symptoms (%) 50.5(43) 0.0(0) 0.000

5 hypomanic symptoms (%) 48.2(41) 0.0(0) 0.000

Distractibility (%) 90.5(77) 59.2(16) 0.001

Racing/crowded thoughts (%) 87.0(74) 48.1(13) 0.000

Irritable mood (%) 81.1(69) 33.3(9) 0.000

More talkativeness (%) 56.4(48) 25.9(7) 0.008

Increased risky activities (%) 30.5(26) 3.7(1) 0.004

Increased goal-directed activity (%) 9.4(8) 0.0(0) 0.195

Reduced need for sleep (%) 3.5(3) 0.0(0) 1.000

Increased self-esteem (%) 0.0 0.0 nc nc

Elevated mood (%) 0.0 0.0 nc nc

t -test for means, Fisher’s exact test for frequencies.

GAF, Global Assessment of Functioning scale; MDE, major depressive episode; nc, not calculable.

202 F. Benazzi

of mixed depressions; 76% of AD was mixed, and

only 14% of non-AD was mixed. These findings

suggest that psychomotor agitation is an important

hypomanic symptom of mixed depression, even if it

needs not always to be present, in line with a

dimensional definition of mixed depression versus a

categorical one (i.e., always requiring specific symp-

tom/s).

All the above findings may partly support the

subtyping of AD as a more ‘mixed’ variant of BP-II

depression. While validating AD as a distinct sub-

type needs further studies, we have shown that AD

has diagnostic utility (Kendell 2003). The diagnostic

utility of this subtyping is related to its possible

important impact on treatment, as antidepressants

used alone may increase psychomotor agitation,

while mood stabilisers and antipsychotics can be

used to treat agitation and the associated hypomanic

symptoms before using antidepressants (Akiskal and

Pinto 1999; Koukopoulos and Koukopoulos 1999;

Baldessarini 2001; Benazzi et al. 2002a, 2004). A

recent study (Bottlender et al. 2004) has shown, in

bipolar-I disorder, that mixed depression was more

likely to switch compared to non-mixed depression

when treated with antidepressants, and in the pre-

sent study it was shown that AD was often mixed.

However, the effects of antidepressants on bipolar

(type I, and especially type II) depression are far

from clear (Amsterdam et al. 2003, 2004; Goldberg

et al. 2003; Post et al. 2003; Benazzi 2004; Ghaemi

et al. 2004; Hadjipavlou et al. 2004). On this

practical basis, AD might be seen as an important

variant of BP-II depression.

Limitations

A single interviewer limited the validity of the

findings. However, the interviewer’s inter-rater

agreement for BP-II diagnosis had been formally

tested, showing adequate reliability (k�/0.73, Be-

nazzi 2003b). Furthermore, reliability of BP-II

diagnosis was found to be high when trained

clinicians used semi-structured interviews as done

in the present study (Simpson et al. 2002), and

clinicians using semi-structured interviews made

more correct diagnoses of mood disorders compared

to structured interviews (Brugha et al. 2001). The

interview was conducted by a clinician studying and

treating mood disorders for a long time. Systematic

use of validated structured/semi-structured inter-

views (and information from key informants),

should have reduced any systematic bias. An inter-

viewer’s bias is unlikely, as the present study aim was

not in the mind of the interviewer when the variables

were collected for different study aims. Results may

not be applicable to that part of BP-II patients

Table III. Comparisons between mixed agitated (M-AD) and mixed non-agitated (M-non-AD) depressions.

M-AD

(n�/85)

M-non-AD

(n�/31)

T, Fisher,

df�/114

P

Variables: mean (SD), (%) (N )

Index age (years) 43.2(12.6) 30.0(10.3) 2.0 0.041

Age at onset first MDE (years) 21.7(11.0) 21.6(9.8) 0.0 0.964

Index GAF 49.2(9.9) 51.9(7.9) �/1.3 0.174

N hypomanic symptoms 4.6(0.8) 4.2(0.5) 2.6 0.010

Female gender (%) 81.1(67) 64.5(20) 0.147

�/4 MDEs (%) 89.4(76) 77.4(27) 0.744

Index MDE symptoms �/2 years (%) 37.6(32) 32.2(10) 0.666

Axis I comorbidity (%) 60.0(51) 54.8(17) 0.673

Index psychotic features (%) 11.7(10) 3.2(1) 0.284

Index melancholic features (%) 15.2(13) 0.0(0) 0.019

Index atypical features (%) 63.5(54) 64.5(20) 1.000

Bipolar I disorder family history (%) 12.9(11) 9.6(3) 0.757

Bipolar II disorder family history (%) 50.5(43) 41.9(13) 0.529

Intra-MDE Hypomanic symptoms:

Distractibility (%) 90.5(77) 93.5(29) 1.000

Racing/crowded thoughts (%) 87.0(74) 96.7(30) 0.177

Irritabile mood (%) 81.1(69) 87.0(27) 0.584

More talkativeness (%) 56.4(48) 45.1(14) 0.300

Increased risky activities (%) 30.5(26) 58.0(18) 0.009

Increased goal-directed activity (%) 9.4(8) 38.7(12) 0.001

Reduced need for sleep (%) 3.5(3) 9.6(3) 0.340

Increased self-esteem (%) 0.0 nc nc

Elevated mood (%) 0.0 nc nc

t -test for means, Fisher’s exact test for frequencies.

GAF, Global Assessment of Functioning scale; MDE, major depressive episode; nc, not calculable.

Agitated depression in bipolar II disorder 203

showing a high substance-related disorder comor-

bidity (Sonne and Brady 1999), called ‘dark’ BP-II

(versus the ‘sunny’ BP-II) by Akiskal et al. (2003).

This subgroup is much more likely to be seen in

tertiary care settings, which usually see the most

severe patients. Instead, in non-tertiary-care set-

tings, as in the present study, private practice, this

subgroup of BP-II (with its related personality

disorders) was found to be uncommon (Benazzi

2000).

The present study findings must be seen as

preliminary, and needing replication in different

settings and by independent groups.

Statement of interest

The author has no conflict of interest with any

commercial or other associations in connection with

the submitted article.

References

Akiskal HS. 2003. Validating ‘hard’ and ‘soft’ phenotypes within

the bipolar spectrum: continuity or discontinuity? J Affect

Disord 73:1�/5.

Akiskal HS, Benazzi F. 2003. Family history validation of the

bipolar nature of depressive mixed states. J Affect Disord

73:113�/122.

Akiskal HS, Pinto O. 1999. The evolving bipolar spectrum:

Prototypes I, II, III, and IV. Psychiatr Clin North Am 22:

517�/534.

Akiskal HS, Bourgeois ML, Angst J, Post R, Moller H-J,

Hirschfeld R. 2000. Re-evaluating the prevalence and diag-

nostic composition within the broad clinical spectrum of

bipolar disorders. J Affect Disord 59(Suppl 1):S5�/30.

Akiskal HS, Hantouche EG, Bourgeois ML, Azorin JM, Sechter

D, Allilaire JF, Chatenet-Duchene L, Lancrenon S. 2001.

Toward a refined phenomenology of mania: combining clin-

ician-assessment and self-report in the French EPIMAN study.

J Affect Disord 67:89�/96.

Akiskal HS, Hantouche EG, Allilaire JF. 2003. Bipolar II with and

without cyclothymic temperament: ‘dark’ and ‘sunny’ expres-

sions of soft bipolarity. J Affect Disord 73:49�/57.

Altshuler LL, Post RM, Leverich GS, Mikalauskas K, Rosoff A,

Ackerman L. 1995. Antidepressant-induced mania and cycle

acceleration: a controversy revisited. Am J Psychiatry 152:

1130�/1138.

American Psychiatric Association. 2000. Diagnostic and Statis-

tical Manual of Mental Disorders, 4th ed, Text Revision

(DSM-IV-TR). Washington, DC: American Psychiatric Asso-

ciation.

American Psychiatric Association. 2002. Practice guideline for the

treatment of patients with bipolar disorder (revision). Am J

Psychiatry 159(Suppl):1�/50.

Amsterdam JD, Brunswick DJ. 2003. Antidepressant monother-

apy for bipolar type II major depression. Bipolar Disord 5:388�/

395.

Amsterdam JD, Shults J, Brunswink DJ, Hundert M. 2004.

Short-term fluoxetine monotherapy for bipolar type II or

bipolar NOS major depression �/ Low manic switch rate.

Bipolar Disord 6:75�/81.

Angst J. 1998. The emerging epidemiology of hypomania and

bipolar II disorder. J Affect Disord 50:143�/151.

Angst J, Gamma A, Benazzi F, Ajdacic V, Eich D, Rossler W.

2003. Toward a re-definition of subthreshold bipolarity:

Epidemiology and proposed criteria for bipolar-II, minor

bipolar disorders and hypomania. J Affect Disord 73:133�/146.

Angst J, Sellaro R, Stassen HH, Gamma Alex. Diagnostic

conversion from depression to bipolar disorders: Results of a

long-term prospective study of hospital admissions. J Affect

Disord (in press).

Baldessarini RJ. 2001. Drugs and the treatment of psychiatric

disorders. Depression and anxiety. In: Hardman JG, Limbird

LE, Goodman Gilman A, editors. Goodman & Gilman’s The

pharmacological basis of therapeutics. 10th ed. New York:

McGraw-Hill. pp 447�/483.

Benazzi F. 2000. Borderline personality disorder and bipolar II

disorder in private practice depressed outpatients. Comp

Psychiatry 41:106�/110.

Benazzi F. 2002. Which could be a clinically useful definition of

depressive mixed state? Prog Neuropsychopharmacol Biol

Psychiatry 26:1105�/1111.

Benazzi F. 2003a. Frequency of bipolar spectrum in 111 private

practice depression outpatients. Eur Arch Psychiatry Clin

Neurosci 253:203�/208.

Benazzi F. 2003b. Depression with racing thoughts. Psychiatry

Res 120:273�/282.

Benazzi F. 2003c. Depressive mixed state: dimensional versus

categorical definitions. Prog Neuropsychopharmacol Biol Psy-

chiatry 27:129�/134.

Benazzi F. 2003d. Diagnosis of bipolar II disorder: a comparison

of structured versus semistructured interviews. Prog Neurop-

sychopharmacol Biol Psychiatry 27:985�/991.

Benazzi F. 2003e. Bipolar II depressive mixed state: Finding a

useful definition. Compr Psychiatry 44:21�/27.

Benazzi F. 2004. How to treat bipolar-II depression and bipolar-II

mixed depression? Int J Neuropsychopharmacol 7:105�/106.

Benazzi F, Akiskal HS. 2003a. Refining the evaluation of bipolar

II: Beyond the strict SCID-CV guidelines for hypomania. J

Affect Disord 73:33�/38.

Benazzi F, Akiskal HS. 2003b. Clinical and factor analytic-

validation of depressive mixed states: A report from the

Ravenna-San Diego collaboration. Curr Opin Psychiatry

16(Suppl 2):S71�/78.

Benazzi F, Helmi S, Bland L. 2002. Agitated depression:

Unipolar? bipolar? or both? Ann Clin Psychiatry 14:97�/104.

Benazzi F, Koukopoulos A, Akiskal HS. 2004. Toward a valida-

tion of a new definition of agitated depression as a bipolar

mixed state (mixed depression). Eur Psychiatry 19:85�/90.

Bottlender R, Sato T, Kleindienst N, Strausz A, Moller H-J. 2004.

Mixed depressive features predict maniform switch during

treatment of depression in bipolar I disorder. J Affect Disord

78:149�/152.

Brugha TS, Jenkins R, Taub N, Meltzer H, Bebbington PE. 2001.

A general population comparison of the Composite Interna-

tional Diagnostic Interview (CIDI) and the Schedules for

Clinical Assessment in Neuropsychiatry (SCAN). Psychol

Med 31:1001�/1013.

Cassano GB, Akiskal HS, Savino M, Musetti L, Perugi G, Soriani

A. 1992. Proposed subtypes of bipolar II and related disorders:

With hypomanic episodes (or cyclothymia) and with hyperthy-

mic temperament. J Affect Disord 26:127�/140.

Coryell W. 1999. Bipolar II disorder: the importance of hypoma-

nia. In: Goldberg JF, Harrow M, editors. Bipolar disorders.

Clinical course and outcome. Washington, DC: American

Psychiatric Press. pp 219�/236.

Dunner DL, Tay KL. 1993. Diagnostic reliability of the history of

hypomania in bipolar II patients and patients with major

depression. Comp Psychiatry 34:303�/307.

204 F. Benazzi

First MB, Spitzer RL, Gibbon M, Williams JBW. 1997. Struc-

tured clinical interview for DSM-IV axis I disorders-clinician

version (SCID-CV). Washington, DC: American Psychiatric

Press.

Ghaemi SN, Ko JY, Goodwin FK. 2002. ‘Cade’s disease’ and

beyond: misdiagnosis, antidepressant use, and a proposed

definition for bipolar spectrum disorder. Can J Psychiatry

47:125�/134.

Ghaemi SN, Hsu DJ, Soldani F, Goodwin FK. 2003. Antide-

pressants in bipolar disorder: The case for caution. Bipolar

Disord 5:421�/433.

Ghaemi SN, Rosenquist KL, Ko JY, Baldassano CF, Kontos NJ,

Baldessarini RJ. 2004. Antidepressant treatment in bipolar

versus unipolar depression. Am J Psychiatry 161:163�/165.

Goldberg JF, Truman CJ. 2003. Antidepressant-induced mania:

An overview of current controversies. Bipolar Disord 5:407�/

420.

Goldberg JF, Harrow M, Whiteside JE. 2001. Risk for bipolar

illness in patients initially hospitalized for unipolar depression.

Am J Psychiatry 158:1265�/1270.

Goodwin FK, Ghaemi SN. 2003. The course of bipolar disorder

and the nature of agitated depression. Am J Psychiatry 160:

2077�/2079.

Hadjipavlou G, Mok H, Yathan LN. 2004. Pharmacotherapy of

bipolar II disorder: A critical review of current evidence.

Bipolar Disord 6:14�/25.

Hantouche EG, Akiskal HS, Lencrenon S, Allilaire J-F, Sechter

D, Azorin J-M, Bourgeois M, Fraud J-P, Chatenet-Duchene L.

1998. Systematic clinical methodology for validating bipolar-II

disorder: data in mid-stream from a french national multi-site

study (EPIDEP). J Affect Disord 50:163�/173.

Hirschfeld RM, Calabrese JR, Weissman MM, Reed M, Davies

MA, Frye MA, Keck PE Jr, Lewis L, McElroy SL, McNulty JP,

Wagner KD. 2003. Screening for bipolar disorder in the

community. J Clin Psychiatry 64:1130�/1131.

Jonas BS, Brody D, Roper M, Narrow WE. 2003. Prevalence of

mood disorders in a national sample of young American adults.

Soc Psychiatry Psychiatr Epidemiol 38:618�/624.

Judd LL, Akiskal HS. 2003. The prevalence and disability of

bipolar spectrum disorders in the US population: Re-analysis of

the ECA database taking into account subthreshold cases. J

Affect Disord 73:123�/131.

Kendell R, Jablensky A. 2003. Distinguishing between the validity

and utility of psychiatric diagnoses. Am J Psychiatry 160:4�/12.

Kendler KS. 1990. Toward a scientific psychiatric nosology.

Strengths and limitations. Arch Gen Psychiatry 47:969�/973.

Kessler RC, McGonagle KA, Zhao S, Nelson CB, Hughes M,

Eshleman S, Wittchen HU, Kendler KS. 1994. Lifetime and

12-month prevalence of DSM-III-R psychiatric disorders in the

United States. Results from the National Comorbidity Survey.

Arch Gen Psychiatry 51:8�/19.

Kessler RC, Rubinow DR, Holmes C, Abelson JM, Zhao S. 1997.

The epidemiology of DSM-III-R bipolar I disorder in a general

population survey. Psychol Med 27:1079�/1089.

Koukopoulos A. 2003. Ewald Hecker’s description of cyclothymia

as a cyclical mood disorder: Its relevance to the modern

concept of bipolar II. J Affect Disord 73:199�/205.

Koukopoulos A, Koukopoulos A. 1999. Agitated depression as a

mixed state and the problem of melancholia. Psychiatr Clin

North Am 22:547�/564.

Kraepelin E. 1921. Manic-depressive insanity and paranoia.

Edinburgh: Livingstone E&S.

Maj M, Pirozzi R, Magliano L, Bartoli L. 2003. Agitated

depression in bipolar I disorder: prevalence, phenomenology,

and outcome. Am J Psychiatry 160:2134�/2140.

Manning JS, Haykal RF, Akiskal HS. 1999. The role of bipolarity

in depression in the family practice setting. Psychiatr Clin

North Am 22:689�/703.

Marneros A. 2001. Origin and development of concepts of bipolar

mixed states. J Affect Disord 67:229�/240.

Parker G. 2000. Classifying depression: Should paradigms lost be

regained? Am J Psychiatry 157:1195�/1203.

Perugi G, Akiskal HS, Micheli C, Toni C, Madaro D. 2001.

Clinical characterization of depressive mixed state in bipolar-I

patients: Pisa-San Diego collaboration. J Affect Disord 67:

105�/114.

Post RM, Leverich GS, Nolen WA, Kupka RW, Altshuler LL,

Frye MA, Suppes T, McElroy S, Keck P, Grunze H, Walden J.

2003. A re-evaluation of the role of antidepressants in the

treatment of bipolar depression: data from the Stanley Founda-

tion Bipolar Network. Bipolar Disord 5:396�/406.

Robins E, Guze SB. 1970. Establishment of diagnostic validity in

psychiatric illness: Its application to schizophrenia. Am J

Psychiatry 126:983�/987.

Sato T, Bottlender R, Schroter A, Moller H-J. 2003. Frequency of

manic symptoms during a depressive episode and unipolar

‘depressive mixed state’ as bipolar spectrum. Acta Psychiatr

Scand 107:268�/274.

Schatzberg AF, DeBattista C. 1999. Phenomenology and treat-

ment of agitation. J Clin Psychiatry 60(Suppl 15):17�/20.

Simpson SG, McMahon FJ, McInnis MG, MacKinnon DF,

Edwin D, Folstein SE, DePaulo JR. 2002. Diagnostic reliability

of bipolar II diagnosis. Arch Gen Psychiatry 59:736�/740.

Sonne SC, Brady KT. 1999. Substance abuse and bipolar

comorbidity. Psychiatr Clin North Am 22:609�/627.

Swann AC, Secunda SK, Katz MM, Croughan J, Bowden CL,

Koslow SH, Berman N, Stokes PE. 1993. Specificity of mixed

affective states: Clinical comparison of dysphoric mania and

agitated depression. J Affect Disord 28:81�/89.

Swann AC, Stokes PE, Secunda SK, Maas JW, Bowden CL,

Berman N, Koslow SH. 1994. Depressive mania versus

agitated depression: Biogenic amine and hypothalamic-pitui-

tary-adrenocortical function. Biol Psychiatry 35:803�/813.

Szadoczky E, Papp Z, Vitrai J, Rihmer Z, Furedi J. 1998. The

prevalence of major depressive and bipolar disorders in

Hungary. Results from a national epidemiologic survey. J Affect

Disord 50:153�/162.

Weissman MM, Leaf PJ, Tischler GL, Blazer DG, Karno M,

Bruce ML, Florio LP. 1988. Affective disorders in five United

States communities. Psychol Med 18:141�/153.

Weissman MM, Bland RC, Canino GJ, Faravelli C, Greenwald S,

Hwu HG, Joyce PR, Karam EG, Lee CK, Lellouch J, Lepine

JP, Newman SC, Rubio-Stipec M, Wells JE, Wickramaratne PJ,

Wittchen H, Yeh EK. 1996. Cross-national epidemiology of

major depression and bipolar disorder. J Am Med Assoc 276:

293�/299.

Weissman MM, Wickramaratne P, Adams P, Wolk S, Verdeli H,

Olfson M. 2000. Brief screening for family psychiatric history.

The family history screen. Arch Gen Psychiatry 57:675�/682.

Williams JBW, Terman M, Link MJ, Amira L, Rosenthal NE.

1994. Hypomania interview guide (including hyperthymia).

Current assessment version (HIGH-C). Clinical Assessment

Tools Packet. Norwood: Center for Environmental Therapeu-

tics.

World Health Organization. 1992. The ICD-10 Classification of

Mental and Behavioural Disorders. Geneva: World Health

Organization.

Agitated depression in bipolar II disorder 205

The World Journal of Biological Psychiatry

Instructions to Authors

The aim of The World Journal of Biological Psychiatry is to increase theworldwide communication of knowledge in clinical and basic research onbiological psychiatry. The composition of The World Journal of BiologicalPsychiatry , with its diverse categories that allow communication of a greatvariety of information, ensures that it is of interest to a wide range ofreaders. It offers the opportunity to educate (through critical reviewpapers), to publish original work and observations (original papers andcase reports), and to express personal opinions (Viewpoints/Letters to theEditor). The World Journal of Biological Psychiatry is thus an extremelyimportant medium in the field of biological psychiatry all over the world.

Manuscripts must be written in standard and grammatical English andshould present new results as well as be of scientific value. Contributionswill be considered for the following categories:

/� Reviews/Mini-reviews/� Original investigations/Summaries of original research/� Brief reports/� Viewpoints/� Case reports/Case series/� Letters to the editors/� Book reviews/� Invitations/Announcements

Please submit manuscripts by e-mail. Alternatively, manuscripts can besent on disk (Microsoft Word files are preferred) accompanied by threepaper copies. Please ensure that all pages are clearly typed, and use doublespacing throughout the manuscript. All manuscript pages should benumbered. Use a clear system of headings to divide up and clarify thetext, with not more than three grades of headings. The desired position offigures and tables should be indicated in the manuscript.

Authors must disclose in their cover letter any commercial or financialarrangements that they have with companies who either have productsfeatured in the manuscript or with companies that produce competingproducts. Any special circumstance, vested interests or sources of bias thatmight affect the integrity of the reported information should also bedisclosed in the cover letter.

Submission of a manuscript implies: that the work described has not beenpublished before (except in the form of an abstract or as part of a publishedlecture, review or thesis); that it is not under consideration for publicationanywhere else; that its publication has been approved by all co-authors, ifany, as well as by responsible authorities �/ tacitly or explicitly �/ at theinstitute where the work has been carried out.

Manuscripts submitted for publication must contain a statement to theeffect that all human studies have been approved by the appropriate ethicscommittee and have therefore been performed in accordance with theethical standards laid down in the 1964 Declaration of Helsinki. It shouldalso be stated clearly in the text that all persons gave their informed consentprior to their inclusion in the study. Details that might disclose the identityof the subjects under study should be omitted.

Reports of animal experiments must state that the Principles of laboratoryanimal care (NIH publication No. 86-23, revised 1985) were followed, aswell as specific national laws (e.g., the current version of the German Lawon the Protection of Animals) where applicable.

The Editorial Office will acknowledge receipt of the manuscript andprovide it with a manuscript reference number. Each manuscript will beassigned to at least two reviewers. The reference number of the manuscriptshould be quoted in all correspondence with the Chief Editor and EditorialOffice.

The editors reserve the right to reject manuscripts that do not comply withthe above-mentioned requirements. The author will be held responsible forfalse statements or for failure to fulfill the above-mentioned requirements.

Manuscript specifications1. The title page shall contain: The title of the article, which should be

concise but informative. A short title not exceeding 40 letters andspaces. The full names, affiliations and addresses of all the contribu-tors, and e-mail and fax number for the corresponding author.

The title page should also include the number of words in text/abstract/title, the number of references and the number of figures and tables.

2. Each article must include a short Abstract of the essential results (notexceeding 200 words).

3. Key words. Up to five key words should be inserted below the abstract.Use terms from the Medical Subject Headings list from Index Medicus.

4. References. All references cited in the text are to be listed (double-spaced) at the end of the paper in alphabetical order under the name ofthe first author. If there are more than six authors, list the first sixauthors and use ‘et al.’ Articles ‘in press’ may be included, but muststate the journal that has accepted them. Personal communicationsshould be avoided. Journal titles should be abbreviated according toIndex Medicus.

Citations in the text should be by author and date in parentheses. Ifthere are two authors, both should be named (e.g., Agar and Douglas1955); if an article with more than two authors is cited, only the firstauthor’s name plus ‘et al.’ need be given (e.g., Komor et al. 1979); ifthere is more than one reference by the same author or team of authorsin the same year, then a, b, c, etc., should be added to the year, both inthe text and the list of references. When styling references, thefollowing examples should be followed:

Brain WR. 1958. The physiological basis of consciousness. Acritical review. Brain 81:426 �/455.

Kuhlenbeck H. 1954. The human diencephalon. A summary ofdevelopment, structure, function and pathology. Basel: Karger.

Teuber HL. 1964. The riddle of frontal lobe function in man. In:Warren T, Akert CH, editors. The frontal and granular cortex andbehavior. New York: McGraw-Hill. p 252�/271.

5. Tables should be typed on separate sheets and numbered consecutivelywith Arabic numbers. All tables should include a brief descriptivecaption.

6. Figures. All illustrations, whether photographs, graphs or diagrams,should be numbered consecutively with Arabic numerals and sub-mitted on separate sheets. Each illustration should be referred to in thetext as ‘Figure 1’, Figures 1�/4’, etc. The figures should have a shorttitle followed by a concise description. All abbreviations and symbolsappearing in the figure must be explained in the legend. Remarks suchas ‘for explanation (or details) see text’ should be avoided. Each figureshould have a legend containing sufficient information to render thefigure intelligible without reference to the text. If a figure has beenpublished previously, acknowledge the original source and submitwritten permission from the copyright holder to reproduce it. Ifphotographs of persons are used, either the subjects must not beidentifiable or their written permission to use the photographs mustaccompany the manuscript. Figures can be sent electronically asseparate files in EPS or TIFF format.

7. Financial disclosure. A copy of the Author Disclosure Declaration (seenext page) should be completed, signed by all authors and submitted.

8. Copyright. It is a condition of publication that authors vest copyright intheir articles, including abstracts, in Taylor & Francis Group Ltd. Thisenables us to ensure full copyright protection and to disseminate thearticle, and the journal, to the widest possible readership in print andelectronic formats as appropriate. Authors may, of course, use thematerial elsewhere after publication providing that prior permission isobtained from Taylor & Francis Group Ltd. Authors are themselvesresponsible for obtaining permission to reproduce copyright materialfrom other sources. To view the ‘Copyright Transfer Frequently AskedQuestions’ please visit http://www.tandf.co.ukljoumalslcopyright.asp.

9. Proofs and offprints. Proofs will be sent as a PDF file to the correspondingauthor’s e-mail address unless otherwise requested. The corrected proofmust be returned to the publisher within 3 days. Authors will be liable topay for excessive alterations in the proof. Offprints can be ordered byfilling out the form accompanying the proofs.The corresponding authors will be supplied with a free copy of therelevant issue.

10. All manuscripts should be submitted to the Editorial Office at thefollowing address:

Dorothea Bode, Editorial Administrator, Department of Psychiatry,Ludwig-Maximilians-University, Nussbaumstrasse 7, 80336 Munich,Germany. Tel: �/49 89 5160 5505. Fax: �/49 89 5160 5530. E-mail:[email protected]

Manuscripts and disks will not be returned.

The World Journal of Biological Psychiatry

Author Disclosure Declaration

It is the policy of the World Federation of Societies of Biological Psychiatry (WFSBP) to ensure balance,independence, objectivity and scientific rigor in The World Journal of Biological Psychiatry.

All authors are expected to disclose to the audience any real or apparent conflict(s) of interest that may have adirect bearing on the subject matter of the article. This includes relationships with pharmaceutical companies,biomedical device manufacturers, or other corporations whose products or services are related to the subjectmatter of the article.

The intent of this policy is not to prevent authors with a potential conflict of interest from publication. TheFederation does not view the existence of these interests or uses as implying bias or decreasing value to thereadership.The Federation feels that this disclosure is important to identify any potential conflict openly, with the intentionthat the readers may form their own judgments about each article with the full disclosure of the facts.It remains for the readers to determine whether the author’s outside interests may reflect a possible partiality ineither the exposition or the conclusions presented.

Please fill in and submit together with your manuscript to the Editorial Office:

Dorothea BodeEditorial AdministratorDepartment of Psychiatry, Ludwig �/ Maximilians �/ UniversityNussbaumstrasse 7D-80336 Munich, Germany

Title of the Article:

Author/Co-Author Names:

I I do not have a financial interest/arrangement or affiliation with any organizations that could beperceived as a real or apparent conflict of interest in the context of the subject of my article(s).

I I have a financial interest/arrangement or affiliation with one or more organization(s) that could beperceived as a real or apparent conflict of interest in the context of the subject of my article(s). I amfully aware that I must disclose to the readers any real or apparent conflict(s) of interest that may havea direct bearing on the subject matter of my articles.

Please tick the box(es) which correspond to your situation, and provide details if needed:

I Ownership of stock options or other financial instruments of companies whose products are mentioned inthe article or who manufacture competing products

I Ongoing paid consultancy with company or competitor (actual or within last 2 years)I Employment with company or competitor (actual or within last 2 years)I Honorarium or compensation for conducting research related to the material contained in the articleI Research and/or educational grant

I verify that the information above is accurate. I acknowledge that my article(s) must give a balanced viewof the therapeutic options; that the use of generic names of products contributes to impartiality; and thatif trade names are mentioned, those of several companies should be used.

Signatures of Author/Co-Authors Date